You are on page 1of 1120

Buckling Experiments:

Experimental Methods in
Buckling of Thin-Walled
Structures
Shells, Built-up Structures, Composites
and Additional Topics - Volume 2
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Buckling Experiments: Experimental Methods in Buckling of Thin-Walled Structures: Shells, Built-Up Structures, Composites
and Additional Topics – Volume 2. J. Singer, J. Arbocz and T. Weller Copyright © 2002 John Wiley & Sons, Inc.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Buckling Experiments:
Experimental Methods in
Buckling of Thin-Walled
Structures
Shells, Built-up Structures, Composites
and Additional Topics - Volume 2

J. Singer
Technion-Israel Institute of Technology, Haifa, Israel
J. Arbocz
Delft University of Technology, The Netherlands
T. Weller
Technion-Israel Institute of Technology, Haifa, Israel

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

JOHN WILEY & SONS, INC.


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
This book is printed on acid-free paper. (8)

Copyright © 2002 by John Wiley & Sons, Inc., New York. All rights reserved.

Published simultaneously in Canada.

No part of this publication may be reproduced, stored in a retrieval system or transmitted in any
form or by any means, electronic, mechanical, photocopying, recording, scanning or otherwise,
except as permitted under Sections 107 or 108 of the 1976 United States Copyright Act, without
either the prior written permission of the Publisher, or authorization through payment of the
appropriate per-copy fee to the Copyright Clearance Center, 222 Rosewood Drive, Danvers, MA
01923, (978) 750-8400, fax (978) 750-4744. Requests to the Publisher for permission should be
addressed to the Permissions Department, John Wiley & Sons, Inc., 605 Third Avenue, New York,
NY 10158-0012, (212) 850-6011, fax (212) 850-6008, E-Mail: PERMREQ@WILEY.COM.

This publication is designed to provide accurate and authoritative information in regard to the
subject matter covered. It is sold with the understanding that the publisher is not engaged in
rendering professional services. If professional advice or other expert assistance is required, the
services of a competent professional person should be sought.

Wiley also publishes its books in a variety of electronic formats. Some content that appears in print
may not be available in electronic books. For more information about Wiley products, visit
our web site at www.wiley.com.

Library of Congress Cataloging-in-Publication Data

ISBN 0-471-97450-1

Printed in the United States of America.

10 9 8 7 6 5 4 3 2 1

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Contents

Preface to Volume 1 xiii


Preface to Volume 2 xvii
Abbreviated Contents of Vol. 1: Basic Concepts, Columns,
Beams and Plates xix

9 Shell Buckling Experiments 623


9.1 Introduction 623
9.1.1 Historical Background - Shells under External Pressure 623
9.1.2 Historical Background - Axially Compressed Shells 628

9.2 Buckling and Postbuckling Behavior of Axially Compressed


Cylindrical Shells 631
9.2.1 Sequence of Events in an Axial Compression Experiment 631
9.2.2 Influence of Rigidity of Test Machine 640

9.3 Model Fabrication for Isotropic Shells 641


9.3.1 Electroforming 641
9.3.2 Mylar Shell Specimens 644
9.3.3 Thermal Vacuum Forming 646
9.3.4 Spin-Casting and Other Thermoforming Processes for
Plastic Models 648
9.3.5 Cold-Worked and Machined Metal Shells 652
9.3.6 Seamless Commercial Drink Cans 654
9.3.7 Realistically Fabricated Shells 654

9.4 Test Setups for Cylindrical Shells under Axial Compression 655
9.4.1 Typical Experiments of the Fifties and Sixties 655
9.4.2 Tohuko University Test Setup for Postbuckling Studies 659
9.4.3 Stanford University High-Precision Test Rig 664
9.4.4 Typical Modern Test Systems for Cylindrical Shells 666

9.5 Recording of Buckling and Postbuckling Behavior 674


9.5.1 Determination of Onset of Buckling 674
9.5.2 Buckling Behavior of Oval Cylindrical Shells 675
9.5.3 High-Speed Photography 677
9.5.4 Strain Gages for Detection of Incipient Buckling 678
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
vi Contents

9.6 Southwell's Method for Shells 679


9.6.1 Application of the Method in Case of External Pressure 680
9.6.2 Stanford University and Georgia Tech Comprehensive
Studies 684
9.6.3 Application to Spherical Shells 689
9.6.4 Application to Stiffened Cylindrical and Conical Shells 690
9.6.5 On the Applicability of Southwell's Method of Shells 693
9.7 Cylindrical Shells under External Pressure, Bending or Torsion 695
9.7.1 External Pressure Loading 695
9.7.2 Bending 702
9.7.3 Torsion 710
9.8 Combined Loading 715
9.8.1 Buckling of Pressure Stabilized Shells 715
9.8.2 Combined Loading Test Setups 717
9.8.3 Repeated Buckling Approach 719
9.9 Conical Shells 722
9.9.1 Conicity Effects and Definitions 722
9.9.2 Early Buckling Experiments of Conical Shells 723
9.9.3 Technion Experimental Program 726
9.9.4 Stanford University Experiments 733
9.9.5 Ghent University Tests on Liquid-Filled Cones 735
9.10 Spherical Shells 738
9.10.1 Spherical Caps 738
9.10.2 More Recent Experiments on Spherical Caps-Effect of
Boundary Conditions 741
9.10.3 Complete Spherical Shells 746
9.1 0.4 Large Fabricated Spherical Shells 753
9.10.5 Spherical Shells Subjected to Concentrated Loads 761
9.11 Toroidal Shells, Torispherical Shells, Buckling nuder Internal
Pressure 762
9.11.1 Toroidal Shells 762
9.11.2 Torispherical Shells under External Pressure 768
9.11.3 Buckling under Internal Pressure 777
9.12 Shells Subjected to Transverse Shear Loads 787

10 Initial Imperfections 809


10.1 Introduction 809
10.2 Early Incomplete Imperfection Surveys 810
10.3 Early Complete Imperfection Surveys 815
10.4 The Awakening of Imperfection Measurement Awareness 820
10.5 Complete Imperfection Surveys on Large or Full-Scale
Cylindrical Shells 820
10.6 Imperfection Surveys on Large Shells of Revolution 825
10.7 Recent Laboratory Scale Imperfection Measurement Systems 828
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
10.8 Evaluation of Imperfection Data 831
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Contents vii

10.9 Characteristic Initial Imperfection Distributions 836


10.9.1 Laboratory Scale Shells 836
10.9.2 Full Scale Shells-Riveted Seams 837
10.9.3 Full Scale Shells-Welded Seams 838

10.10 Imperfection Data Banks 840


10.11 Probabilistic Design Methods 841
10.11.1 Closed-Form Solution 841
10.11.2 Monte Carlo Method 843
10.11.3 Response Surface Method 845

10.12 Residual Stresses 847


10.13 Imperfection Measurements and Data Banks in Columns and
Plates 852
10.14 Concluding Remarks 856

11 Boundary Conditions and Loading Conditions 863


11.1 Column Buckling 863
11.2 Plate Buckling 865
11.2.1 Column Behavior 866
11.2.2 Flange Behavior 867
11.2.3 Plate Behavior 868
11.2.4 Elastically Supported Unloaded Edges 869
11.2.5 Effect of Boundary Conditions for Shear Loading 871
11.2.6 Experimental Verification 873
11.2.7 Effect of Boundary Conditions on Postbuckling
Behavior 875

11.3 Buckling of Circular Cylindrical Shells 875


11.3.1 Effect of Boundary Conditions Using Membrane
Prebuckling 876
11.3.2 Effect of Boundary Conditions Using Rigorous
Prebuckling 884
11.3.3 Effect of Elastic Boundary Conditions Using Rigorous
Prebuckling 886
11.3.4 Load Eccentricity Effects 897

11.4 Concluding Remarks 899

12 Stiffened Plates 905


12.1 Built-up Structures, Local and General Instability 905
12.2 Buckling and Postbuckling Strength of Stiffened Plates 905
12.2.1 Introduction 905
12.2.2 Analysis of Stiffened Plates 906
12.2.3 Mode Interaction in Stiffened Panels 906

12.3 Experiments on Stiffened Plates Subjected to Axial


Compression 907
12.3.1 Local Buckling Tests 907
--`,`,`````,`,````,,``,``,,`-`-`,,`

12.3.2 Early Small-Scale Tests 907


12.3.3 University College, London, Small-Scale Tests 910
Copyright Wiley 12.3.4 Manchester University Experiments 915
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
viii Contents

12.3.5 Welded Steel Ship Grillage Tests 921


12.3.6 Large-Scale Civil Engineering Tests 931
12.3.7 Single-Stiffener Panels 936
12.3.8 Aerospace Tests 939

12.4 Sandwich Plates 943


12.4.1 Sandwich Structures 943
12.4.2 Buckling Tests of Sandwich Plates 944

13 Stiffened Shells 955


13.1 Global and Local Buckling of Stiffened Shells 955
13.1.1 Introduction 955
13.1.2 Closely Stiffened Shells 955
13.1.3 Linear Smeared Stiffener Theory 956
13.1.4 Eccentricity and Discreteness Effects 959
13.1.5 Boundary Effects 961
13.1.6 Adequacy and Bounds of Validity of Linear
Smeared Stiffener Theory 964

13.2 Model Fabrication for Stiffened Shells 966


13.2.1 Machined Shells 966
13.2.2 Small-Scale Welded Specimens 97 4
13.2.3 Realistically Fabricated Small-Scale Stiffened Shells 979
13.2.4 Plastic Models 983
13.2.5 Large Stiffened Shells 985

13.3 Experiments on Stiffened Cylindrical Shells Subject to Axial


Compression 991
13.3.1 Technion Axial Compression Tests 991
13.3.2 Axial Compression Tests on Small Welded Shells 994
13.3.3 Testing of Large Shells Subject to Axial Compression 996

13.4 Experiments on Stiffened Cylindrical Shells under External


Pressure, Bending and Torsion 999
13.4.1 External Pressure Tests 999
13.4.2 Bending Tests 1006
13.4.3 Torsion Tests and Combined Loading 1009

13.5 Stiffened Conical and Spherical Shells 1013


13.5.1 Stiffened Conical Shells 1013
13.5.2 Technion Tests on Stiffened Conical Shells 1013
13.5.3 More Recent Tests on Stiffened Conical Shells 1017
13.5.4 Stiffened Spherical Shells 1020

13.6 Experiments on Stiffened Curved Panels 1022


13.6.1 Curved Panel Experiments 1022
13.6.2 Early Lockheed-UC Berkeley Tests 1025
a. Deflectometers 1025
b. Load distribution 1026
c. Scale factor 1027
d. Retest technique 1027
e. Experimental method 1027
13.6.3 Lockheed Palo Alto Stiffened Cylindrical Panels 1027
a. Potting of panel ends 1032
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Contents ix

b. Installation of panel in test frame 1032


c. Application of loads in test frame 1033

13.7 Special Stiffened Shells 1034


13.7 .1 Corrugated Shells Subject to Axial Compression and
Bending 1034
13.7 .2 Corrugated Shells Subject to External Pressure 1038
13.7 .3 Elastically Supported and Sandwich Shells 1040
13.7.4 Tube-Stiffened Shells 1043

14 Composite Structures 1053


14.1 Background 1053

14.2 Flat Panels 1054


14.2.1 Unstiffened Panels 1054
a. Theoretical Considerations 1054
b. Axial Compression Loading Experiments 1063
c. Shear Loading 1085
d. Combined Loading 1102
e. Crippling 1105
14.2.2 Stiffened Panels 1117
a. Axial Compression Tests 1118
b. Shear Loading Tests 1147
c. Combined Loading 1155

14.3 Wing Box Structures 1157

14.4 Curved Panels and Shells 1165


14.4.1 Unstiffened Panels and Shells 1165
a. Theoretical Considerations 1165
b. Axial Compression Experiments 1169
c. Bending Experiments 1208
d. Shear Buckling Experiments 1212
e. Lateral Loading Buckling Experiments 1215
f. Combined Loading Buckling Experiments 1220
14.4.2 Stiffened Panels and Shells 1223
14.4.3 Corrugated Cylinders 1229
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

14.5 Concluding Remarks 1232

15 Nondestructive Buckling Tests 1243


15.1 Nondestructive Methods for Buckling Tests 1243
15.2 Vibration Correlation Techniques (VCT) 1244
15.2.1 Correlation between Vibration and Buckling 1244
15.2.2 Vibration Correlation Techniques (VCl) for Determination
of Boundary Conditions and Buckling Loads in Columns 1246
15.2.3 VCT for Determination of Boundary Conditions and
Buckling Loads in Plates 1252
15.2.4 Correlation between Vibrations and Buckling for Shells 1256
15.2.5 VCT for Determination of Boundary Conditions in Shells 1258
15.2.6 Application of VCT to Practical Boundary Conditions and
Realistically Fabricated Shells 1262
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
X Contents

15.2.7 VCT for External Pressure and Combined Loading 1270


15.2.8 VCT for Direct Prediction of Buckling Loads in Shells 1271
15.2.9 Other Recent Vibration Correlation Methods 1274
15.2.10 The Status of VCT 1283

15.3 Static Nondestructive Methods 1284


15.3.1 Experimental Determination of End Fixity 1284
15.3.2 Force/Stiffness Techniques 1288
15.3.3 Further Application of Force/Stiffness Methods 1290

16 Plastic Buckling Experiments 1299


16.1 Plastic Buckling Phenomena 1299
16.1.1 Introduction 1299
16.1.2 Inelastic Column Theory 1299
16.1.3 The Flow Theory versus Deformation Theory Paradox-
The Cruciform Column 1301
16.2 Plastic Buckling Experiments 1305
16.2.1 Plastic Buckling of Columns 1305
16.2.2 Plastic Buckling of Plates 1309
16.2.3 NACA Langley Tests in the Forties 1311
16.2.4 More Recent Plastic Buckling Tests on Plates 1313
16.2.5 Plastic Mechanisms in the Buckling of Thin-Walled Steel
Structures 1320
16.2.6 Plastic Buckling Tests on Cylindrical Shells under Axial
Compression 1327
16.2.7 Plastic Buckling of Cylindrical Shells Subjected to External
Pressure 1338
16.2.8 Plastic Buckling of Cylindrical Shells Subjected to
Bending or Torsion 1345
16.2.9 Plastic Buckling Tests on Conical, Spherical and
Torispherical Shells 1360

16.3 Combined Loading Tests in Plastic Buckling 1369


16.3.1 Biaxial Loading in Columns and Plates 1370
16.3.2 Biaxial Loading in Cylindrical Shells 1373
16.3.3 Caltech Tests on Combined External Pressure and Axial
Tension 1376
16.4 Southwell's Method in the Plastic Range 1389
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

16.4.1 Extension of Southwell's Method to Inelastic Columns 1390


16.4.2 Southwell Plots in Plastic Buckling of Shells 1393
16.5 Some General Remarks on Plastic Buckling 1403

17 Influence of Holes, Cutouts and Damaged Structures 1413


17.1 Effect of Holes and Cutouts on Plates and Shells 1413
17.1.1 Introduction 1413
17.1.2 The Effect of Holes and Cutouts in Plates 1413
17.1.3 The Effect of Holes and Cutouts in Shells 1418
17.1.4 Reinforcements 1423

17.2 Experiments on Plates with Holes and Cutouts 1426


Copyright Wiley
17.2.1 Metal Plates and Webs 1426
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Contents xi

17.2.2 Composite Plates 1434

17.3 Experiments on Shells with Holes and Cutouts 1437


17.3.1 Metal Shells and Curved Panels 1437
17.3.2 Composite Shells and Curved Panels 1445

17.4 Stability and Strength of Damaged or Dented Shells 1450


17 .4.1 Buckling and Strength of Damaged Structures 1450
17 .4.2 Damaged Stiffened Shells 1455
17 .4.3 Buckling of Delaminated Composite Shells and Panels 1462

18 Buckling under Dynamic Loads and Special Problems 1471


18.1 Dynamic Buckling Phenomena 1471
18.1.1 Background 1471
18.1.2 Dynamic Buckling Criteria 1473
18.1.3 Theoretical Considerations in Dynamic Pulse Buckling 14 76
a. Elastic buckling of a simply supported bar 1477
b. Elastic buckling under eccentric loads 1479
c. Dynamic plastic flow buckling of bars 1481
d. Technion theoretical model 1482

18.2 Impact Induced Buckling Experiments 1487


18.2.1 Column Buckling 1487
18.2.2 Plate Buckling 1495
18.2.3 Buckling of Arches and Spherical Shells 1501
a. Arches 1501
b. Spherical shells 1504
18.2.4 Buckling of Cylindrical Shells 1510
a. Buckling of rings and cylindrical shells subjected to
radial loading 1510
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
b. Buckling of cylindrical shells under axial impact 1514

18.3 Propagating Buckles 1524


18.3.1 Propagation of Bulges in Inflated Elastic Tubes 1525
a. Volume-controlled inflation experiments 1526
b. Determination of propagation pressure of bulges 1528
18.3.2 Propagation of Buckles in Long Tubes and Pipes under
External Pressure 1529
a. Propagation buckle experiments 1530
b. Propagation pressure in presence of axial tension 1532
18.3.3 Propagating Buckles in Long, Confined Cylindrical Shells 1535
a. Linearly elastic shells 1535
b. Elastic-plastic shells 1536
18.3.4 Buckle Propagation in Long Shallow Panels 1538

19 Thermal Buckling and Creep Buckling 1547


19.1 Introduction 1547
19.1.1 High-Temperature Effects in Structures 1547
19.1.2 Structural Responses to High Temperatures 1550
19.1 .3 Thermal Protection Systems 1555

19.2 High-Temperature Testing 1559


Copyright Wiley 19.2.1 Early Thermal Stress and Thermal Behavior Experiments 1559
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
xii Contents

19.2.2 Methods for Rapid Heating 1569


19.2.3 Measurements at High Temperatures 1584

19.3 Thermal Buckling 1596


19.3.1 Origins of Thermal Buckling 1596
19.3.2 Early Thermal Buckling Experiments on Plates and Shells 1598
19.3.3 Thermal Buckling Tests on Plates 1603
19.3.4 Thermal Buckling Tests on Shells 1611

19.4 Creep Buckling 1626


19.4.1 The Concepts of Creep and Creep Buckling 1626
19.4.2 Early Creep Buckling Experiments on Columns 1631
19.4.3 Creep Buckling Tests on Plates 1637
19.4.4 Creep Buckling Tests on Shells 1644

20 Some Comments on Measurements 1669


20.1 Introduction 1669
20.2 Strain 1671
20.2.1 Measurement of Strain 1671
20.2.2 Mechanical Strain Gages 1672
20.2.3 Optical Strain Gages 1673
20.2.4 Acoustical and Pneumatic Strain Gages 1675
20.2.5 Electrical Strain Gages 1676
20.2.6 Semiconductor Strain Gages 1678
20.2.7 Fiber-Optics Strain Sensors 1679
20.2.8 Strain Gage Circuits and Instrumentation 1680

20.3 Displacement Sensors 1680


20.3.1 Displacement Measurements in Buckling Tests 1680
20.3.2 Potentiometers and LVDT's 1681
20.3.3 Other Displacement Sensors 1683

20.4 Optical Methods 1684


20.4.1 Basic Optical Methods 1684
20.4.2 Photoelasticity and Photoelastic Coatings 1685
20.4.3 Moire Methods 1686
20.4.4 Holographic Interferometry and Speckle Methods 1689

20.5 Data Acquisition Systems 1689


20.6 Additional Sensing Devices 1691
20.6.1 Force Transducers-Load Cells 1691
20.6.2 Pressure Transducers 1692
20.6.3 Temperature Measurements 1692
20.6.4 Accelerometers and Vibration Measurements 1693
20.6.5 Acoustic and Thermal Emission Sensors 1694

20.7 Summary 1695


--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Author Index 1701


Subject Index 1707

The complete indexes for Volume 1 are included at the end of this book.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Preface to Volume 1

The motivation to write this book was the realization that in the vast literature on buckling
of thin-walled structures, and in particular in the many textbooks that have appeared during
the last few decades, the experiments have usually been relegated to the background and to
the secondary task of verification of theory. The authors felt, therefore, that a book written
from the viewpoint of the experimenter, emphasizing the strong interdependence of experi-
ment and theory, giving a detailed and critical review of the many important buckling ex-
periments carried out all over the world, in short a handbook assessing the state-of-the-art
was direly needed.
The book does not provide "cookbook recipes," but rather presents selected typical ex-
periments, which are often described in great detail, with some comments focusing on ques-
tions raised during the tests, the methods employed and the actual test atmosphere. The
choice of adopting or rejecting a certain technique is then left to the judgment of the reader.
In some cases minute details of an experiment were presented, since we felt that the accu-
mulated experience would be useful to the less experienced experimenter.
The wise experimenter should approach his or her tests with a fairly sound theoretical
background. We felt therefore that a certain amount of theory is essential also in this book.
Hence a brief review of buckling and postbuckling theory and numerical analysis is presented
in Chapters 2 and 3, and additional brief introductions to specific topics precede other chap-
ters. The aim of these reviews is to remind the reader of the theoretical basis, with emphasis
on the buckling phenomena and behavior, and of the computational tools available, and also
to provide the essential information for simple calculations.
Most of the fundamental theoretical ideas presented in Chapters 2, 3 and 5 are based on
many earlier tests referred to at the end of each chapter. As appropriate to a book devoted
to experimental methods, the theoretical derivations are rather concise, but are up-to-date
and include some novel approaches.
In a state-of-the-art handbook one cannot expect all readers to follow the text in an orderly
fashion, more probably they will often try to obtain specific information for their problem
by perusal of just the specific chapter of interest. We have also tried to accommodate these
readers, though they will find it helpful to refer back to other chapters, as indicated in the
chapter of their main interest. As the book is primarily concerned with test setups and
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

procedures, there is a slight overlap between the chapters that are ordered according to the
type of structural element tested. For example, some of the test rigs in Chapter 8 have also
been employed for stiffened plates, primarily covered by Chapter 12. Or some of them have
been built for metal and composite plates, mainly referred to in Chapter 14. Similarly, some
of the test rigs and procedures of Chapter 9 cover stiffened or composite shells as well,
pertaining to Chapters 13 and 14, respectively. We have, however, made an effort to avoid
actual duplications, and instead have referred the reader where appropriate to the discussion
in the relevant chapter.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
xiv Preface to Volume 1

One of the guidelines (or "Leitmotivs") throughout the book has been to emphasize the
potential interaction between the disciplines. For instance, the civil engineering tests and
aerospace experiments have been intentionally intermingled, to point out the similarity in
problems and phenomena.
On initial compilation of the book, the authors considered the advisability of discussing
some of the older experiments, in view of the rapid development of instrumentation and data
acquisition and reduction system, that makes the earlier equipment obsolete. However, as
the work progressed it became clear that the classic experiments of Fairbairn, von Karman,
Prandtl and some other outstanding investigators of the first half of the twentieth century
certainly deserve serious discussion on account of the questions they asked which have
proved sustainable and are still fully applicable today. Furthermore, the very extensive stiff-
ened shell experiments of the sixties and seventies, primarily motivated by the "golden age"
of space launcher development, outshine most more recent tests. They therefore justify de-
tailed consideration, as they are still the main source of experience (or data bank) to which
a young experimenter should turn.
Though fairly extensive, the lists of references (well over 2000) are by no means all
inclusive. Most of the significant experiments have been quoted, but certainly not all. For
example, due to limited accessibility, the references from the former Eastern Block are rather
sparse. However, in their choice of references the authors have endeavored to emphasize
how the important research activities transcend national boundaries and specific disciplines.
They expose buckling experimentalists to the vistas of benefits to be gained from the ex-
perience accumulated throughout the many laboratories all over the world, as well as clari-
fying the disadvantages of restricting themselves only to their immediate field of application.
Due to the special nature of the book, the authors requested information from many
colleagues at universities, research institutes and industry all over the world, to amplify the
data available in the literature. Gratitude is expressed to the hundreds of colleagues who
kindly provided the valuable information. photographs and sketches on their experimental
investigations, that assisted in the accurate, complete and up-to-date presentation of their
work. Obviously all this information is appropriately acknowledged throughout the book. In
some sections, it was felt fitting to quote verbatim from some papers, reports and corre-
spondence, and this is shown in the text by bracketing with double quotation marks.
The senior author (J. Singer) would like to express his appreciation to the late Professor
Charles (Chuck) D. Babcock of the California Institute of Technology, with whom he shared
the initial stages of conception of the idea of the book in the early eighties.
The senior author thanks in particular, Professors P.C. Birkemoe (University of Toronto),
S.R. Bodner (Technion), C.R. Calladine (Cambridge University), G.A.O. Davies (Imperial
College London), D. Durban (Technion), G.D. Galletly (University of Liverpool), S. Kyr-
iakides (University of Texas), A. Libai (Technion), N.W. Murray (Monash University, Mel-
bourne), H. Ory (RWTH Aachen), K.A. Stevens (Imperial College London), who were so
kind to read portions of the manuscript and whose comments contributed to the relevant
discussions.
The authors would also like to thank Mrs. B. Hirsch of Technion, Mrs. A. van Lienden-
Datema of TU Delft, Ms. S. Bryant of Caltech and Ms. Kirsten Maclellan of UCLA for
their devoted typing of the manuscript; Mrs. R. Pavlik and Mrs. D. Rosen of Technion, Mrs.
P.E.C. Zwagemaker of TU Delft and Mrs. B. Wood of Caltech for preparation of drawings,
and the librarians Mrs. S. Stern, Ms. A. Szmuk and Ms. S. Greenberg of Technion, Mrs. J.
Anderson and Mrs. P. Gladson of Caltech, and Mr. W. Spee of TU Delft for their kind
assistance. Thanks also to Mr. A. Grunwald, chief technician of the Technion Aerospace
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Structures Laboratory for his many-faceted assistance.


The authors would also like to extend their thanks to the Lena and Ben Fohrman Aero-
space Structures Research Fund, the Jordan and Irene Tark Aerospace Structures Research
Fund and the Caltech Sherman Fairchild Distinguished Scholars Fund for their generous
support.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Preface to Volume 1 xv

Thanks are also due to the editors and staff of John Wiley & Sons for their continuous
cooperation.
Last but not least, a word of praise to our wives Shoshana Singer, Margot Arbocz and
Ruth Weller. It is no exaggeration to say that without their encouragement and patient un-
derstanding we could not have completed this book.

Josef Singer
Johann Arbocz
Tanchum Weller

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Preface to Volume 2

This is the second volume of the two-volume book Buckling Experiments, a handbook as-
sessing the state of the art of experimental methods in buckling of thin-walled structures
from the viewpoint of the experimenter, emphasizing the strong interdependence of experi-
ment and theory. Though the Preface that appears in volume I (which emphasizes the phi-
losophy that guided our presentation and is therefore repeated here) covers both volumes,
some additional remarks are warranted.
The two volume~ are closely related and interconnected. While volume l addresses basic
concepts, columns, beams, arches and plates, volume 2 considers shells, stiffened plates and
shells, composite structures, plastic buckling, cutout and damage effects, dynamic loads,
thermal buckling, nondestructive tests and measurements. There are, however, extensive
cross-references between the two volumes, that follow from the affinity of the behavior of
ditierent structural elements, and of different test setups.
Volume 2 is considerably bulkier than volume l, resulting from the choice of a logical,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

contents-derived dividing line between plates and shells, and from a detailed exposition of
the more recent developments in the field. Again, as in volume l, we have been aided by
hundreds of colleagues all over the world, who kindly provided us with valuable information,
which is appropriately acknowledged throughout the book. Though the list of references has
more than doubled, it is still by no means complete. As in volume 1, we have again quoted
verbatim from some papers and reports, which is shown in the text by double quotation
marks or by indented smaller print.
It should be noted that: all figures reproduced from American Institute of Aeronautics and
Astronautics (AIAA) publications are reprinted with permission; all figures reproduced from
American Society of Mechanical Engineers (ASME) publications are reprinted with permis-
sion; all figures reproduced from Society of Experimental Mechanics (SEM) publications are
reprinted with permission from the Society of Experimental Mechanics, Inc., Bethel CT; all
figures reproduced from American Society for Testing Materials (ASTM) publications are
reprinted with permission; all figures reproduced from Institution of Mechanical Engineers
(!MechE) are reprinted with permission; all figures reproduced from Elsevier publications
are reprinted with permission; all figures reproduced from Academic Press publications are
reprinted with permission; and all figures reproduced from Springer-Verlag publications arc
reprinted with permission and their copyright is owned by Springer-Verlag.
Thanks are extended to Mr. Joel Stein and the editors and staff of John Wiley & Sons,
New York, for their continuous cooperation in the completion of this volume.
The authors, and in particular the senior author (J. Singer), would like to repeat their
thanks extended in the Preface that precedes volume 1 to many colleagues and co-workers
for their assistance, and express additional thanks also to Mrs. B. Hirsch, Mrs. A. Goodman,
Mrs. M. Zeitelbach, Mr. V. Salit, Ms. M. Ilan, Mrs. N. Rein and Mrs. S. Mokady, and all
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
xviii Preface to Volume 2

of the Technion. Thanks are due once more to the Jordan and Irene Tark Aerospace Structures
Research Fund for their continued support.
Finally, the authors would like to repeat their praise to their wives, Shoshana Singer,
Margot Arbocz and Ruth Weller, for their encouragement and support. J. Singer would like
to add his special thanks to Shoshana for her continuous, valuable assistance and advice in
the final stages of the work.

Josef Singer
Johann Arbocz
Tanchum Weller

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Abbreviated Table of
Contents - Vol. 1

Vol. 1: Basic Concepts, Columns, Beams and Plates v


Preface XV

1 Introduction 1
1.1 Experiments as Essential Links in Structural Mechanics 1
1.2 The Role of Experiments in Structural Stability 3
1.3 Motivation for Experiments 5
1.4 Bridging Gaps Between Disciplines 9

2 Concepts of Elastic Stability 15


2.1 Physical Concepts - Types of Observed Behavior and
Their Meaning 15
2.2 Mathematical Models for Perfect Structures 94

3 Postbuckling Behavior of Structures 131

3.1 Introduction 131


3.2 Asymptotic Imperfection Sensitivity Analysis 134
3.3 Direct Solutions of the Nonlinear Stability Problem 154

4 Elements of a Simple Buckling Test - a Column Under


Axial Compression 181

4.1 Columns and Imperfections 181


4.2 Von Karman's Experiments 182
4.3 The Basic Elements of a Buckling Experiment 185
4.4 Demonstration Experiments 187
4.5 Southwell's Method 194
4.6 Application of the Southwell Method to Columns, Beam
Columns and Frames 197
4.7 Remarks on the Applicability of the Southwell Plot 207

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
XX Abbreviated Table of Contents - Vol. 1

5 Modeling - Theory and Practice 217


5.1 Mathematical and Physical Modeling 217
5.2 Dimensional Analysis 218
5.3 Similarity 220
5.4 Application to Statically Loaded Elastic Structures 223
5.5 Loading Beyond Proportional and Elastic Limits 228
5.6 Buckling Experiments 229
5.7 Scaling of Dynamically Loaded Structures 237
5.8 Scaling of Composite Structures 259
5.9 Model Analysis in Structural Engineering 272
5.10 Analogies 282

6 Columns, Beams and Frameworks 289


6.1 Buckling and Postbuckling of Columns 289
6.2 Crippling Strength 309
6.3 Torsional-Flexural and Distortional Buckling 320
6.4 Lateral Buckling of Beams 328
6.5 Interactive Buckling in Columns and Beams 344
6.6 Beam-Columns 356

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
6.7 Buckling of Frameworks 367

7 Arches and Rings 409


7.1 Background 409
7.2 Shallow Arches 410
7.3 Rings and High Rise Arches 434
7.4 Lateral Buckling of Arches 440

8 Plate Buckling 453


8.1 Buckling and Postbuckling of Plates 453
8.2 Experiments on Axially Compressed Plates 470
8.3 Determination of Critical Load and Southwell's Method
in Plates 516
8.4 Experiments on Shear Panels 538
8.5 Web Crippling 561
8.6 Biaxial Loading 570
8.7 Guidelines to Modern Plate Buckling Experiments 577

Author Index to Vol. 1 603


Subject Index to Vol. 1 611

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
9
Shell Buckling Experiments

9.1 Introduction
In contrast to columns, which have a neutral postbuckling path, and plates, which exhibit a
stable postbuckling behavior, shells usually have a very unstable postbuckling behavior that
strongly influences their buckling characteristics. Thin shells, however, are very efficient
structures that can support very high buckling loads and hence their buckling and postbuck-
ling have presented scientific and engineering challenges for decades. The extensive theo-
retical studies, discussed briefly in Chapters 2 and 3, have clarified the phenomena, connected
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

initial postbuckling behavior with imperfection sensitivity, developed analysis procedures,


and established the imperfection sensitivity of some shell/load combinations. Unfortunately
the impact of these modern methods of analysis on engineering practice has as yet been very
small, and empirical "knock-down" factors are still primarily relied on in the design of
buckling-critical shells (see discussion in [1.14] and [1.18]).
One of the reasons for this slow and incomplete technology transfer from researcher to
designer is probably the relative complexity of the analysis, as well as the difficulties en-
countered in correlating theory with experimental results. Another reason may have been the
lack of experimental investigations that were closely coordinated with theoretical studies,
though thousands of shell buckling tests have been carried out.

9. 1. 1 Historical Background-Shells under External Pressure


The earliest shell buckling tests were probably carried out in 1845-1850 by Fairbairn and
Hodgkinson in England on thin-walled tubes under axial compression and bending, in con-
nection with the design of the Britannia and Conway Tubular Bridges, as mentioned in
Chapter 8 (see Figure 8.1 and [8.2]). A few years later, Fairbairn carried out extensive
experimental studies on tubes under external pressure, reported briefly in 1857 to the British
Association for the Advancement of Science [9.1] and in more detail in 1858 to the Royal
Society of London [9.2]. The studies were undertaken "to determine the laws which govern
the strength of cylindrical vessels exposed to uniform external force, and their immediate
practical application in proportioning more accurately the flues of boilers," since due to
insufficient knowledge, the increase in working pressures resulting from the "desire to econ-
omize" was "accompanied by an increase of dangerous and fatal accidents from boiler
explosions."
Fairbairn's reports were not only the first accounts of a comprehensive experimental pro-
gram in shell stability (35 shells subjected to external pressure and 7 to internal pressure),
but they, in particular his 1858 Royal Society Paper, present a well documented discussion
of a systematic and carefully planned study that provides enlightening reading even today.
Figure 9.1 a (reproduced from [9.2]) shows the vertical cross-section of Fairbairn's large 8 ft
Copyright Wiley
Provided byBuckling
IHS MarkitExperiments:
under license withExperimental
WILEY Methods
in Buckling of Thin-Walled Structures:
Licensee=McDermott Inc -Shells, Built-Up User=G,
India/8215328006, Structures, Composites
Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
and Additional Topics – Volume 2. J. Singer, J. Arbocz and T. Weller Copyright © 2002 John Wiley & Sons, Inc.
624 Shell Buckling Experiments

to)

( b)

REM/IRKS

N. 8 30
~ --·
-

~
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
0. 8 39 .0 4 3 32

P. 8 40 .0 4 3 31
~
, __

Figure 9.1 Fairbaim·s (857 test setup for Study or the resistance of tubes l O collapse under external
pressure (from [9.2]): (a) Venical cross-sect ion or cast-iron pressure vessel, placed in a
locomotive pit with a test shell in it. (b) Typical table of results for tubes of (R/1) = 93.
indicating collapse modes of shoner shells

(2.44 m) long and 2 in. (5.08 em) thick cast iron pressure vessel, which was placed in a
locomotive pit with a test shell in it. A typical table of re.~ults from [9.2] for tubes of (Rit)
= 93 is shown in Figure 9.1 b, indicating the collapse modes for some of the shorter shells.
The 32 cyli ndrical tubes tested under external pressure covered a large range of geometries
defined by (Rit) = 29.4-140 and (LIR) = 2.28- 30. Secondary effects, like the effect of lap
joints, the importance of out-of-roundness. the infl uence of change of material, and even the
behavior of elliptic tubes, were also studied. The strength of similar cylindrical shells under
internal pressure was also tested to show the "comparative weakness of tubes subjected to
external pressure." Practical applications of the results are discussed. and Fairbairn points
out the structural inefficiency of the then customary boiler designs with "immense excess
of strength in the outer shell" loaded by internal pressure and " the extra thickness of boi ler
plate which causes it. being so much material thrown away. adding nothing to the strength
whi lst the nues (subjected to external pressure) remain in so dangerous ly weak a condition."
A separate study of the collapse of glass cylinders and globes [9.3] was also carried out to
verify and extend the results obtained earlier for the wrought iron riveted tubes.

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Introduction 625

For nearly half a century, Fairbairn's experiments, and the formula in which he summa-
rized his results, were the basis of boiler design and the starting point for all the discussions
on improved empirical formulae. Some of the best known of these were the formulae of
Grashof [9.4], Love [9.5], Nystrom [9.6], Unwin [9.7], Belpaire [9.8] and Wehage [9.9], and
their derivation was sometimes accompanied by criticism of Fairbairn's conclusions. There
were also lively discussions in the engineering community about their relative merits. For
example, in 1876 there was such a vigorous debate in successive issues of London Engi-
neering in which Wilson, Unwin, Fletcher (the Chief Engineer of the Manchester Steam
Users Association) and the editor of Engineering participated (see [9.1 0]). In the debate
some other isolated boiler tests were also mentioned, but no systematic experiment other
than those of the late Sir William Fairbairn. A comprehensive and critical review of the
experimental investigations, primarily those of Fairbairn, and the various empirical formulae
based on them, was published in 1881 by Roelker in the United States. One should note the
last section of his paper, which contains detailed recommendations for future experiments
and suggestions for their reporting which are applicable even today.
Some small series of tests were indeed carried out in the 1860s, 70s and 80s, as for
example those of Messrs. Russels and Sons, and of Fletcher, discussed by Clark in his
Manual of Rules [9.12], and later in Engineering [9.13], or the tests on the old boilers and
superheaters of the steamship Pharos reported by Richards in the Engineer [9.14]. But all
these were tests and collapse records of flues from operational steam boilers, sometimes
sensibly worn. The exception was an extensive series of tests of 18 circular wrought iron
tubes of (RI t) = 30.8-66.7 and (L/ R) = 0.67-3.96, representing the geometries used in ship
boilers, carried out at the German Navy Docks, Danzig, in 1887-92. These tests were dis-
cussed in detail by C. Bach, one of the leading German researchers in strength of materials

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
at the time [9 .15]. One may note that in the last two decades of the 19th century Bach carried
out a wide range of experimental investigations at the materials and structures testing lab-
oratory he established in Stuttgart, which included also the first experimental study of the
buckling of spherical and tori spherical caps [9 .16].
Bach began his paper on the Danzig experiments by pointing out that, though Fairbairn's
test specimens were too small and weak to be relevant to practical designs in the 1880's,
scientists and engineers in England, France and in particular Germany had been preoccupied
for decades with mathematical processing of Fairbairn's results. This was also emphasized
by the investigators at the beginning of the 20th century, especially Carman at the University
of Illinois [9.17] and [9.18], and Stewart at the U.S. National Tube Company [9.19], who
carried out very extensive systematic experimental studies of tubes under external pressure.
Both Carman and Stewart noted with surprise that an important problem, "the solution of
which has not only great scientific interest but also valuable technical applications, should
remain for so many years with so little experimental work." (Both, however, overlooked the
German Navy Danzig tests published more than a decade earlier.)
Carman's and Stewart's experiments covered a broad range of geometries. Carman tested
132 tubes in 1905-06 with (Rit) = 7.9-35.7 and (L/R) = 1.8-153.3, seamless cold drawn
flues, lap welded steel boiler flues and brass tubes; and Stewart tested 514 Bessemer steel
lap-welded tubes in 1902-04, with (Rit) = 6.23-32.0 and (L/R) = 6.96-80.7. Two com-
prehensive test programs indeed. Stewart's 1906 paper [9.19] represents a well-planned ex-
perimental study, defining at the outset the factors on which the resistance of tubes to external
pressures were assumed to depend and investigating the influence of these factors. For ex-
ample, "since it was anticipated that the out-of-roundness of the tube would exert a con-
trolling influence on its behavior," an autographic calipering apparatus was constructed and
used to record automatically the divergence from roundness of tubes at many stations along
their length. From his experimental results, Stewart developed empirical formulae, charts and
curves for designers, essentially a similar empirical approach to that of Fairbairn and of the
investigators who reprocessed Fairbairn's experimental results in the second half of the 19th

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
626 Shell Buckling Experiments

century. Hence Stewart's careful experimental investigation was actually only a broader and
more modern version of Fairbairn's work of 50 years before, without any innovative con-
cepts.
Carman [9.17] and [9.18], on the other hand, tried to correlate his experimental results
with "rational formulae," with the predictions of the theoreticians Bryan [9.20], Bassett
[9.21] and Love [9.22], who had attempted in the two preceding decades to treat this "dif-
ficult" problem of "the stability of an elastic system." One may note here that it was Fair-
bairn's experiments, reported in his 1858 paper, which attracted the attention of these inves-
tigators, who arrived at a mathematical solution for long shells near the end of the 19th
century. They obtained a formula for the critical pressure of long tubes,

1 (n2 ~ 1) 1 ~E v2 ( R
Pu = 12 t )' (9.1)

where tis the thickness of the tube, R its radius, E Young's modulus and n the number of
circumferential lobes or waves into which it buckles or collapses. (Their notation differs
slightly since they write 2h for the thickness instead of t.) For minimum collapse pressure
n = 2, as indeed observed in experiments for very long tubes, yielding

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Pee = 4I 1 ~E v 2
( t )'
R (9.2)

Note that Eq. (9.2) is very similar to the earlier solution obtained by M. Bresse in 1866 and
independently by M. Levy in 1884 for a uniformly compressed ring [2.26] and [2.27] and
can now be derived from it simply by substitution of El (1 ~ v 2 ) forE there and (t 3 112) for
I there. But as pointed out by Carman, before 1906, "no use had been made of this formula
[Eq. 9.2] in experiments or in engineering practice." Carman also noted that two earlier
attempts to derive rational formulae, by Grashof and Unwin [9.4] and [9.7] were unsuccessful
because their authors "did not appreciate the very limited range of Fairbairn's experiments
as regards length of tubes," and these were the only systematic experimental results available
at the time. For thin-walled tubes, with (Rit) > 20, Carman was able to correlate his ex-
perimental results with the theoretical conclusions of Bryan and others, but he still had to
resort to purely empirical formulae for the thicker tubes.
A few years later Slocum, from the University of Cincinnati [9.23], evaluated the com-
prehensive experimental data of Carman and Stewart, by comparison with Love's formula,
Eq. (9.2), for thin tubes and with the classical Lame's formula for thick ones. He concluded
that the experimental data provided the appropriate correction factors for these rational for-
mulae as required by "the imperfections of certain types of commercial tubes," and that "all
that now remains to be done is the experimental determination of the correction constant for
other types of commercial tubes." Later studies, however, proved this conclusion to be over-
simplified.
The buckling of short shells under external pressure was studied experimentally by Cook
in London nearly a decade later [9.24] and [9.25]. Cook tested machined solid drawn steel
tubes of (RI t) = 97-206 and (L/ R) = 0.68-4.22, and correlated his results with the theory
for short shells derived by Southwell at about the same time [9.26]. One may note that
similar theoretical results were obtained contemporaneously by von Mises in Germany [9.27].
In 1917 also Carman returned to the problem [9.28], carrying out another extensive ex-
perimental program which included over 150 tubes (seamless cold-drawn steel tubes, seam-
less brass tubes, aluminum, glass and hard rubber tubes, with (Rit) = 8-50), and correlating
his results with the newer theoretical treatment of Southwell [9.29]. After discussing the
theoretical studies and the earlier experimental investigations, Carman focused on the "crit-
ical length," beyond which the collapse pressure is independent of the length, and pointed
to disagreement between his experimental results and Southwell theoretical formula L, =
kVJ53Ti. It is worth noting that Southwell himself did not consider the "critical length" of
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Introduction 627

importance, but stressed the discontinuous dependence of the collapse on the number of
lobes. Southwell also carried out experiments on seamless steel tubes, with (RI t) ~ 18, to
confirm the shape of his theoretical curves [9.29]. Cook also returned to the problem in 1925
[9.30] with a series of tests on thin brass tubes of (RI t) ~ 63 and (L/ R) = 0.48-13.3. Cook
pointed out the importance of in-plane constraints and devised spherical surface end-fittings
which "secured freedom from longitudinal constraint," while "thin rubber bands placed
around the lines of the contact made perfectly water-tight joints." Very satisfactory agreement
with predictions, by Southwell's analysis, was obtained for collapse pressures, but the number
of lobes observed was less than predicted, except for n = 2 for very long tubes. Cook pointed
out that "it must be remembered that the final shape of the cross-section is not necessarily
that which exists at the moment of collapse," an important observation that is well appre-
ciated by shell buckling investigators today.
On the other hand, in experiments performed by von Mises in 1918 (and reported a decade
later [9.31]) on a series of large thin-walled mild steel shells, of (Rit) = 400 and (L/R) =
0.3-0.6, loaded by hydrostatic pressure, the observed number of circumferential waves
agreed very well with the predicted ones. The-for the period-unusually high (Rit) = 400
in these tests should be noted.
It may be pointed out here that the motivation for the study of the collapse of cylindrical
shells under external pressure in the 19th century came primarily from the designs of steam-
boiler flues and of similar larger pipes and containers, used for drainage of farm lands, for
distillation, for refrigeration and for food canning, which were subjected to pressure loads
that could induce buckling (see [9.32]). Further study was stimulated early in the 20th century
by problems arising in the design of submarines, tunnel liners and dome structures. However,
nearly all the early experiments on shells were made on comparatively thick tubes and shells,
which usually failed by plastic buckling, or due to yielding of the material, rather than by
elastic instability. With the exception of the 1918 von Mises test program, the experiments
with very thin shells, where elastic buckling predominates, appeared only in the late twenties
and thirties, with the widespread application of thin shells to airplane structures, in which
buckling under axial compression gained prominence.
However, before we discuss axial compression, the extensive studies of collapse of cylin-
drical shells under external pressure in the late twenties and early thirties, motivated primarily
by submarine design, warrant a closer look since they contributed significantly to the un-
derstanding of shell buckling and postbuckling behavior, in particular that of stiffened shells.
Many of these studies were carried out at the U.S. Navy Experimental Model Basin in
Washington, D.C. (later called the David Taylor Model Basin) under the auspices of the
ASME (American Society of Mechanical Engineers) Special Research Committee on the
Strength of Vessels under External Pressure [9.33]-[9.35]. As pointed out in [9.33], this
ASME committee and the related research program were organized due to "the lack of
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

adequate design formulas" and "the dearth of comprehensive experimental data" for the
larger, shorter and thinner shells employed in the early thirties. The experiments carried out
in 1929-34 at the U.S. Navy Experimental Model Basin excel in their precision, care of
execution and correlation and include 36 short and thin shells, in a wide range of geometries,
(Rit) = 74-263 and (LIR) = 0.125-1.0. For example, in [9.33] the importance of overall
and local imperfections and the technique of their measurement is discussed, and the lack
of recorded imperfection measurements is pointed out to be a primary weakness of previous
experimental investigations. Also the essential requirements of geometrical and material sim-
ilarity for reliable model tests are emphasized.
In 1932 Saunders and Windenburg published a valuable paper on the use of models [9.34],
which reports on what is probably the first systematic study of scaling problems and model
manufacturing techniques for shells. Even today, the paper presents an outstanding account
of the model approach with its associated problems and provides very instructive reading.
Among the topics emphasized in it are, for example, the importance and technique of mea-
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
628 Shell Buckling Experiments

surement of the shell contour (its initial imperfections) and the problems of attachment of
stiffening rings in models.
The measurement of initial imperfections, roundness and thickness variations is also
stressed by Jasper and Sullivan in the discussion of their 1931 tests of 14 arc-welded tubes,
of (R/t) = 17-31, and of 4 seamless tubes, of (R/t) = 9-11, carried out at the A.O. Smith
Corporation in Milwaukee [9.36]. Note that industry had begun to appreciate imperfection
measurement.
The second paper published by the ASME Special Research Committee on the Strength
of Vessels under External Pressure [9.35], summarized the U.S. Navy experimental results,
correlated them with the theoretical formulae for short tubes of Southwell, von Mises and
Tokugawa [9.26], [9.27], [9.31] and [9.37], and presented a simple approximation to the von
Mises formula (which they preferred), the U.S. Navy Experimental Model Basin formula:
2.42£ (t/D) 512
Per = (1 J (9.3)

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
")1/4 [
- v-. (LID) - 0.45(t/D) 112

where L is the length of tube and D its diameter. Equation (9.3) is still used today by most
designers.
Two other important test programs on buckling of cylindrical shells under external pressure
also belong to this group of investigations. The experiments carried out by Tokugawa for
the Japanese Navy during 1925-28 [9.37] and those performed by Sturm at the University
of Illinois during 1928-29 and at the ALCOA Aluminum Research Laboratories during
1931-33 (reported nearly a decade later in [9.38]). Tokugawa's investigation [9.37] included
4 large-scale steel models (details of which were, however, classified) and 96 small-scale
brass models of (Rit) = 100 and 200 and (LIR) = 0.4-3.0. Most of the models were ring-
stiffened shells, and the study focused on the two distinct modes of collapse of ring-stiffened
(called "cross-stiffened" by Tokugawa) circular cylindrical shells: local and general insta-
bility and their transition, as well as on the boundary conditions introduced by the ring
frames. In the accompanying theoretical study, Tokugawa derived independently a formula
for buckling of cylindrical shells subjected to hydrostatical pressure, similar to that of von
Mises [9.31], but included it in a "frame factor," to be determined by experiment, which
made the formula applicable to ring-stiffened shells. It was probably the first comprehensive
study of stiffened shells.
Sturm [9.38] emphasized in his analytical and experimental work the difference between
hydrostatic and lateral pressure, the influence of end conditions and the effects of plasticity
and "out-of-roundness." The two large-scale 6-m-long black sheet iron thin-walled pipes,
with (R/t) = 143-159, tested at the University of Illinois represented specimens fabricated
by current structural engineering methods and used a number of novel test techniques. For
example, the smaller-diameter pipe was placed inside the larger one and the external pressure
was obtained by pumping air out of the space between them. The inner pipe then served as
a mandrel preventing excessive radial deformation and thus permitting repeated buckling
tests without a measurable permanent set. The second series of experiments at ALCOA
included 33 aluminum alloy extruded and fabricated tubes, with (Rit) = 49-213 and (L/R)
1.7-25.0, and confirmed the predicted marked effect of fixed edges on the collapsing pres-
sure. The most important contribution of Sturm's studies was, however, the confidence it
gave to structural engineers in the methods of calculation of shells subjected to external
pressure.

9.1.2 Historical Background-Axially Compressed Shells


Though thin-walled circular tubes under axial compression were tested already in 1846 by
Fairbairn and Hodgkinson [8.1], their buckling was first recognized in the beginning of last
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Introduction 629

century by "wrinkling" or "secondary flexure" in columns. In 1905-08 Lilly carried out an


extensive series of experiments on mild steel tubes at Trinity College in Dublin, Ireland
[9.39]-[9.41], accompanied by a theoretical study. He was primarily concerned with the
economic design of columns (an early optimization approach) and hence his tubes were
rather thick-walled with (Rit) = 4-40. He clearly showed, however, the "remarkable wave
phenomena" that occur "in connection with secondary flexure" and that the load producing
failure becomes smaller the larger (R It), and stressed that "the true strength to compression
of the tube is the load which produces the wave formation." Lilly's experiments represent
the first systematic study of elasto-plastic buckling and collapse of axially compressed cir-
cular cylindrical shells. In 1908, Mallock [9.42] showed similar instability failure patterns
and initiated a nearly inextensional study, "a particular case of the general theory of the
form of the folds of drapery," which was much later developed independently by Yoshimura
into his postbuckling pattern [9.79]. Another series of carefully conducted tests on solid
drawn steel tubes of (Rit) = 11.2 and 19.2 under axial compression was presented by Mason
in 1909 as part of his study on yielding of mild steel under combined stress [9.43]. He
observed that early yielding by flexure occurred in some of the thinner tubes, due to wrin-
kling, and that collapse under axial compression "always occurred by wave like flexure of

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
the walls of the tubes."
In the same period, the problem of wrinkling, buckling, or elastic instability of circular
cylindrical shells was investigated theoretically by Lilly [9.40], Lorenz [2.37], [9.44],
Timoshenko [9.45], [9.46], Southwell [9.29] and later by Dean [9.47], yielding the main
results of the classical linear theory discussed in Chapter 2.
In practice, however, buckling of cylindrical shells under axial compression became im-
portant as their use in aircraft structures broadened, first as thin-walled columns and then in
the stressed-skin construction of fuselages and wings, introduced in the late twenties and
thirties (see [9.48]). It soon became the central design problem of aerospace structures.
The early experiments of Barling and Webb on annealed steel tubes [9.49] carried out at
the Royal Aircraft Factory in 1914-15 were motivated by the design of aeroplane struts.
Their tubes were short and thick, (RI t) = 2-30, and they observed wrinkling, or as they
called it, "crinkling," stresses much below the yield stress for the thinner tubes, when
(Rit) > 16.7. "As for the explanation of crinkling stress," they "believe that it is of the
nature of elastic instability," and show that for thin tubes, with (t I R) < 0.06, the crinkling
stress is directly proportional to (tl R). They noted also that "inward crinkling reduces the
moment of inertia of the strut," a premonition of the Brazier instability perceived a decade
later. The design requirements of the Royal Aircraft Factory also initiated the study of the
failure of short tubular struts of high-tensile steel by Popplewell and Carrington at Man-
chester [9.50]. They tested their nickel-chrome tubes, of (Rit) = 5.8-29.5, in the hardened
and annealed conditions and confirmed Barling and Webb's conclusion that the crinkling (or
wrinkling) stress varies as (tl R) when (tl R) < a constant, which for the nickel-chrome tubes
is 0.1 (it depends on the yield stress of the material).
The first comprehensive series of experiments correlated with theoretical predictions were
those carried out by Robertson at the Royal Aircraft Establishment in 1915 and later at
Manchester but reported only a decade later [9.51] and [9.52]. Four sets of short tubes were
tested: nickel-chrome steel tubes with (Rit) = 12-96, mild steel tubes with (Rit) = 46-175,
tubes made of thin high-tensile steel strips with (R It) = 110-500, and tubes made of silver
spruce strips with (Rit) = 5-75. Robertson emphasized the difference between the collapse
behavior of thick tubes, which is determined by the material behavior (the yield stress), and
that of thin tubes, which is determined by elastic instability (buckling), indicating the dif-
ferent collapse patterns as well as the different autographic (stress-strain) diagrams and show-
ing the transition value of (t I R) for the various sets of tubes. He also demonstrated for some
of his largest tubes, made of high-tensile steel strips, with (Rit) = 500, that the buckling
deformation is primarily elastic and disappears upon removal of the load. In these tests he
employed a mandrel to restrict the buckle depth in order to prevent permanent plastic de-
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
630 Shell Buckling Experiments

formations, a technique that was reinstated and widely used only four decades later. Rob-
ertson pointed out that the comparison with available theoretical analysis, based on elastic
theory of thin shells, applied only to the thinner specimens and showed that the experimental
collapse loads were about 0.6 of the theoretical ones. He suggested that geometric imper-
fections were the main reason for this discrepancy. In his tests the observed number of lobes
was always less than the predicted one in approximately a similar ratio.
The rapid growth of air transportation in the thirties and the associated development of
semi-monocoque aircraft shell structures motivated the careful experimental studies on cy-
lindrical shells under compression of Lundquist at NACA Langley [9.53] and of Donnell at
the California Institute of Technology [9.54]. At the same time, the increased employment
of thin-walled columns in civil engineering structures prompted an extensive study by Wilson
and Newmark at the University of Illinois Engineering Experiment Station [9.55]-[9.57]. At
about the same time in Europe, a German civil engineer, Fliigge at the University of Got-
tingen, independently studied the stability of cylindrical shells and carried out experiments
on shells under axial compression [2.48]. Lundquist at NACA tested 45 duraluminum cyl-
inders of (RI t) = 333-1415 and (L/ R) = 0.25-6.0; Donnell at Cal tech tested 19 small steel
and 21 small brass shells of (Rit) = 160-1440 and (LIR) = 2-32; Wilson and Newmark at
the University of Illinois tested 33 machined seamless steel tubes and 39 large fabricated
steel cylinders of (Rit) = 35-990 and (LIR) = 1.4-24.7; and Fliigge at Gottingen tested 3
rubber shells and 13 celluloid shells of (Rit) = 90-136 and (LIR) = 1.8-5. These compre-
hensive tests covered a large range of geometries as well as materials and could therefore
serve as a database for designers.
Comparison of experimental buckling stresses with predictions by the then well-known
and accepted linear classical theory, which for a Poisson's ratio of v = 0.3 can be written
as
uc~ = 0.605 E (tl R) (9.4)
showed, however, great scatter, with experimental values being very much lower than the
theoretical values, though failure in most of these relatively thin shells was due to elastic
instability. The length of the cylinder was found to have very little effect on the buckling
load, unless it was very short, and similarly the effect of end fixity was very small. The ratio
of experimental to predicted buckling load or stress, later to be known as "knock down
factor," p = (uexr/ ucJ), varied between 0.10-0.65, depended on the method of fabrication of
the shells, and showed a definite tendency to decrease with increasing (R It). This discrepancy
between theory and experiment was too great to be accepted, it was a challenge to researchers
and became the motivation for the development of large deflection theory for imperfect shells
in the decades to come.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

The engineers, however, could not wait and embarked on further testing programs. As a
matter of fact, from the mid-thirties to the late fifties, most shell buckling tests were for
design data only and their results usually exhibit wide scatter and nonrepeatability (see also
[9.58]). In the last three decades, however, more careful experiments, more closely accom-
panied by analysis, have been carried out that have not only helped the understanding of
buckling and postbuckling behavior but also have begun to influence the designers.
Furthermore, whereas in the past reviews on shell buckling (as for example [3.16], [9.59]-
[9.62], and [9.78]), relegated experiments to the secondary task of verification of theory,
more prominence began to be given in the eighties to certain aspects of experimental work,
as for example imperfection measurements in [1.14], and surveys focusing on experimental
studies began to appear. Examples of these are [1.15], which discusses work carried out at
the Institute for Aerospace Studies of the University of Toronto in the seventies; [9.63],
which summarizes buckling tests of fabricated steel cylindrical shells in the United States
till 1982; [1.16], which outlines buckling research performed at Det norske Veritas in Oslo
in the late seventies and early eighties; [9.64], in which empirical design rules for unstiffened
cylindrical, conical and spherical shells are proposed as a result of a very comprehensive
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Buckling and Postbuckling Behavior 631

literature survey; or [9.58], [1.17] and [1.24], which review the status of experimental buck-
ling investigations on shells till 1982 and are the basis of this chapter.

9.2 Buckling and Postbuckling Behavior of Axially


Compressed Cylindrical Shells
As mentioned in the historical introduction, the center of attention of shell buckling for more
than half a century has been the perplexing behavior of the isotropic circular cylindrical
shell under axial compression (see, for example, [9.65]). Hence considerable experimental
effort has also been devoted to this problem, which will therefore be the first case of shell
buckling to be discussed in detail.

9.2. 1 Sequence of Events in an Axial Compression Experiment


If one applies axial compression to a thin cylindrical shell of medium length in a typical
test fixture (shown schematically in Figure 9.2) and plots the axial load versus axial short-
ening, one obtains a curve like Figure 9.3 (the full and dashed lines). The test fixture is
taken to be deformation controlled and the axial load is increased from zero by a slow
decrease in the distance between the end plates (increase in axial shortening). The prebuck-
ling behavior is approximately linear and the shell essentially retains its original shape till
it buckles at A. There, the shell not only evades any further load increase by buckling, but
the load drops abruptly to a fraction of the buckling load at B (as shown by the dashed line
AB in Figure 9.3), while the shell snaps swiftly into the typical two-tier postbuckling pattern
(sometimes called "diamond pattern") shown in Figure 9.4. If the test machine is very rigid,
AB will be parallel to the load axis as shown in Figure 9.3. If it is less rigid, the load will
drop to B 1, at a larger axial shortening than A, as shown by the dash-dot line AB 1, the slope
of which depends on the elasticity of the test fixture. One may note that in an actual test,
curves AB or AB 1 may not be entirely linear when the displacement control is too slow to
follow the rapid drop in load. At B (or B 1) the cylindrical shell has assumed its first stable
postbuckling equilibrium state.
If now the axial shortening is decreased, the stable postbuckling equilibrium becomes
unstable and the cylindrical shell jumps back, in some steps, along the dashed curve BE (or
a similar curve B 1E, not shown in the figure) to its unbuckled pattern, at a load E below the
initial buckling load. Then it returns to 0 along the original line EO.

LOADING
SCREW JACK
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

END PLATES

SHELL

MEASUREMENT OF

MEASUREMENT

Figure 9.2 Schematic view of a typical deformation controlled test rig for axial compression (from
[9.72])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
632 Shell Buckling Experiments

LOAD F

p I
l
f.:

[\
/! \
E j : \
.,; ! \
"" .
G ~~__...-:.
8 e,
c
u.....-----
o o,

AXIAL SHORTENING

Figure 9.3 Typical buckling process of a cylindrical shell under axial compression. Note the large
difference between the experimental buck ling load (A) and the predicted one (f) for a
perfect shell

If on the other hand, after the shell has snapped 10 B (or B, ). the axial shortening is
increased, the load carried by the cylindrical shell rises slightly unti l, at a second critical
value of the axial shortening, the first postbuckl ing pattern becomes unstable and jumps.
along CD (or CD,), into another two-tier pattern wi th the number of circumferential waves
reduced by one. This process of "secondary buckli ng" will repeat itself if the axial shortening
is further increased, as long as the deformations remai n elastic.
The theoretical curve for the corresponding perfect cylindrical shel l is shown in Figure
9.3 as a dotted line OFGB. which then joins the experimental postbuckling curve BC. As
pointed out in Chapter 2, the main cause for the large difference between the predicted
buckling load F for the pcrfccl shell and thai observed in a typical lest A is lhe unslablc
initial postbuckl ing behavior of the shell and the rcsu lling impelfection sensitivity. Further-
more, the buckling pattern predicted by class ical linear theory for a perfect shell is a chess·
board pallern shown by a drawing in Figure 9.5 (reproduced from [9.65]), wh ich differs
considerably from lhe two-tier postbuckling pauern (Figure 9.4) observed in a typical test.
The theoretical chessboard pattern of Figure 9. 5 (whi ch is really for an infi nitely long shell)
represents the highly unstable initial buckling pattern and has never been observed directly
in this form in an actual experiment Bu1 high-speed photography of very care fully performed
axia l compression tests, high-speed recording of changes in the photoelastic isocli nic pat·

Figure 9.4 1ypical stable 1wo-1ier poslbuckling pauern for a cy-


--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`--- lindrical shell unde1· axial compression (from [9.77])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Buckling and Postbuckling Behavior 633

Figure 9.5 Theoretical chessboard type buckling pattern for a long cylindrical
shell under axial compression (from [9.65])

terns, and some experiments with mandrels inside the shells that restrict the buckle depth
have revealed traces of an unstable initial buckle pattern that (somewhat) resembles the
chessboard pattern predicted by classical linear theory and have clearly shown the transition
from an initial buckling pattern to another, completely different postbuckling pattern.
In the sixties, a number of experimental studies of the unstable deformation states by high-
speed photography were carried out (see [9.66]-[9.74]). At the Institut fiir Strukturmechanik
of the German Aerospace Research Establishment (DLR) in Braunschweig, Esslinger and
Meyer-Piening studied the buckling process on a number of axially loaded polyester (Mylar)
cylindrical shells with a Fastax camera having a maximum speed of 5200 frames per second
[9.70], [9.72] and [9.73]. The thin shells had a radius of 100 mm, wall thickness 0.254 mm,
and thus (R/t) = 394, and were of three lengths yielding (L/R) = 1.41, 2.25 and 3.30. The
Mylar foil had a Young's modulus of 5.4 X 104 daN/cm 2 (7.8 X 10 2 ksi) and a high elastic
limit (uyl E being nearly 10 times that of structural steel). The shells had a simple longitu-
dinal lap joint and were cast in heavy aluminum end plates.
The shells were tested in the rigid and precise deformation controlled test system RZ 100
of the DLR Braunschweig (see Figure 9.38). The axial buckling loads of the thin Mylar
shells did not exceed 3000 N (661 lbs) and therefore a smaller loading fixture could have
been used. However, the large RZ 100 system was employed because of its capability for
slower and better controlled axial deformation. For most of the high-speed photography the
slowest axial shortening of 0.039 mml sec was applied.
For the photography experiments the shells were sprayed with a silver paint, which gave
them a matte silvery tone, and were placed in the test rig between special cylindrical distance
pieces to provide space for the many lamps needed. These distance cylinders and the back-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

ground board were painted black to accentuate the test shell. The high light intensity required
for the high-speed photography was provided as shown in Figure 9.6, but it posed a severe
problem. The lamps produced immense heat, which affected the test cylinders, changing
their axial strain and elastic properties. To minimize these heating effects the following
measures were adopted:
1. The radiation of the two floodlights of 2000 W was reduced with heat filters. The 12
750 W lamps grouped around the picture section could not, however, not be shielded
since the heat filters would have projected reflections on the camera objective.
2. As already mentioned, the shells were sprayed with silver paint, which reflected most
of the radiated heat.
3. The lighting period was minimized with an automatic control that fed full voltage to the
lamps only during the brief instant that the camera was rolling at very high speed.
One fringe benefit of the heating effect was the introduction of a bias that ensured the
desired initiation of buckling on the heated side facing the camera.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
634 Shell Buckling Experiments

ARRANGEMENT OF
LAMPS f

c-$·· '' ' '

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Figure 9.6 DLR Braunschweig test setup for high-speed photography of the unstable initial buckling
patterns in cylindrical shells under axial compression (from [9.70]) a. test cylinder; b. base
plate of test rig; c. loading columns of test rig; d. camera; e. two 2000 W floodlight;
f. special 750 W lamps; g. background; h. spacing cylinder

Only about 0.3 seconds camera running time at sufficiently high speed were available for
recording the buckling process. The starting of the control system that operated the lamps
and the film had therefore to be fixed within less than 0.1 seconds. This emphasizes the
camera timing problem in photographing the rapid elastic buckling process of a cylindrical
shell, pointed out also by other investigators (for example in [9.67]). Triggering the camera
by the initiation of buckling is very difficult since as the camera takes about half a second
to reach its maximum speed, the buckling process (which, from initiation to the stable two-
tier postbuckling pattern, takes about 5-30 ms), would be completed long before that. If,
however, the shell can be buckled several times at virtually the same axial load, i.e., re-
peatability can be assured, as was the case with these Mylar shells (because of the ability
of Mylar to remain elastic for large strains), the camera can be synchronized with the ex-
pected initiation of buckling. The automatic control of the RZ 100 system here ensured
repeated identical loading speed and hence also repeatability of the load-shortening curve.
The buckling process for one of the longer Mylar shells, (L/ R) = 3.30, is shown in Figure
9.7 (reproduced from [9.73]), which presents selected frames from the high-speed film cov-
ering the unstable initial postbuckling region, from just before buckling to the stable post-
buckling equilibrium. Buckling commences here simultaneously at two points, at the bottom
end and in the middle of the cylinder, as can be clearly seen in the third frame, 0.33 ms
after the onset of buckling. Each of these buckles becomes the source of a growing unstable
buckle pattern that spreads over the cylinder. The buckles in the middle region soon pre-
dominate in this shell, as can be seen from the sixth frame (0.83 ms) onwards, whereas the
bottom buckles become weaker and eventually disappear. The buckles in this initial unstable
postbuckling state are small rounded squares (see for example in the fourth frame, 0.50 ms,
or sixth frame, 0.83 ms) and can be considered to represent the chessboard pattern predicted
by linear theory. As a matter of fact, with some imagination one can regard them as corre-
sponding to a circumferential wave number n = 18, as predicted from linear theory for
"square" buckles. Many additional similar buckles appear around the initial ones, till they
cover a large part of the cylinder surface (as can be seen for example in the tenth frame,
1.5 ms, or the eleventh one, 1.7 ms). Their shape is still roughly square, but they have
become larger than the initial buckles, and in the twelfth frame, at 2.0 ms, they colTespond
to a circumferential wave number n = 13. From this point the shape of the buckles begin
to change. They become more elongated in the axial direction, until the usual stable two-
tier diamond pattern has evolved in the twentieth frame, at 24 ms. By comparing, say, the
third frame at 0.50 ms with the twentieth after 24 ms, one observes very clearly that the
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Buckling and Postbuckling Behavior 635

0
. .
I 01 fll t~
.
G. • l ~. ~ . ,

:~,.s1
II t11 f)l fli
o,s• •.-:

t,J
A' n n tJ ~. s ~ ,' , •

,J
t~ tl !~~ - tt
, £,G • ,G . S.•

8 ·t)98
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

f) ) 1i zo l ..

Figure 9.7 The buckling process of a thin Mylar cyli ndrical shell under axial compression recorded
with high-speed photography (courtesy of Professor M. Esslinger. Shell II of [9.731). The
numbers on the pictures represent the time in milli seconds from initiation of buckling. Note
that the vertical dark strips, visible already at time 0. arc not prcbuckling deformations but
miiTOf image~ Of lhC iJ1uminatiOn sySicm

initial unstable buckli ng pattern differs very significantly in shape and size fro m the stable
postbuckling pattern reached eventually, which is visible to the naked eye after the snap-
tlu-ough.
Similar conclusio ns were reached earlier by Tennyson [9.66]. [9.68] and [9.7 1]. who stud-
ied the ini tial buckling modes of ci rcular cy lindrical photoelastic shells subjected to axial
compression. by recording the change in the 45° isoclinic pauems with high-speed cameras.
The shells were manufactured by spin-casting a liquid photoelastic plastic in a rotating
apparatus. Their (Rir) "" 100-400, (L/R)"" 2.5-4, and their thickness variation was on ly
1.5- 6 percent. They behaved completely elast ically, permitting repeatable tests (as required
for camera synchronization), and yielded buckling loads wi thin I 0-14 percent of the classical
predicted values.
In elasticity, and in particu lar in photoclasticity, isoclinics of parameter 0 defi ne the loci
of points in a stressed body whose pri ncipal stresses are incli ned at the angle 0 to a set of
orthogonal coordinate axes. Principal pl anes have zero shear stress. and employing this con-
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
636 Shell Buckling Experiments

dition and the property of isoclinics, the classical buckling shapes can be used to derive the
45• isocl inics (see [9.681), which can then be compared with those viewed through a plane
reflection-type polariscope set at 45°, in hi gh-speed pho tography (sec Figure 9.8. reproduced
from [9.68]). Tennyson showed there that the pred icted isoclinics, derived from the classical
buckling mode. resemble those observed by the Fastax camera (at 2000 frames/sec) in the
first frames. though (because of the localization of the o nset of buckling in a small region)
the observed ones arc not well defined. One can. however, deduce that the ini tial buckl ing
configuration in iL~ localized region has two circumferential half-waves and three axial hal f·
waves. In later frames. the pauern changes to a different o ne wi th larger buckles, developing
finally to five large diamond-shaped buckles circumferentially with only one axial half-wave.
This behavior is summarized in Figure 9.9, which shows the rad ial deflection pattern in the
initial and final stages, emphasizi ng again the very signi ficant difference between the unstable
init ial buckling pattern and the final stable postbuckli ng configuration that one usually ob-
serves in tests.
In addi tional st udies [9.71] Tennyson reconfirmed these conclusions, that buckli ng initi-
ation in an isotropic cyl indrical shell under ax ial compression is a very local ized phenomenon
wi th an initi al postbuckling pattern that somewhat resembles the classical one and differs
significantly from the large deflectio n postbuckli ng diamond shapes into whi ch the shell
eventually deforms.
A further confirmation was obtained by Tennyson in ex periments studying the effects of
ax isymmetric initial shape imperfections on the buckling behavior of cylindrical shells
19.208]. Tile initial asymmetric buckling mode 45 isoclinics (see Figure 9.10a) showed there
the formation of a periodic square wave pattern (the classical chessboard pattern) that de-
veloped into a postbuckli ng mode wi th a smaller number of much larger buckles (see Figure
9.10b).
A similar process was observed in another high-speed photography study carried o ut by
Almroth, Holmes and Brush at the Lockheed Palo Alto Research Laboratory [9.69]. They
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 9.8 The buckling process of a geomcllically ncar· p<:rfccl c ircular cyli nd•ical shell (R fr = 11 7
and LIR = 2.45) under axial compression. as viewed lhrough a plane refleclion polariscope.
The high speed pho1ographs of 45• isoclinics arc al 2000 frames/sec (from [9.681)

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Buckling and Postbuckling Behavior 637

8 ~SUC KLED C£NTER-LINE SHIFT INITIAL BUCKLED SHAPE


m • 3. n • 10

FINAL SUCKLED SHAPE


m• 1, n • ~

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
w(0.5,Y)

Figure 9.9 End view or


the radial deflection panern in the initial and final stages of buckling of a
cylindrical shell tmder axial compression (from [9.68])

tested very thin electrodeposited ni ckel cyli nd rical shell s. (R/1) = 857 and (L! R) = 1.67,
under axial compression. and photographed the buckl ing process o n two of their specimens
with a Fastax camera at a shutter speed of about 8000 frames/sec. The camera was also
synchronized in these tests with the expected ini tiation of buckl ing obtained from an earlier
test, 10 be repeated in the photographed one (except one test on a cyl inder that had not been
prebuckled, where the buckling load was progressively approached, with a resulting waste
of several rolls of film). In all the fi lmed tests, buckling in itiated with a small local buckle
about midway between the end plates. A series of similar small rounded buckles then prop-
agated from the initial one circu mferentially in a wedge-shaped manner. g rowi ng to a pattern
that could perhaps be considered as represent ing a chessboard pattern (frame 7 in Figure
9. 11 ). Later a different Stahle pattern of diamonds developed, each ahout twice the size of
the initial rounded buckles (sec frame 14 in Figure 9.11 ).
In o rder to achieve repeatabili ty of buckli ng tests o n thin shells. a technique of restricting
the inward buckle motion by a closely fitt ing internal mandrel was developed by Ho rton at
Stanford Uni versi ty [9.75]-[9.77 ] and concurrentl y by Almroth et al. at Loc kheed Palo Alto
(9.69 ]. A cross-section of such a mandrel in a typical test is shown in Figure 9.12. If the
max imum buckle depth is restricted to no mo re than the thickness of the shell, as in this
case. the buckli ng process can be repeated many times in the same test set up without any

(O) (b)

Fi~:ure 9.10 Buckling process of an ax isymmetric imperfect circular cylindrical shell (R/1 = 223,
Ll R = 2.82) under axial compression. as viewed through a plane reflection polariscope.
The high-speed photographs of 45• isocl inics are at 3500 frames/sec (from [9.208]): (a)
the initial buckling pauern, (b) the postbuckling panem

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
638 Shell Buckling Experiments

(O) (b)

Figure 9.11 Buckle initi atio n and progression for a very thin electrodeposited nickel cylindrical shell
(Rir = 857. UR = 1.67) under axial compression. recorded with high·speed photography
at 8000 frames/sec (from [9.69]). Only two pictures are reproduced here: (a) frame 7
showing the initial buckling pallem, representing a chessboard· like pauern, and (b) frame
14 showing the stable diamond postbuckling pauern

degradation of the shell, i.e., repeated buckling m the same buckli ng load. Even a slightly
larger radial clearance, of about tluee times the th ickness, permilled repeated buckl ing within
I percent of the buckling load (see [9.69]).
The specimens used in the earl ier S tanford University Tests [9.751 and [9.761 were elec-
troformed ni ckel shells of 0.004 in. (10.2 1-'111) wal l th ickness, 2.906 in. (73 .8 mm) diameter
and 8 in. (203 mm) length. i.e., (R/1) = 363 and (L/ R) = 5.51. Later tests [9.771 used thin-

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
walled cylindrical 7075-T6 alumi nu m shells, of nominal thickness 0.005 in. ( 12.7 I-'m),
machined carefu lly from a thi ck-walled tube in stages, the final machining being cruTied out
with the tube shrunk onto a special mandrel. In these aluminum alloy shells, (R It) = 3 13
and (L/ R) = 3.52 and their circu larity was checked to be wi thi n ± 0.001 in. (2.5 mm). In
both sets of tests the radial gap bet ween the shell and the concentric mandrel was equal to
the wall thiclmess (see Figure 9. 12), and adequate clearance at the top of the mandrel per-
mitted buckling without contact with the end plate.
In all these inward buckle restricted shells, a rounded buckle pattern appeared at buckling,
which covered part of the shell, usually one or two rows of buckles extendi ng around the
circumference of the shell. When the load was increased after each buckling, as was done
in the case of the nickel shells (see [9.75]), the buckles spread row after row until the shell
was entirely covered with the rounded pallern shown in Figure 9. 13. T his pallern can be
considered to resemble somewhat the theoret ical chessboard pattern of Figure 9.5. It is

ST'EEL EN0 PLATE

70 76- T6
CYLINDRICAL.
,_~

ALU t.ONUM
MANDREL

ENCASTRE
ENOS

f igure 9.12 Cross-section of a 707S-T6 aluminum alloy cyli ndJical shell in the text fixture, showing
the internal mandrellhal res1ric1s inward buckle motion (from [9.77])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Buckling and Postbuckling Behavior 639

Figure 9.13 Buckles covering the entire surface of an axially


compressed thin nickel shell, which has a close-
fining internal mandrel (from [9.76], courtesy of
Professor W. Horton)

someti mes also inexactly referred to as the "diamond pauern" (for example in [9.65)), though
it is sti ll a long way off the usual two-tier, sharp-edged diamond pattern that appears in the
stable postbuck ling state (Figure 9.4) of axially compressed cyl indrical shells when there is
no restriction on inward buckle motion.
T he important conc lusions from the high-speed photography studies and experiments with
restricted buckle depth are therefore that buckling initiates in a local unstable rounded pattern
that somewhat resembles the classical chessboard pattern bu t that, after propagating circum-
ferentially, transforms into an entirely different stable postbuckling pauern of diamonds of
about twice the size of the initial rounded buckles.
One could also interpret the buckling process in the following manner, as proposed by
Esslinger and Geier (9.74]: The chessboard pattern (Figure 9.14a) with which buckling ini-
tiates is of infini tesimal amplitude and hence cannot be detected by the naked eye. As the
buckles deepen so that they can be seen by the naked eye, the chessboard panern transforms
to a rounded diamond pattern (Figure 9. 14b) forrned by the deepening and growing of the
inward buckles, with a simu ltaneous shri veling of the outward buckles to narrow ridges.
These ridges move obliquely and thus give the rhombic appearance. Their intersections are
the centers of the initial outward buckles. This rounded rhombic pattern is unstable and
finally transforms into the common stable postbuckling pattern of larger sharp-edged dia-
monds.
Tn passing. it may be pointed out, as Fung and Sechler did in [9.78], that on physical
grounds the development of the sharp-edged diamond pattern in the deep postbuckli ng region
of thin cylindrical shells is very plausible. For
[l]n a cylindrical shell the bending •·igidity is proportional to 1-' (I = wall thickness) whereas the
rigidity against membrane extension is proportional to 1. Hence, when 1 is very small, the resistance
to extension is very much larger than that to bending and. in order to seek a lower strain energy
configumlion. 1he cyli nder wiiJ take a form which involves liHie sLreLching of t.he mid-surface.
Since a cylinder is a developable surface and a diamond panem formed by a regular arrangement

Figure 9.14 Esslinger and Geier"s interpretation of the initial


buckling process in a cylindrical shell under axial
compression: (a) the chessboard pattern, (b) rounded
lol (b)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
diamond pattern
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
640 Shell Buckling Experiments

of flat triangular surfaces is also developable: with proper dimensions and ideaJJy sharp ridges a
cylinder can be transformed imo the diamond pauem with no membrane extension.
Yoshimura pro,•ed this in 1951 [9.79] and its Engl ish translation, [9.80], and the developable
polyhedral surface (see Figure 9.1 5) that thi n cyli ndrical shells approach in the deep post-
buckling regime is called the Yoshi mura buckling panern .
.. In a real cylinder, the bending is concentrated around the ridges and, as a consequence,
a small amount of stretching is introduced. The thinner the cylinder wall. the stronger is the
tendency to form such a developable diamond pauern" [9.78].
The onset of buckling in an isotropic cylindrical shell under axial compression is a local
occurrence and the buckling loads are therefore independent of the length of the shell (except
for very short and very long shells). whereas the mi ni mum postbuckJing loads. and the
correspondi ng postbuckling panerns, are strongly length dependent. On the other hand , initial
geometric imperfections and disturbances will significantly inll uence the onset of buck.ling,
and hence the buckling load , but wi ll hardly atlect the minimum stable postbuckl ing load.

9.2.2 Influence of Rigidity of Test Machine


Before we leave this discussion of the physical events in an ax ial compression experiment
on a thin cylindrical shell. some comments are warranted on the inlluence of test machine
rigidity on the buckl ing load. As mentioned in Chapter 3, von Karman and Tsien concluded
in their analysis (3. 1) that the elasticity of the testing machine should have a signi ficant
effect on the snap-through process, and therefore on the buckl ing load , and Tsien proposed
an energy criterion that would assess the influence of the testing machi ne [7. 15]. However.
nearly all tests in very rigid and in very clastic testing machines lead to the same buckling
load. In the six ties this problem was studied extensively (see [9.65), [9.75], [9.76] and [9.8 1)-
(9.83]). Already early experiments on spherical shells and the preliminary tests by Horton
et al. in 1961 [9.81] indicated the absence of a clear difference in the buckling load obtained
in a rigid or clastic test machine. Four of these experiments were conducted in testing
machines whose spring constants could be measured reliably. Then in 1963 Mossakovski i
and Smelyi (9.82 ( reported a significant difference, of about 12.5 percent, in the buckling
stress between their "soft" and "hard" machine. They tested 34 cylindrical shells in three
testing machines (whose rigidities were, however, not determined quantitatively) and claimed
that Tsien's energy criterion gave the correct influence of the testing machine. They were
the only investigators who repotted such agreement. but more recent calculations (see [9.83])

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 9.15 Yoshimura buckling panern


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Model Fabrication for Isotropic Shells 641

show that their experimentally observed difference is only a quarter of that which would be
predicted with Tsien's criterion, if it were valid. In their experiments in 1964, which included
the high-speed photography already discussed, Almroth et al. [9.69] tested their high-quality
electrodeposited nickel specimens in a testing device that could be operated as a "dead
weight" (i.e., extremely "soft") machine or a "rigid," hard one and found no difference in
the buckling load obtained for rigid or dead weight. Though the large amount of inertia in
their system limits their dead weight loading, their results strongly reinforce those of [9.81],
invalidating the Tsien criterion. Later careful experiments by Horton and Bailey on steel and
aluminum shells [9.77] and by Babcock on electroformed copper shells [9.83] provide un-
qualified evidence for this conclusion. Horton and Bailey tested 100 beverage cans whose
automated manufacture ensured geometric and material consistency, 50 in a hard 30 ton
universal testing machine and 50 in a soft machine, obtained by insertion of a leaf spring,
after the stitiness of both test rigs was measured. They also carried out multiple tests on
machined aluminum shells with a closely fitting mandrel, described earlier, for different
measured testing machine stiffnesses obtained with the aid of leaf springs. They concluded
unequivocally that testing machine rigidity has no influence on the initial buckling load or
buckling mode. Babcock [9.83] tested eight high-quality thin electroformed copper shells of
(Rit) = 760-1160 (having thickness variations of ±2 percent and very low shape imper-
fections) with three types of loading systems. One was a rigid controlled displacement type
machine, one was a soft system, obtained with a simple loading ring, and one was a very
soft system, obtained with a pressurized flexible tube inserted in the end ring. The three
loading systems had stiffness ratios of 555:42: l. Again the buckling load was found to be
entirely independent of the stiffness of the loading device. A similar conclusion was arrived
at for shallow spherical caps in [9.223] and for complete spherical shells in [9.84], to be
discussed later.
One should, however, remember that whereas the buckling load is independent of the
stiffness of the loading device, the postbuckling behavior depends on this stiffness, as was
shown for example in Figure 9.3.

9.3 Model Fabrication for Isotropic Shells


Before describing typical test setups and procedures for isotropic cylindrical shells under
axial compression, one has to consider model fabrication, which for shells is of major im-
portance because their buckling behavior is often very sensitive to details, in particular initial
imperfections. The discussion will focus here on cylindrical, conical and spherical shell
models, but where the techniques apply also for other types of shells, even stiffened shells,
these too are included. As pointed out by Babcock [9.58], the main problem with shell model
making is that models must be made so that the difference in buckling loads and behavior
from one nominally identical specimen to another must be less than the parameter in the
experiment under investigation. This led to the quest for much better shell models, which
hopefully would reduce the large scatter exhibited by tests for decades (see for example
Figure 9.16).

9.3. 1 Electroforming
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Electroforming was one of the methods for making better shells favored by many investi-
gators in the sixties. It was introduced to shell experiments by Thompson of University
College, London, in 1960 for spherical shells made of copper [9.85], and then applied to
cylindrical shells in 1962-64 by Babcock and Sechler at Caltech [9.86], who used copper,
and by Sendelbeck at Stanford University [9.87] and Almroth et al. at Lockheed Research
Laboratories, Palo Alto [9.69], whose specimens were made of nickel. In electroforming,
sometimes also called electrodeposition, the shells are plated on either reusable or fusible
mandrels (usually wax sprayed with an electrically conducting silver paint). The quality of
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
642 Shell Buckling Experiments

Ptl<Pipcl

i.O'-,~~--'ClASSICAL THEORY

o.ar- •

0,6
I
. I
• •

:-.::a,
• e • LOWER BOUND OF
TEST RESULTS I
0.4til.ft. ·. '/.
0.2
\!":-'
... ,.,./.
* • ...,. ·, /.'· "'
.. ' .
• •:-..~'.•"/ ' '

Figure 9.16 Scatter of test data (from various investigations) for cylindrical shells subjected to axial
compression (from [9.116])

the shell is that of the mandrel, but plating is usually accompanied by initial stresses in the
plated material that tend to relieve themselves when the shell is removed from the mandrel
and cause imperfections. The magnitude of these initial stresses depends on the plating bath
conditions and can be reduced by the addition of plating additives, which may also increase
the relatively low proportional limit of the plated model. The plating additives are usually
proprietary and sometimes mysterious (for example, addition of black molasses to a copper-
fluorborate bath or of saccharine to a sulfamate nickel bath have been reported to be useful),
which makes electroforming a trial and error process or somewhat like black magic. The
electroformed specimens were indeed better shells (with smaller geometrical imperfections),
which when tested yielded significantly better knock-down factors, as can be seen for cylin-
drical shells in Figure 9.17 (from [9.58]) when compared with Figure 9.16 (reproduced from
[9.116]) or with the data range of Weingarten et al. [9.103] shown in Figure 9.17, but the
scatter was not changed significantly. Some of the many buckling experiments with electro-
formed shells carried out in the sixties are reported in [9.75], [9.86]-[9.88], [9.93], [9.94]
and [9.97] for cylindrical shells, [9.91], [9.92] and [9.95] for conical shells and [9.84], [9.85],
[9.89], [9.90], [9.96] and [9.98] for spherical shells, with others cited in Babcock's survey
[9.58]. Only a few of these papers and reports give details of the electroforming process
used (for example [9.85], [9.87]. [9.88], [9.89], [9.91] and [9.97]). Since the high expecta-
tions regarding the quality and results of the electroformed specimens did not materialize,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

they had practically disappeared by the early eighties, except for spherical shells (for example
[9.98]).
In recent years there has, however, been a revival of clectroforming. At the very active
shell buckling research center of the French Institut National des Sciences Appliques (INSA)

Perl Pel
b. BABO:X:K a
~
SECI'iLER, 1963
1.0 ¢' HORTON a OURHAII, 1963
V ALMROTH, HOLMES 8 BRUSH, 1964

~08
X. HORTON Bl DURHAM, 196!5
X +
+ SENDELBECK a
CARLSON, 1967
\.
" 'V~ a o
0 BABCOCK, 1967
!5 x ' ',......,
+ o<:g o 0 ARSOCZ, 1966

f6
IXl

rn
0
8 + +
~
;:-- ........ 0
""·-----
• HORTON &.CRAIG,I969
DATA RANGE, WEINGARTEN,
MORGAN 8: SEIDE, 1960

~0.4
u
0
""'
g02
"'z
:J 0 '==~"'=-~!;-;;---:;±o;---,:!=:--::!=:--:-!:-.;:­
g 200 400 600 600 1000 1200 1400
IXl RADIUS 10 THICKNESS RATIO, Rlt

Copyright Wiley Figure 9.17 Axial buckling load for electroformed cylindrical shells (from [9.58])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Model Fabrication for Isotropic Shells 643

in Lyon, Jullien and his students have developed techniques for electroforming high-quality
copper and nickel shells and employed them extensively in their experimental programs of
shell buckling under different loadings [9.99]-[9.102]. Electroforming was first employed
(see [9.99] or [9.100]) to obtain nickel cylindrical shells with accurately predetermined ge-
ometrical imperfections, both modal imperfections (periodic and local) and industrial im-
perfections (typical fabrication dents). This was essentially a modern development of earlier
similar studies with specifically designed imperfections [9.86] or [9.91]. The accurate near-
perfect reference cylinders, of (Rit) = 375-625, carried buckling loads of 80 percent or
more of the classical one, and the scatter in the artificially imperfect cylinders was small
enough to permit discernment of the influence of the various geometrical imperfections, as
can be seen in Figure 9.18 (reproduced from [9.99]). Later, electrodeposition was used to
obtain corrugated nickel shells, consisting of a series of truncated cones and short cylinders
[9.101], with (Rit) = 410-440, yielding either concave or convex axisymmetric "corruga-
tions" (a "slim" or "wide" hip). The machined mandrels upon which the specimens were
electrodeposited were made of duraluminum, which was dissolved chemically (with caustic
soda) for separation from the shell.
Jullien and his coworkers also developed electrodeposition of stiffened shells. In 1985
[9.99] nickel cylindrical shells were electrodeposited on an aluminum mandrel that had thin
steel rings inserted between aluminum rings. The crystalline structure of the nickel deposited
onto the steel rings approaches epitaxis, and thus an excellent bond is formed. After the
aluminum mandrel was dissolved, ring-stiffened shells of (Rit) = 440 were obtained. More
recently they have electroformed good quality ring- and stringer-stiffened copper shells by
MODAL INOUSTRI AL
IMPERFECTIONS IMPERFECTION

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
NEAR PERFECT LOCAL LOCAL PERIODIC
CYLINDER CONCAVE' CONVEX SINUSOIDAL

~R~ :tiJ~ ~~~~ICAL


fD ZL17~-.~ 7 l~ Jq
1

~~'~Z~m
.

UJ rev
_J_ A 2L/ --- _;___ "
l\j) r~·"25-70

-~no ~~ {G]""'o f.··~•


0

W '"

~Att"' I 1-2.2 :Ef6·'


ll
L
lQ~. --WELD-TYPE
1

•,,~16->5

10 --------------------------------- '

-0
b .' •
' i\ ',
b i . ""-¢....---~
fE

• "'
06 \• --- • "' "'
I "- - - - -<:f>----- - - - ---<>--·
I -~ •
o.t' \ ... ~-.:_~
·------
0.2_

0
o':---~'--:!-'----,3~
1 _ _4.LI--~~ _6L_A-/t--~---------;;--

Figure 9.18 Int1uence of various geometrical imperfections on the buckling stress of electroformed
nickel cylindrical shells subjected to axial compression (from [9.99], courtesy of Professor
Copyright Wiley J.F. Jullien)
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
644 Shell Buckling Experiments

a similar technique [9.102). The copper shells were electrodeposited on aluminum mandrels
in which copper stringers and/or rings had been prepositioned and bonded temporarily (see
Figure 9.19).
The work of Jullien at INSA has brought electroformi ng back as a significant method of
model fabrication. but it remains a quite difficult method , which requires high skill, expe-
rience and some art to produce accurate specimens with uniform thickness and material
propenies. The difficulty is demonstrated in Figure 9.20 (reproduced from [9.84]). which
shows the normalized buckling load of complete electrofonned spherical shells as a function
of specimen number. As Babcock already noted in 1974 [9.58[: " [Tjhe lesson from this is
fairly obvious; it takes even a ski lled person a considerable amount of ti me to be able to
produce and test consistently good specimens." Remembering that the "skilled person" mak-
ing the specimens of Figure 9.20 was R.L. Sendelbeck from Stanford University, a real ex pelt
in electroforming, and that the " lesson" was poi nted out by an outstanding experimental ist
C.D. Babcock, who was also experienced in electrofotming, underlines the warni ng. The
authors, who have also had some experience with this method. concur with Babcock's note
of caution. But for the skillful and patient experimenter the method has great potential, as
can be seen at the right-hand side of Figure 9.20 (spherical shells buckling at 80 percent of
the load of perfect shells!) or in the results of Jullien and his co-workers. The newcomer is
advised, however, to study carefully the trials and tribulations of his predecessors mentioned
above.

9.3.2 Mylar Shell Specimens


Another method initiated in the sixties was fabrication of shells from Mylar and other similar
polyester films. primarily on account of the high (uJ E) values, by an order of magnitude
larger than that of the usual structural material. which facilitates postbuckl ing experiments
(as discussed in Chapter 5, Section 5.6). The method is most suitable to developable surfaces,
such as cyli ndrical or conical shells. Mylar specimens are inexpensive and can be buck led
deep into the postbuckling region many times without noticeable degradation of the shell
quality, due to its high (uJ E) value = 0.016. compared to 0.0022 for medium steel). In the
sixties Mylar specimens were widely employed (see for example [9.103)- (9.1071 and others
cited in Babcock's survey [9.58)). but one disadvantage found in My lar was its anisotropic
material propenies. For example, as much as ± 15 percent variation in tensile modulus,
depending on the orientation of the specimen wi th the full sheet axis, was observed [9.1 08)
but was ignored by most investigators.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Much effort was invested in the seventies in making Mylar specimens more accurate and
the bonding at the seams and the mounting in the end plate more precise. For example,
Yamaki and Otomo [9.1 09]. who used a simi lar polyester film (Mitsubishi Di afoi l no. 250),

Figure 9.19 Elcctrodcposited stringer-stiffened copper shell fa b-


ricated with the techniques developed at INTA Lyon
(counesy of Professor J.F. Jullien)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Model Fabrication for Isotropic Shells 645

oaf-
0
' rP 0 0
0.61- 0 0
0 0
0
0 0

::L-
0

0 0°
000
0o 0
10 --2""'oc-----io--+.4o----'50·
SPECIMEN NUMBER

Figure 9.20 Test performance of electroformed nickel spherical shells-a "learning curve" (from
[9.84])

checked the degree of anisotropy of Young's modulus by mirror extensometers, parallel and
perpendicular to the rolling direction, and found about 6 percent difference. Similar mea-
surements of Poisson's ratio with resistance strain gages also revealed very small differences.
They therefore decided to regard the material as isotropic and used it extensively in these
and later postbuckling tests [9.110] and [9.111].
At the DLR (formerly DFVLR) Institut fiir Strukturmechanik in Braunschweig, Mylar foil
has been used extensively for cylindrical and conical specimens [9.112]-[9.115] or [9.248]
and others referred to in Singer's survey [1.24]. Geier and Heidemann [9.115] carefully
measured anisotropy and thickness of the Mylar foil. The anisotropy did not exceed 12
percent and the thickness variations about 5 percent. Mean values were calculated and the
variations were lumped together as foil tolerances. The total scatter these tolerances could
cause in buckling loads, being evaluated at 6.8-20.7 percent, was found to be smaller than
the scatter of the measured buckling loads. An equivalent isotropic modulus was also esti-
mated for external pressure and axial compression. After these preliminaries, Mylar was then
regarded also by the DLR investigators as isotropic and uniform in the continuation of their
work.
Indeed, the quality of Mylar and other polyester foil (like Melinex, Diafoil, etc.) improved
considerably since the large variations of ± 15 percent in modulus found in 1968 [9.108].
But one should note that, for example Melinex films, employed for cylindrical shells in some
studies, can be nearly isotropic, such as the Melinex Type 226 used by Tooth and Fernandez
[9.117], or have marked orthotropic properties, such as the film used by Toh and Spence
[9.118], which exhibited an orthotropy of more than 30 percent. Caution is therefore advised
regarding the directional properties of the polyester foil to be used in an experimental pro-
gram.
The users of Mylar specimens, in particular those at DLR Braunschweig, have continu-
ously striven to obtain more perfect shells. One aspect of the manufacture in this direction
is the care taken to utilize the slight curvature of the rolled foil in forming the curvature of
the cylinder or the cone, and thus to minimize the initial circumferential bending stresses
that might be imparted in the process. Other aspects are the bonding of the seam (with a
joint strip for the thinner foils) in a special double-wall polished steel cylindrical die, using
suction to ensure adherence of the Mylar specimen to the circular die, and careful precise
milling of the ends in the same die (see [9.115] for details). Special care is also taken in the
placement of the shell in the resin-filled grooves of the end plates. For example, one rec-
ommendation is to lower the Mylar specimen till it touches the bottom of the groove and
then raise it again by about 0.1 mm, leaving a thin film of resin, which should yield more
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

uniform axial loading.


With the improved quality of Mylar and similar polyester films, these foils have become
today the favorite of many investigators for buckling and postbuckling studies of thin shells
with (Rit) = 400-1000. They are now commonly used for specimens with developable
surfaces (for example, [9.119], [9.120] and the references quoted above). Mylar has also
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
646 Shell Buckling Experiments

been used to model tanks with variable wall thickness because it can easily be laminated
and the laminated sheets wrapped round a specially prepared wooden form [9.121].
One should point out that the aim of experiments with Mylar models is primarily to study
elastic stability phenomena and not to obtain design information, which would be distorted
by its very high (uJ E), unrealistically far from that of practical construction materials. Also,
there are other ways of making near perfect, or intentionally imperfect shell models, as will
soon become evident.

9.3.3 Thermal Vacuum Forming


Another widely used method of fabricating plastic models of shells is thermal forming, also
developed in the sixties and seventies. In this process the plastic material, which is usually
flat, is heated and made to conform to a mold. PVC (polyvinyl chloride), polyethylene,
Lexan, PMMA (polymethyl methacrylate), solid urethane plastic (SUP) and some epoxy
resins (such as Araldite) have been used.
Some typical examples for PVC spherical caps, elliptical and torispherical shells are the
domes tested in the mid-sixties at the McDonnell Douglas Corporation [9.122] and at the
U.S. Army Watertown Materials Engineering Laboratory [9.123]. Both programs employed
thermal vacuum forming. Adam and King of McDonnell Douglas [9.122] described the
fabrication process in detail. Their specimens were made of a plasticized PVC material
(Union Carbide Bakelite), a mixture of 86 percent vinyl chloride and 14 percent vinyl acetate,
that yielded a more ductile material than the more common rigid polyvinyl chloride sheet.
This plasticized PVC could already be formed at 160°F (7l C), compared to the 275°F 0

(135oC) required for unplasticized PVC. Test of coupons perpendicular to and in the direction
of rolling showed the material to be nearly isotropic, with an average E = 465,000 psi (3, 170
MPa), and a high (u) E) "=' 0.012 (not far from that of Mylar). Similar values were obtained
for the polyvinyl chloride sheets used for torispherical shells in [9.123] and for spherical
caps in [9 .124]. The domes were vacuum thermoformed into female 6061 aluminum molds
and, without removing the domes from the molds, their inside surfaces were machined to
produce the required uniform wall thicknesses.
The dome forming setup is shown in Figure 9.21, which exhibits the aluminum mold with
its perpendicularly drilled small vacuum holes. The vacuum forming method was modified
by adding external pressure, as can be seen in the figure. This addition helped forming the
rigid vinyl into the sharp radii, as in the case of shallow elliptical and torispherical shapes.
In the fabrication process, the PVC sheet was clamped on the mold with a clamping ring,
to which then the external pressure plate was bolted, using an "0" ring for the pressure seal
(see Figure 9.21). The assembly was placed in a mechanically convected oven and soaked
at a temperature of 240oF (ll6°C) for at least 6 hours to assure a uniform temperature
distribution. The vacuum pump was then turned on, pulling the PVC sheet down to conform

CLAM~ING
RING

VACUUM

Figure 9.21 Vacuum-forming setup for PVC spherical domes (from [9.122])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Model Fabrication for Isotropic Shells 647

to the mold cavity. With the plastic sheet being held down by the vacuum produced pressure,
the external pressure (15 psi) was applied to the surface of the plastic shell. The assembly
was then cooled down slowly (for a minimum period of 8 hours) to room temperature, the
slow cooling preventing formation of residual stresses in the domes. The external pressure
plate was then removed and the mold-dome assembly was installed on the lathe base plate,
the PVC dome being held to the mold cavity by the vacuum (kept through a rotating pressure
joint) also during the machining of the internal surface of the shell. After the finished domes
were checked for wall thickness variations with a dial indicator reading to 0.0001 in. (0.0025
mm), they were fixed in a grooved plexiglas ring with a Hyson 2038 epoxy cement (with
hardener "'C"), providing nominal fixed-edge support.
All the aluminum molds and PVC domes were machined on the same lathe-and-baseplate
combination, eliminating concentricity problems and reducing the cost of tooling. The only
tooling necessary for a PVC dome of one particular shape was the female mold and its
corresponding tracing template, which the hydraulic tracing attachment of the lathe follows.
By changing the diameter of the tracing stylus to compensate for the dome thickness, the
inside surface of the plastic dome could be machined with the same template that was used
for the mold, eliminating the effect of template differences on the shell wall thickness var-
iations. To reduce the influence of the different coefficients of thermal expansion, the machine
room temperature was kept constant.
With this fabrication process Adam and King obtained precision domes with very small
thickness variations of ± 0.00025 in. ( ± 0.006 mm), less than ± 1 percent of the thickness,
at minimal cost. The spherical cap specimens of (Ron,) t) = 245-357 appeared to be of high
quality, yielding consistent high buckling coefficients of about 80 percent the classical value.
The first domes of the series had larger thickness variations, up to ± 3 percent of the thick-
ness, but still yielded similar consistently high buckling coefficients, indicating that thickness
variations are not a critical imperfection in comparison to shape imperfections (variations in
the radius of curvature). The fabrication process has been discussed here in detail because
it presents a good example of careful planning of the manufacture of shell specimens. Though
today the machining would be done with more sophisticated computer-controlled equipment,
the planning considerations remain valid.
The fabrication technique for PVC specimens employed by Adachi and Benicek [9.123]
for their torispherical shells was similar to that just described, as were also those used in
[9.124]-[9.126] for spherical caps. One may note that though the specimens of Wang et al.
[9.124] were also of high quality, they were somewhat less perfect than those of [9.122],
yielding buckling coefficients of 70 percent of the classical, or less for the relevant fixed
ends. This may have been due to their larger (R It) = 300-600 or to the use of Ultracal
plaster molds instead of the aluminum ones for the vacuum forming. Also, Kruger and
Glockner's spherical caps [9.126], with (Rit) = 200-500, showed more scatter and buckling
coefficients, usually about 60-67 percent of the classical for clamped ends, though their
fabrication process closely resembled that of [9.122]. On the other hand, Tillman's spherical
caps [9.125], manufactured using a very similar technique, appear to be of comparable
quality to those of [9.122].
A few years later, high-precision experiments on the buckling of spherical caps were
carried out at the Institute of Space and Aeronautical Science of the University of Tokyo
[9.127]. After detailed examination of several common processes for fabrication of shell
specimens (spinning, air vacuum forming, hydroforming, machining, electroforming, explo-
sive forming), Sunakawa and Ichida decided to use thermal vacuum forming of flat PMMA
(polymethyl metacrylate, i.e. perspex) plates in a manner similar to that employed in [9.122]
for PVC. Their molds, shown in Figure 9.22a, again use both vacuum and external pressure.
The temperature and forming pressures were carefully measured and controlled during
the forming process, as seen in Figure 9.22b. The resulting specimens, covering a wide
range of geometries, (R It) = 100-3000, and of shell segment geometrical parameter,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
648 Shell Buckling Experiments

~~---

(a J

2
uJ
150
a::
:::l
l:i 100
a::
uJ
ll.
~ 50
uJ
1- I
0
0 2 3 4 5 6 7 B
TIME lhrl

(b)

Figure 9.22 Mold and forming procedure for PMMA spherical caps (from [9.127]): (a) mold used in
forming, (b) an example of the forming procedure

A(== 2[3(1 - v 2 )] 114 VHTi) = 2.64-22.5, were indeed nearly perfect, with thickness varia-
tions within ± 1.5 percent of the mean and yielded buckling coefficients of about 80 percent
the classical for A < 13, decreasing for larger values of (see Figure 9.110).
More recently, PVC hemispheres with an axisymmetric thickness variation for imperfec-
tion sensitivity studies were also made by the vacuum thermoforming process (see [9.128]),
and this variation was obtained within 6 percent for most of the specimens produced.
Like Mylar specimens, thermally formed plastic shells can be repeatedly buckled, on
account of their relatively high (aJ E) values. Some investigators however, restricted the
buckle depth to 2-3 shell thicknesses with a mandrel, for better repeatability of buckling (as
for example in [9.122]). One should note that dynamic loading of plastic specimens warrants
some caution since fracture failures have occurred, for example on arches at Caltech (see
[9.58]) or in vibration tests at the University of Calgary, Alberta, Canada [9.129], and in
this case the more ductile variants of PVC or similar plastics are to be preferred.

9.3.4 Spin-Casting and Other Thermoforming Processes for Plastic


Models
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Another technique for making cylindrical shell specimens is spin-casting, developed by Ten-
nyson at the Institute for Aerospace Studies of the University of Toronto in the early sixties
[9.130]-[9.132] and [9.66]. The test cylinders were made from a liquid photoelastic epoxy
plastic, spun cast in an acrylic tube. The tube was attached to a lathe, which produced
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Model Fabrication for Isotropic Shells 649

sufficient rotational energy to spin-cast the polymer inside the shell form. The plastic used
was a photostress epoxy resin (type A), in which resin and hardener were mixed in a ratio
of 85 parts to 15 parts. A room temperature curing time of approximately three hours was
necessary. Casting of the epoxy had to be done in a condition of low relative humidity or a
cloudy condition permeated the plastic during its curing.
Because of irregularities in the wall thickness and internal diameter of the acrylic form,
mass balancing was necessary to prevent vibration of the apparatus during spinning (which
would produce thickness variations in the circumferential direction). To ensure a circular
cross-section in which to cast the photoelastic shells, a Hysol plastic liner shell was first cast
in the acrylic form. High rotational velocities on the order of 800 rpm were necessary in
order to distribute the liquid evenly and yield a precise circular shape. The centrifugal force
also drove out the air bubbles, which were trapped in the liquid and thus produced a smooth
imperfection-free shell wall. The transparent acrylic tube also permitted detection of any
remaining trapped air bubbles, which could be removed by local short-time heating. Careful
alignment of the form yielded shells with thickness variations of usually less than ± 4 per-
cent, and ± 6 percent at most. (More details of the technique are given in [9.130] and
[9.131]).
The process produced nearly perfect shells, with (Rit) = 98-443, which under axial
compression yielded buckling loads of 86-90 percent of the classical one. These epoxy shells
exhibited linear elastic buckling behavior and hence permitted many repeated tests without
any apparent degradation (each shell was tested as many as 20 times with repeatable results),
which made them also very suitable for combined loading tests, from which interaction
curves are obtained (see [9.132]). Their photoelastic properties also provided another means
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

for studying their postbuckling and buckling behavior (sec for example [9.68] or [9.71]).
A similar spin casting technique was also used by Gorman and Evan-Iwanowski at Syr-
acuse University Applied Mechanics Laboratory [9.133], to fabricate photoelastic cylindrical
shells, of (Rit) = 157, for study of prebuckling deformations.
For doubly curved shells of revolution, other types of thermoforming methods were de-
veloped. For hyperboloid shells, such as natural draft cooling towers, Veronda and Wein-
garten at the University of Southern California developed a hot pressure molding technique
for PVC specimens [9.134]. In the process, a wooden male mandrel in the shape of a
hyperboloid of revolution was first machined on a lathe with a hydraulic tracer accessory
following a precise template. With the mandrel, a split female plaster of paris mold was
fabricated, the meridional split in the mold facilitating the removal of the formed shells. The
specimens were made from cylinders of PVC sheet, bonded along a longitudinal seam. These
cylinders were then inserted in the female mold, sealing flanges were attached at the ends
and the whole assembly was heated to 185°F (85oC) in a hot air circulation oven. When the
PVC softened, it was blown outwards onto the mold by compressed air at a pressure of 5
psi. After cooling, the sealing flanges were removed and the PVC hyperboloid shell was
withdrawn from the mold and accurately trimmed to length. The specimens were carefully
mounted in the groove of a clamping fixture, which was filled with cerrobend (a low-melting
alloy), which, upon cooling, clamped the bottom ends of the shells while their top edges
remained free.
Though, as in other thermoforming processes, it was difficult to maintain constant thick-
ness, accurate micrometer measurements indicated thickness variations of not more than ± 2
percent. Young's modulus for the PVC was determined for each formed specimen from its
load-deflection curve, the average being E = 460,000 psi (3,200 MPa), similar to the values
measured in [9.122] and [9.123 ]. For reference, three cylindrical shells were fabricated by
the same technique as the five hyperboloidal ones. Comparing these cylinders of (R/ t) "=
200 for axial compression with the Mylar cylindrical shells of Weingarten et al. [9.103]
showed here a buckling coefficient of 74-79 percent of the classical one, compared to 50-
81 percent for the six Mylar cylinders of similar (R/t) there, which indicated that the PVC
specimens were of comparatively high quality.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
650 Shell Buckling Experiments

A similar thermoforming process was employed by Mungan at Ruhr University Bochum


in Germany for his Araldite (epoxy resin) hyperboloidal shells [9.135]. The Araldite material
used was a cold-mix epoxy resin of CIBA-GEIGY CY 219 with hardener HY 977 and
accelerator DY 060, in proportions 100:50:2. Tests on tensile and bending specimens yielded
an average value of E = 500,000 psi (3,450 MPa). At room temperatures (up to 25°C) the
hardened material has a linear stress-strain curve and about 5 percent maximum recoverable
strain, a very high elastic strain capacity, which, as pointed out in Chapter 5, Section 5.9,
permits repeated testing well into the postbuckling regime. In the tests of these Araldite
shells, the strain measurements on the unloaded models confirmed that after a certain time
nearly all deformations recovered. In order to eliminate, however, the time effect (of the
delayed elasticity of the material) in strain measurements, it was necessary to record the
strains after a constant time interval, a five-minute interval being found sufficient.
The models (with maximum and minimum diameters of 600 mm and 400 mm, respec-
tively, and a height of 1200 mm) were cast between two molds shown in Figure 9.23. The
internal (male) mold consisted of 12 aluminum alloy segments, 6 of the top half and 6 of
the bottom one, and the external (female) mold had 3 steel segments. The liquid epoxy resin
filled the 5 mm spacing between the molds, beginning from the lower part and rising to the
top. The hardening of the material was completed at 20oC in a week. The external mold was
then carefully removed and the model was machined on the internal mold to the desired
thickness, yielding relatively perfect shells, though with significant thickness variations. The
models were relatively thick, (Rit) ~ 100-150, and hence were not appropriate for post-
buckling studies. The capability of the material for strain recovery was, however, utilized in
the repeated buckling tests required for the experimental interaction curves in the case of
combined axial compression and external pressure. This was after it had been verified that
repeated tests under one type of loading repeatedly yielded the same buckling load and
pattern.
A similar casting process for another type of Araldite epoxy resin, Araldite E with hardener
HY 956 in proportions 4:1, was used by Wurm at the Technical University Munich for his
doubly positive curved shells [9.136]. For the casting, a precise, disassembling internal (male)
aluminum mold and a fiberglass reinforced polyester external mold were used. Since this
Araldite is also a very effective adhesive, the molds were sprayed with Teflon to facilitate
removal of the model. After hardening (which required only one day) and removal of the
outer mold, the shell was again machined to required thickness while still on the internal
mold.

G SECTION 1-1
.
)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

TOP VIEW

Figure 9.23 Casting molds for Araldite hyperboloid shells (from [9.1351)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Model Fabrication for Isotropic Shells 651

Morton, Murray and Ruiz at the University of Oxford used a somewhat similar process
for their very thin complete spherical shells of (R! t) = 890, made of Araldite MY 753
[9.137]. They employed accurately NC machined aluminum castings for male and female
hemispherical molds, the female one being made in two parts to facilitate removal of the
finished castings. Also, here the molds were sprayed with a releasing agent to prevent stick-
ing of the epoxy shells to them, especially to the male one. The casting process consisted
of placing a weighed amount of epoxy resin in the female mold and pressing the male
against it, causing the resin to flow upwards and fill the mold. The complete spherical shells
were made from two cast hemispheres joined on an equatorial ring and supported by a
relatively thick Perspex skirt. Careful measurements showed the thickness of the shells to
be fairly uniform, within ± 9 percent of the mean thickness (and with the thickest region
occurring over a small area around the pole), and the lack of sphericity to be within 0.05
percent of the radius (or 45 percent of the mean thickness). Hence the models could be
considered to be of fairly good quality.
The same casting process was also employed by Ross and Mackney at Portsmouth Poly-
technic (now University of Portsmouth) in England for their hemispheroid domes [9.138].
These 50 shells of Rlt > 50 and of different aspect ratios (domelight/base radius), varying
from oblate hemiellipsoids to prolate hemiellipsoids, were made of solid urethane plastic
(SUP) and cast between precisely machined male and female aluminum molds, like the
hemispheres of [9.137].
For wind tunnel experiments of wind loading on models, for example of cooling-towers,
their material should have a relatively low modulus of elasticity lest the critical dynamic
pressure required to cause buckling be beyond the capability of the wind tunnel. Hence
Ruhwedel at Ruhr University Bochum chose a polyurethane elastomer, Vulkollan, made by
Pleiger in Hattingen, Germany, for his hyperboidal shells [9.139]. Vulkollan, withE = 90-
300 MPa (varying according to hardness), bridges the gap of model materials between
rubber-like materials and the relatively hard plastics, such as PVC or epoxy resins. It is very
ductile and has a failure strain of 300-650 percent, according to the type of Vulkollan, and
for strains below 2-3 percent the stress-strain diagram is linear. On account of fairly high
compressive strengths, the strength-to-modulus ratio (a) E) = 0.020 for the harder Vulkollan
50 (similar to Mylar or PVC) and 0.0025 for the softer Vulkollan 40, ensuring elastic be-
havior also in the postbuckling region for the harder types of Vulkollan.
The models were made by a casting process similar to that used for the epoxy resin shells
of Mungan, Wurm and Morton eta!. [9.135]-[9.137] discussed earlier. Both the internal and
external aluminum molds could be disassembled for removal of the finished model. The
models were also machined to the desired thickness, which sometimes varied here with
height, while still on the internal mold. Ultrasonic thickness measurements, at 198 points
for each model, showed the shells to be of good quality with thickness variations of ± 1-2
percent of the design mean thickness.
Before we leave the topic of thermally formed or cast plastic models, it may be of interest
to note that thermal vacuum forming was also employed by Knapp [9.140], [9.141] for the
models of his prebuckled pseudo-cylindrical shells, or pseudo-cylindrical concave polyhedral
shells. These prebuckled cylindrical (PC) shells are cylinders formed by the triangular poly-
hedral surface of the Yoshimura postbuckling pattern (see Figure 9.15), discussed earlier,
which sustain higher buckling external pressures than the corresponding true cylindrical
shells and hence appear to have potential for undersea structures. The PC shell models were
fabricated by a process similar to that employed for spherical caps, torispherical or hyper-
boloid shells discussed earlier [9.122], [9.123] and [9.134]. Bakelite rigid vinyl (PVC) sheet
is wrapped round an aluminum male mold, built up from accurately machined polyhedral
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,

ring segments, and is then vacuum drawn onto the mold at a temperature of 100°C (see
Figure 9.24, reproduced from [9.141], which shows the formed vinyl sheet on the mold).
Thickness variations of less than ± 3 percent were measured on the 25 models tested.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
652 Shell Buckling Experiments

Figure 9.24 1l1ermal vacuum fonning of a prebuckled pseudo-cylindrical shell model (from (9.141 1.
counesy of Professor R.H. Knapp)

9.3.5 Cold-Worked and Machined Metal Shells


Many investigators still make thei r shell specimens by cold working o r machini ng of metal.
Spinning, explosive forming and hydroforming arc commonly employed forms of cold work·
ing.
Explosive forming for spherical caps and similar closed shells of revolution was developed
in the mid-sixties at the Martin Company in Denver (9. 142 (- (9.144) for 6061 and 2014
alumi num alloy specimens. In the process, annealed (0 temper) Aat sheet was cut into square
or circul ar blanks, which were positioned on a female laminated fiberglass epoxy mold. The
soft blank was subjected to an explosive load, which pressed it into the cavity of the mold.
The formed aluminum alloy shell was then solution heat treated and water quenched to the
T4 condition and kept cold to retain this condition unti l it was resized by applyi ng an
additional explosive load in order to ensure its geometry. After resizing. the shell was arti-
ficially aged to the T6 temper by precipitation heat treatment. Then it was chemically milled
w the required thickness. By establishing the chemical milling rate and determini ng the
immersion time. reasonably close control of thickness was achieved. The geometry and
thickness of the shells were care fully measured before installation in the test rig. Though
explosive forming yielded spherical shells of fairly uniform thickness (usually within ±2
percent of the mean thickness. except for the thinnest shells, for which variations were within
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

±9 percent), it appears that for high-quali ty specimens this technique requi res a fairl y long
develo pment process.
Spinning. whi ch, on account of its relative simpl ici ty. is used extensively in practice for
fabrication of shells of revolution when accuracy of shape is not essential. was also employed
for making specimens of spherical caps and deep spherical shell s. conical shells and o ther
shells of revolution (see [9.145). [9.1 46) and o thers quoted in Babcock's survey [9.58)).
However, di fficu lties in maintaining good thickness and geometry tolerances have been en-
countered with this technique, which is highly dependent o n the experience of the operator
and may also resu lt in indeterminate residual stresses.
Hydroforming is ano ther techn ique sometimes used for metal shells specimens. For ex·
ample, Dumesni l and Nevi lle at the Universi ty of Florida made spherical caps from 3003-0
aluminum sheet by this process [9. 147]. They clamped circular blanks between a thick plate
and a female die and formed their spherical caps by forcing oi l between the blank and the
thick plate. Fairly good spec imens resulted. which were withi n the foll owing tolerances:
radius of curvature ± 5 percent of nominal th ickness and thickness :t 10 percent of nom inal,
within :t 5 percent in most cases.
An interesting application of hydroforming was the formi ng of cylindrical shells by plastic
ex pansion. carried out by Guist at NASA Ames Research Center [9. I 48]. T he specimens
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Model Fabrication for Isotropic Shells 653

were made from portions of commercially available 6 in. (15.2 em) diameter type 304 welded
stainless steel tubing, which were sealed with fixed diameter end caps and surrounded by a
precision cylindrical form. They were then subjected to internal hydrostatic pressure and
axial compression in a servo-controlled hydraulic compression test machine, which main-
tained an axial compressive load slightly less than the internal pressure load on the end caps
while allowing the tubes to shorten as they expanded. The remaining axial tensile load served
to prevent buckling. After expanding the tubes about 42 percent over the full length of the
cylindrical form to a diameter of 8.5 in. (21.6 em), which yielded a thickness of 0.0130 in.
(0.330 mm), about 16 percent less than the original one, they were removed from the form

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
and cut to length. Then pairs of clamp rings were installed at each end of the specimens, so
that the ends of the cylinders protruded 0.02 in. (0.5 mm) and only they made contact with
the loading surfaces, after being ground to precision flatness. This led to better load unifor-
mity. The initial geometric imperfections were then measured. The specimens made by this
plastic expansion process, with (Rit) = 327, were of similar quality as the electroformed
ones and upon axial loading sustained buckling loads 62-71 percent of the classical loads.
Hydrospinning, a shear spinning technique in which a flat plate is formed by shearing it
onto an accurate mandrel in a spinning process with a hydraulically pressed tool, was used
by Weller and Singer at Technion in Haifa [9.149] to form the conical blanks from high-
strength 17-7 PH alloy steel plate, which were then machined on another mandrel to ring-
stiffened conical shells. The hydrospinning represented a cold-drawing process that trans-
formed the original semiaustenitic 17-7 PH steel into a martensitic state. The shear-spun
blanks were, however, highly prestressed and required an aging heat treatment for about 15
hours at 5l0°C, which was arrived at after extensive trial and error. The result were blanks
of fairly high yield strength and about 80 percent stress relief. The machining of the conical
shells to their stiffened configuration (though it actually belongs to Chapter 13, it is briefly
discussed here for completeness) was carried out in two stages. First the inner surface was
turned and ground to fit the mandrel accurately, and then the shells were mounted on the
mandrel and their stiffening rings were turned at low cutting speed and low depth of cut, in
order to prevent high local pressure of the cutting tool. The process was more expensive
than grinding (which was tried and discarded due to the appearance of local thermal buckling
in the very thin skin of the shell) but yielded precise thin shells, with ( p.,J t) = 265-615
and thickness variations of less than 4-5 percent of the mean thickness.
Machining from thicker stock and special forgings is primarily used for stiffened shells,
but many accurate isotropic shells of revolution, including cylinders, cones and spherical
caps, as well as some more complex shapes, have been made by this method. If the radius-
to-thickness ratio is not large, the specimen can be machined without the support of an
internal support mandrel, but for thin shells a mandrel is necessary. For cylindrical shells a
thermal shrink fit, either by cooling of mandrel or heating of shell and by relying on the
difference in thermal coefficient of expansion, is probably the best method for reliability and
ease of removal.
Among the many investigators using machined specimens, those at the U.S. Navy David
Taylor Model Basin stand out for the quality of their shells and for their ingenuity. Most of
their work relates to stiffened shells, which are discussed in Chapter 13, but their machined
spherical caps [9.150]-[9.153 J, are excellent examples of very precisely machined specimens,
some with predetermined machined geometrical imperfections [9.152]. Another very recent
example of precise machining from thick stock is the torispherical shells machined from
245-mm-diameter billets of mild steel by Blachut et al. [9.154]. A computer numerical
control (CNC) lathe was used, and very good shape accuracy was obtained for these shells
of (Rit) ~ 150, with thickness variations of less than 4 percent from the average, except at
the edges of the domes, resulting in excellent agreement with elastic-plastic theory predic-
tions for perfect shells.
Other recent examples include cylindrical shell specimens machined from welded steel
tubes by Steinhardt and Schulz at University Fridericana in Karlsruhe, Germany [9.155].
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
654 Shell Buckling Experiments

They made their thinner shells, however, from thin brass stock hard soldered along the seam.
Another example is the cylindrical shells, with and without cutouts, of Strenkowski and
Witmer from MIT in Cambridge, Massachusetts [9.156], which were milled from 6061-T6
aluminum tubing. Almroth and Holmes of Lockheed, Palo Alto, California, similarly ma-
chined their cylindrical specimens with rectangular cutouts from 6061-T6 aluminum tube
stock. Other examples are an extensive series of aluminum alloy (Hiduminium 48 and 66)
cylinders with end closures of various shapes, as well as some steel spherical caps, machined
by Galletly and his associates at Liverpool University for their elasto-plastic buckling tests
[9.158]-[9.164] and similar shells made by Gill and his co-workers at Manchester University
[9.165] and [9.166].
Before we leave the topic of machined specimens, it may be of interest to note the use
of chemical milling in the fabrication of very thin conical and cylindrical shells at MIT
in the late fifties [9.289]. The shells tested there were to simulate realistic missiles with

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
(Rit) ~ 1000. Since the specimens were to have an average radius of curvature of 4-5 in.,
a skin thickness in the vicinity of 0.004 in. was required. The thinnest commercially available
7075-T6 aluminum alloy sheet was 0.020 in. thick. It was therefore decided to etch the
sheets to the required thickness (20 percent of the original one) using a solution of Ferlon
(a strong alkali) and water. The etching process was conveniently set up and proved to be
quite satisfactory, yielding a maximum variation in skin thickness of 5 percent. The physical
properties of the etched material were verified experimentally. The specimens were then
fabricated by wrapping the etched sheet around forms and joining it at a l /2 in. lap joint
with an epoxy resin adhesive. Because of the very small stiffness of the etched material, it
had to be handled with special care to prevent premature buckling before the shell was
fabricated. Finally, after the ends of the shells were squared off in a lathe, they were fixed
with Cerrobend (a low-melting-point metal) in grooves in the aluminum end plates.

9.3.6 Seamless Commercial Drink Cans


One simple method, originated by Horton at Stanford University, in the early sixties (see for
example [9.77]), for obtaining accurate isotropic cylindrical shell specimens inexpensively
is the use of seamless beer or soft drink cans. Since these thin-walled cans have to be within
the close tolerances demanded by the automated filling and sealing process, they represent
a source of specimens with accurately repeated dimensions. Recently this idea has been
revived by Arbocz and Elishakoff at Delft University of Technology in a test program related
to stochastic stability analysis [9.167] and at the Technion Aircraft Structures Laboratory in
a test program on buckling of shells under axial impact. A special production run of empty
steel beer cans without paint and without closing lids has been made, yielding specimens
with (RI t) = 300. The shells have, however, an integral bottom end closure, which has to
be removed, some axisymmetric thickness variations with height and quite significant resid-
ual stresses. These have to be studied further for more precise evaluation, but even now,
though undefined, they are practically identical for the many specimens made by the process.

9.3. 7 Realistically Fabricated Shells


Whereas the efforts of most model makers were devoted to building nearly perfect shells,
in order not to overemphasize the relative imperfections due to the small scale, a different
avenue was taken by some investigators who tried to build their models realistically by the
same fabrication technique as the full-scale shells-essentially a return to the philosophy of
some of the tests of the thirties (for example Sturm [9.38] or Wilson and Newmark [9.55])
or the fifties (for example Harris et al. [9.168]). This path was adopted particularly in civil
and ocean engineering. For example, Miller [9.169] tested 41 small-scale unstiffened and
ring-stiffened steel shells as part of an evaluation of over 700 axial compression test results,

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Test Setups for Cylindrical Shells under Axial Compression 655

the test specimens being made in a similar manner to the controlling large-scale ring-stiffened
test cylinders. Another example is the tests carried out by Ostapenko and his co-workers at
Lehigh University, Bethlehem, Pennsylvania [9.170]-[9.174], where the specimens were
cold-rolled and then welded at the joint, as customary in civil and ocean engineering practice.
Though the walls of these specimens were not very thin, (R It) = 40-124, and their buckling
was primarily inelastic, the tests were important because of the extensive investigations of
the influence of the residual stresses caused by welding.
The tests by Smith and his co-workers at the Admiralty Research Establishment, Dun-
fermline, Scotland, on tubes representing offshore and steel bracing members with simulated
damage effects [9.175] and [9.176) are examples from ocean engineering. Typical model
shells fabricated by methods used in aerospace practice are the Alclad 2024-T3 specimens
with (Ril) = 300 used in thermal buckling tests on cylindrical shells at the Technion, Haifa
[9.177] and [9.178], as well as the steel specimens used in thermal buckling tests in the
U.S.S.R. [9.179], or for example the dural cylinders, with (Rit) = 170 and 440, and with
cutouts, tested by Jung in Germany [9.180) or the steel cylindrical shells with (R It) = 570
employed in the tests in the U.S.S.R. of local buckling due to a concentrated force [9.181].
Further examples of specimens simulating shells used in tests in other industries in the
U.S.S.R. are discussed in Singer's review [1.24].
Obviously one advantage of such "more realistic" specimens is that the experimental
results can be readily applied to empirical design methods, without any doubts about material
effects.
As pointed out in reviews on shell buckling experiments [ 1.24] and [ 1.17], in the late
seventies the ocean engineering industry motivated extensive test programs on large models
of shell structures, since in order to provide reliable data for welded marine and offshore
structures, it was thought advisable to fabricate rather large specimens, which would repre-
sent the practical geometric imperfections and residual stresses accurately. Most of these
large shells were stiffened shells, which are discussed in Chapter 13, but one unstiffened
example of the high quality large specimens was the Det norske Veritas aluminum alloy
welded 2-m diameter spherical shell segments (Figure 9.25) tested by Odland and Didriksen
in Oslo in 1977 [9.182]. It may be of interest that the specimens were manufactured by
SAAB-SCANIA Flygdivisioncn-an indication of the technology transfer. Each model was
assembled from three identical segments, which were shaped individually by stretching, and
finally joined together by meridional argon welds. The shell thickness was measured ultra-
sonically, and the variations found were less than 6 percent. Geometric imperfections were
measured and mapped (sec Chapter 10 and [9.183]) and found to be on the order of one
shell thickness.
In concluding this section on isotropic shells, it may again be mentioned that the fabri-
cation of stiffened shells is discussed in Chapter 13, Section 13.2.

9.4 Test Setups for Cylindrical Shells under Axial


Compression
As pointed out in the introduction to this chapter, after the basic experimental studies of the
thirties, structural engineers in the aeronautical industries concluded that since they could
not yet rely on theory for prediction of buckling, they had themselves to provide the empirical
data needed for design. This resulted in many experimental parametric studies in the forties,
fifties and sixties, carried out primarily in industry.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

9.4. 1 Typical Experiments of the Fifties and Sixties


A well-known example of this effort are the tests of Harris. Suer, Skene and Benjamin at
the Missile Division Laboratory of North American Aviation, Downey, California, in the late

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
656 Shell Buckling Experiments

fil(u re 9.25 Large Dct norskc Vcl'itns 2-m diameter alummurn alloy spherical shell segment in test rig
(from [9.1821. courtc>y of Dr. J. Odland)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

fifties [9.168). [9.184]- [9. 187]. "ini tiated to develop reliable design infonnation on the crit-
ical and compressive buckling stress for thin walled circu lar cylindrical shells with and
without internal pressure." The study commenced with axial loading. but it was then extended
to combined loading of axial compression. bending and tension. In the fifties. empha~is was
placed in missile structures on unstiffened (monocoque) circular cylindrical shells stabi lized
by internal pressure (the body of the well-known Convair. General Dynamics. Atlas balli stic
missile was such a pressurized thin -walled monocoque stainless steel shell, wi th R11 .,. 1500).
Hence. pressurized shells featured notably in most of the test programs of thi s decade and
also in the North American st udy. A schematic drawing of the o riginal test rig for axial
compression is shown in Figure 9.26a. The test cyli nders were fabricated by wrapping very
thin 0.005 in. (0.13 mm) thick aluminum alloy foil. or ~imilar and even thinner half-hard
stainless-steel foil. around a fonn and adhesive bonding a single 0.375 in. (9.5 rnm) longi-
tudinal seam. Some of the later cylindrical shells were joined by seam welding. which proved
to be a simpler and yet satisfactory technique. The te~t cylinders snugly fitted the lower and
upper heads of the test rig and were clamped to the head~ by rubber-lined steel straps. which
prevented slippage and leakage between the heads and the shell. To hold the cyli nder more
tightly to the heads. eight lugs were screwed agai n~t each strap. This type of clamping with
metal straps (see Figure 9.27) has been extensively employed in many industrial tests because
of its simplicity, but it presents boundary conditio ns that are not too well defined. as wi ll be
funher discussed in Chapter I I. Internal pressure was pneumatical ly supplied, hut for rea~ons
of safety the cylindrical shell~ were partially filled with water whenever the internal pre~sure
exceeded 2 psi. The axial compression loading wa; by a hydraulic strut (see Figure 9.26a)
and wa~ measured by a load cell. The buckling loads were detennined visually ~ince the
buckles usually snapped into position. In unpres~urited cylinders the typical diamond-shaped
buck.le pattern appeared. but with internal pressure the pattern changed by decreasi ng the
axial and circumferential wavelengt hs, though in different proportions. In pressurized shells,
incipient buckling was often indicated by the format ion of very shallow axisymmetri c ripples
(which could be seen in the reflective surface of the shell ). which subsequently became
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Test Setups for Cylindrical Shells under Axial Compression 657

PRESSURE
RELIEF VALVE

PRESSURE -· GAGE LOAD STRUT

WATER INLET
CLAMP
.,
"
l .1._
5!r~,
HEAD! I
HYDRAULIC kQAD STRUT
PRESSURE INLET
INTERNAL PRESSURE
INLET
f
PRESSURE SEAL

CIRCULAR SHELL

CELL

CLAMP
HEAD
DRAIN "

(a)

PRESSURE GAGE

CIRCULAR SHELL

COMPRESSION
LOAD CELL
HEAD

LOAD STRUT BENDING LOAD CELL


BENDING LOAD STRUT

(b)

Figure 9.26 North American Aviation lest rig for pressurized cylindrical shells-schematic (from
[9.168] and [9.185]): (a) for axial compression loading only, (b) for combined axial com-
pression, bending and torsion loads

narrow diamonds (see Figure 9.28). The authors then presented a statistical semi-empirical
design procedure (90 percent and 99 percent probability design curves for different Rlt
ratios), which was adopted by most of the aerospace industry and was further amplified in
1970 by Almroth et al. [9.188] to include also stiffened shells. They stressed, however, that
such a design method was still only an interim solution in the absence of "totally satisfactory
methods," such as have now been developed based on computations for shells with known
or predicted initial imperfections, discussed in Chapters 3 and 10.
For the extension of the experimental studies to other loading cases and load combinations,
the test rig was developed to apply axial compression, bending, torsion and external pressure
(see Figure 9.26b). Over 200 stainless steel and aluminum alloy cylindrical shells, with
(Rit) = 1000-2730, were tested. The initial imperfections of the test shells were not mea-
sured, as it was assumed that they were equivalent to those appearing in full-scale aircraft
structures and hence presented the appropriate empirical base for the design procedure. The
measured variations in thickness of less than l 0 percent of the mean were also considered
to be representative of the full-scale ones.
In the sixties, another extensive experimental program on unpressurized and pressurized
cylindrical and conical shells was carried out by Weingarten et al. at Space Technology
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
658 Shell Buckling Experiments

(b)

(a)

Figure 9.27 Typical buckle pauems for pressurized thin cylindrical shells under axial compression
(Rit)"' 1000- 2700 (from [9.168])

Laboratories (now TRW), Los Angeles [9.103) and [9.104], which yielded lower bounds for
cylindrical and conical shells that were extensively used in industry. Mylar test specimens
of (Rit) = 100-2000, were employed with some comparative experiments on steel shells,
which exhibited effects of microscopic plastic deformations that significantly reduced their
repeated buckl ing loads. In the Mylar shells the plastic deformations were much less and
did not affect the buckling loads. This might have detracted from the empirical value of the
data, but si nce the lower bound curves bounded also most previous test data, nearly all from
metal shells, which compensated for this deficiency, the bounds were val id.
The test specimen assembly of Weingarten et al. for internally pressurized cylindrical
shells is shown in Figure 9.29. The shells were of 8 in. diameter and 8 in. length and were
cast into grooves in the end caps with a potting material. The grooves were stepped to provide
a locating diameter during casting. The alignment of the two end caps during assembly was
provided by the center post and ball bushing assembly. The center post of the top end cap
also provided a mounting for a differential transformer, which measured the relative axial
displacement between the caps. To avoid wrinkling of the very thin (0.002 in. and 0.003 in.)
Mylar shells, the top end cap weight had to be counterbalanced and great care was required
in installation of the specimens.
Tntemal pressure was applied with compressed air through ports in the bottom end cap
and measured by manometers. The axial compression was applied with a motorized screw
mechanism in series with a load cell, a ball bearing and a floating cover plate, load and
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

dcfonnation being recorded si multaneously with an X- Y recorder. In the tests, a fixed pres-

(O ) (b)

Figure 9.28 Shallow axisymmeuic ripples and subsequen< buckle pauern for a pressurized thin cylin-
dJical shell under axial compression (from [9.168]): (a) ripples at 1500 lb load, (b) buckle
pauern (max. load = 1590 lb)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Test Setups for Cylindrical Shells under Axial Compression 659

STATHAM
LOAD-CELL __ _

DIFFERENTIAL
TRANSFORMER

Figure 9.29 Space Technology Laboratories (now TRW) test specimen assembly for internally pres-
surized cylindrical shells subjected to axial compression (from [9.104])

sure was applied and then the axial load was gradually increased till snap buckling occurred
or till an ultimate load was reached. The axial load was then decreased till the buckles
disappeared. Then the internal pressure was increased by a fixed amount and the shell was
again buckled by axial load, the repeated buckling causing no degradation of the Mylar
shells. At higher internal pressures, shallow axisymmetric ripples appeared prior to buckling
or collapse, similar to those observed in the North American tests (see Figure 9.28).
For casting the specimens into the end caps, two different common low-melting-point
alloys were used: Cerrobend, which expands while cooling in some series of tests, and
Cerrolow. which contracts a much smaller amount (about 1/25th) while cooling in others.
For no, or very low, internal pressure, consistent differences in buckling load were observed
depending on the casting material used. As can be seen in Figure 9.30, the buckling stresses
obtained with shells cast with Cerrolow were consistently higher than those for similar ones
cast with Cerrobend. Because of its expansion upon solidifying, the casting with Cerrobend
gave the shells more eccentricity at the ends, representing a shortwave local axisymmetric
imperfection, which caused the observed reduction in buckling stresses. Note the significant
effect of an apparently secondary installation factor.

9.4.2 Tohoku University Test Setup for Postbuckling Studies


In the sixties and seventies, as high-powered computers became available, new nonlinear
theoretical studies of the immediate and deep postbuckling behavior of shells thrived. They
were accompanied by careful experiments aiming at the verification of these theories. The

05
0 0 CERROI.OW
0.4 x CERROBEND

0.3 fl 0
0

' l
O'eR
0.2

0.1

0
0 500 1000 1500 2000
Rtt

Figure 9.30 E!Iect of end conditions, caused by solidification characteristics of casting materials, on
the buckling stress of unpressurized cylindrical shells under axial compression (from
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
[9.104])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
660 Shell Buckling Experiments

extensive theoretical and experimental studies with Mylar shells on the deep postbuckling
behavior by Thielemann, Esslinger, Geier and their co-workers at the DLR in Braunschweig,
already mentioned in Sections 9.2 and 9.3 [9.74], [9.105], [9.112]-[9.116], and [9.189], are
representative of these investigations. The studies by Yamaki and his co-workers at the
Institute of High Speed Mechanics, Tohoku University, Sendai, Japan [9.109]-[9.111],
[9.190], [9.191] and [2.44], are another good example. Their experimental studies are typical
of carefully performed buckling and postbuckling tests of cylindrical shells. The specimens
in these experiments are also made of Mylar-type Diafoil polyester film, for the reasons
pointed out in Section 9.3.
The test apparatus for axial compression is shown schematically in Figure 9.31 (from
[9.110]). In the figure, (1) is the table of a tension test machine, (2) is a base plate, (3) is a
test cylinder and (4) is a stiffening steel plate fixed to the cover plate. To the crosshead (5)
of the test machine, a reduction gear bo" (6) is fixed and, by turning the lever (7), the
cylinder (8) can be moved in the axial direction. A strain-gage load cell (9) with the capacity
3000 N ( = 300 kg) is connected to (8), which is directly calibrated with the test machine.
The loading head (1 0) has three equidistant legs, on both faces of which strain gages (11)
are bonded. The compressive load on each leg can be finely adjusted by turning the bolt
(12), observing the strain output from (11). The loading head is constrained to move in the
axial direction by a frictionless ball bushing (13), supported by a three-leg stand (14). With
this loading system, tests were performed, controlling the axial shortening of the shell. How-
ever, it is to be noted that the shell is also subjected to a dead load of 50.2 N ( = 5.12 kg),
due to the weights of the loading head, stiffening as well as cover plates.
The deflection-measuring apparatus (16), including from (15) to (20), is the same as that
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

used in earlier tests on cylindrical shells under hydrostatic pressure [9.109]. It is attached to
an annular disc (15), which can be rotated around the base plate (2). (17) is a rolling contact
point, utilizing a ballpoint of a commercial writing pen, and (18) are strain gages bonded to
both sides of a 0.2 mm-thick phosphor bronze plate attached to the end of an aluminum
lever. The displacement of (17) can be measured through the bending strain of (18), the
calibration being made with a micrometer (19), and the cantilever set ( 17) to (19) can also
be moved in the axial direction by using the rack mechanism (20). With this apparatus, both
axial and circumferential distributions of the deflection can be recorded precisely without
distorting the buckled surface of the shell. The edge shortening of the shell is measured
utilizing the strain gages (21) bonded to a thin cantilever fixed at the end of a dial gage
(22), which in turn is used for calibration and is attached to a stand also fixed to the annular

Figure 9.31 Schematic diagram of Yamaki's test setup for cylindrical shells under axial compression
(from [9.110], courtesy of Professor N. Yamaki)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Test Setups for Cylindrical Shells under Axial Compression 661

disc (15). Two sets of these are used in series to obtain the average value of edge shortenings
at the diametrically opposite points on the stiffening plate (4). A six-element dynamic strain
meter is used for strain measurements, while an X-Y recorder records the results.
In the tests the relation between the compressive load P and edge shortening 8 is obtained
by lowering or reversing the loading head (10). Typical results for specimen no. 200-3, with
a Batdorf geometry parameter Z = 200, are shown in Figure 9.32. As shown by solid lines,
P increased linearly with 8 until the primary critical load ( = 980 N) was reached, at which
a dynamic snap-through buckling occurred with a sudden load reduction. The stable post-
buckling configuration reached was of an asymmetric type with two tiers of staggered buck-
les, having the circumferential wave number n = 14. With further increases in 8, secondary
snap-through buckling occurred successively to the new equilibrium state, with n reduced
by one each time. On the other hand, when the end shortening was reduced, the shell lost
its postbuckling stability at another critical point along the equilibrium path, at a lower 8,
and snapped back to the immediate previous path with n increased by one. With further
reduction of the end shortening, the process was repeated and the shells finally returned to
the prebuckling state. The asymmetric postbuckling configuration was preserved in these
repeated snapping processes, though the wave number changed. However, in the deep post-
buckling region, when 8 was larger than 0.5 mm, the aforesaid asymmetric configuration
became locally unstable, with a tendency to degenerate into symmetric central buckles. In
this case a fully symmetric postbuckling configuration (with respect to the central section of
the shell) with uniformly distributed one-tier buckles could be easily realized with a slight
adjustment of the shell wall with fingertips. These additional symmetric postbuckling equi-
librium paths are shown in Figure 9.32 by dotted lines, representing another system, in which
as 8 decreased the number of circumferential waves increased successively by one. At higher
critical values of 8 the stability of the symmetric postbuckling configuration was, however,
lost with initiation of local torsional deformations. Similar postbuckling behavior was also
observed for long shells with Z greater than 200, but for relatively short shells, with Z :s
100, the postbuckling patterns were always symmetric, and the asymmetric two-tier ones
could not be obtained.
To see the overall distribution of buckled waves, circumferential distributions of w (pos-
itive inward) were recorded along various circular sections, with a typical result (for P =
478 N, 8 = 0.112 mm and n = 14) shown in Figure 9.33. In the figure, x and e stand for
axial and angular coordinates, respectively. One may note that the longitudinal seam, which
is located at the right-hand side of the figure, had no significant effect on the uniform
distribution of the waves. From such circumferential and axial distributions of the deflection

1000
p
INI Asymmetric

800 Symmetric

0 01 0.2 0.3 04 8 !mml


0.6

Figure 9.32 Variation of the axial load P and wave number n with axial shortening 8, for Yamaki's
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Diafoil polyester specimen with Z = 200 (from [9.110], courtesy of Professor N. Yamaki)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
662 Shell Buckling Experiments

x/L
1.0
I

r=~~~· ~0.9

~V\?\:7\/V'v\:f'~~ os
'pr\J'vovov'\f\JV\l\f\fV~f\F' o.?
w(mm)
'rj\:f'if'if'if\;t\:1\f'\J\f\.f\)'~ 06
-ol [ I.e---~---
"" ~0.5
I

~::
i"""=~~_,..--.__=~~~~==~........-07'""=~~~~.-J 0.1

0" 90" 180" 27o" e 3so'

Figure 9.33 Circumferential distributions of the deflection w, showing the postbuckling waves around
a typical cylindrical shell with Z = 200 (from [9. 110], courtesy of Professor N. Yamaki)

w, contour lines could be obtained for typical postbuckling configurations. Some of these
contour lines are shown in Figure 9.34.
The details of Yamaki's test setup and procedure, as well as the relevant measurements,
have been presented here to indicate the care and effort necessary to perform a meaningful
postbuckling experiment for a thin shell and to point out the precise information that can be
obtained. This information not only deepens our understanding of the postbuckling behavior
of the shell, but can also be used to evaluate and verify postbuckling analyses, as mentioned
earlier. Regarding the physical behavior, the experiments considered (of [9.110]) clearly show
for the case of axial compression the decrease in the number of circumferential waves as
one proceeds into the postbuckling region. This concurs with the results of other postbuckling
experiments on Mylar cylindrical shells under axial compression (for example, those sum-
marized in Chapter 3 of [9.74], or in [9.189]. Under hydrostatic pressure, however, similar
experiments [9.109] showed that the number of circumferential waves remained constant

Ia) P•478N, ti•OII2mm, n•l4


--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,

Figure 9.34 Contour lines (in mm) for a typical postbuckling configuration of a cylindrical shell with
Z = 200 under axial compression (from [9.110], courtesy of Professor N. Yamaki)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Test Setups for Cylindrical Shells under Axial Compression 663

until a torsional pattern occurred in which the cylinder finally failed at excessively high
pressures. Also, under torsion [9.111] the number of buckling waves remained unchanged
with further advance into the postbuckling region. As under compression, Yamaki's results
for these two loading cases (discussed in more detail below) concur with those observed in
postbuckling tests by other investigators (for example in [9.74], [9.105] or [9.115]). It may
be of interest to recall (from Chapter 8) that for rectangular plates under axial compression
the number of longitudinal waves increases as one advances into the postbuckling region.
Regarding the comparison between theory and experiment for axial compression, Yamaki
and his co-workers obtained fairly good correlation between their carefully measured ex-
perimental results and their calculations [9.191] performed later. Such a comparison is shown
for one case in Figure 9.35. Two typical presentations are shown in the figure: (1) load-
shortening relations and (2) load-deflection relations. Both sets of curves are for the asym-
metric postbuckling mode. As already mentioned, a symmetric postbuckling mode for the
same case also exists, which has been measured and calculated (see [9.191]). Esslinger and
her co-workers [9.112] also showed fair correlation between experiment and theory in the

1·0>,...-----------------,
Z•500
(Asymmetnc)

-Theory
---Experiment
0

0 0·5 1·0 1·5 2·0 2·5

(a) LOAD- SHORTENING

l.: z •500
{Asymmetric)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

"

(b) LOAD- DEFLECTION

Figure 9.35 Load-shortening and load-deflection relations for the asymmetric postbuckling mode of
an axially compressed shell with Z = 500, (Rit) = 405, and v = 0.3 (from [9.191],
courtesy of Professor N. Yamaki)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
664 Shell Buckling Experiments

deep postbuckling region. This correlation, which is better than that obtained at buckling
and in its immediate neighborhood, is an indication that the geometric imperfections, which
are the prime cause of scatter in the buckling loads of isotropical shells (and are discussed
in more detail in Chapter 10), lose their dominant influence in the far postbuckling region.

9.4.3 Stanford University High-Precision Test Rig


In the quest for better axial compression experiments, leading to buckling loads that are
closer to the classical one and to reduction of scatter, attention was also directed to better
test rigs. An interesting loading rig was developed by Sendelbeck and Hoff at Stanford
University, in which thin cylindrical electroformed nickel shells buckled in a range from 80
to 96 percent of the classical critical stress [9 .192]. The two features that make this loading
rig better are: (1) a simple procedure for accurate axial alignment of the shell and (2) tapered
circular end caps, which ensure its circularity.
The test rig (which uses a conventional I 0,000 lb capacity Tinius Olsen testing machine
for movable cross-heads) is shown in Figure 9.36. It consists of a 1.75 in. (44.5 mm) diameter
steel shaft upon which two end fixtures slide freely but maintain a close fit (see Figure
9.36a). The sliding fit condition was accomplished by first boring out each end cap to the
same diameter as the shaft and then heating the aluminum housings several hundred degrees
until they slid easily over the shaft. Number 600 mesh carbide power was mixed in a light
oil to form a paste, which was then applied to the shaft. Pushing the shaft back and forth
in the slowly cooling housing made simultaneous lapping of the two parts possible, which
resulted in a good sliding fit at room temperature when a silicone surface lubricant was
applied. Since the mating of the two parts was continuous over the entire length of the
housing end caps, longitudinal wobbling was minimized.
The end fixtures have a tapered circular shank over which the shell specimen is fitted.
Thus, the circular shape of the end sections is maintained and at the same time coincidence

T~IN C:VUNORICAL. SHELL


(TEST SPECIMEN)

g) FOil SEVERELY AESTAAINING END MOTION b) FOA ALLOWING UNRESTRAINED END MOTIONS

Figure 9.36 Cutaway views of Stanford test rig for cylindrical shells under axial compression (from
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
[9.192]): (a) for severely restrained end motions, (b) for allowing unrestrained end motions
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Test Setups for Cylindrical Shells under Axial Compression 665

of the axes of the central shaft and the specimen is ensured. A trough surrounding the shank
holds the material that is used for encasing or "potting in" the ends of the shell.
Sendelbeck and Hoff used as potting material a cold-setting epoxy resin (2 parts Epon
resin No. 828 to 1 part curing agent RF 61 of the Shell Oil Company). After completion of
a test, the assembly was heated to about 250°F (121 °C). At this temperature the epoxy
became pliable enough to be removed from the end caps, which could then be used again.
After the shell had been potted, the 1.75 in. guide shaft could be removed (Figure 9.36b)
or left in place (Figure 9.36a), depending upon which type of test the shell was to undergo.
Tests performed with the guide shaft in place had the effect of laterally restraining end
motions. Tests with the shaft removed allowed unrestrained end motions.
This assembly of shell and end fixtures was then screwed on an alignment platform that
was located on top of the upper cross-head of a conventional test machine. In Figure 9.36a
the assembly is shown with the guide shaft in place. The load applied by the upward move-
ment of the cross-head was transmitted here to the shell assembly by the guide shaft via a
shallow cap at the top and a load cell, a connecting short chain and short rod at the bottom.
When the guide shaft was removed (Figure 9.36b), the load was transmitted by a short
chain and a straight 318 in. (9.5 mm) diameter steel rod via a central spherical thrust bushing
plug at the top and a load cell at the bottom. This 318 in. rod passed through a 112 in. hole
in the alignment platform. Since the centerline of this hole was coincident with the axis of
the shell, the position of the rod in the hole indicated the degree of alignment between the
shell axis and the load transmitting axis. By manipulating three alignment screws on the
alignment platform until the rod located itself in the center of this oversized hole, it was
possible to align the rod with the axis of the shell, this operation taking about a minute. The
maximum initial error that could occur prior to load application, when the rod touched one
side of the hole, was a 0.4 percent side load.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

In both setups, the end shortening of the shell was measured by four dial gages located
evenly around it, which also indicated the extent of any end capping tipping.
The loading rig was evaluated by testing near-perfect shells and comparing the consistency
and level of buckling stresses with those obtained in earlier investigations of near perfect
shells. The 15 electroformed nickel shells tested had a nominal diameter of 2.90 in. (73.7
mm) and covered a range of (Rit) = 470-960. The mounting of the shells over the tapered
end fixtures in the test rig reduced the edge distortion and resulted in smaller initial imper-
fections or initial out-of-roundness. The alignment deviations were also very small in both
test setups.
Sendelbeck and Hoff measured the initial deviations from circularity (the initial geometric
imperfections), taken at 10 percent length from the base, of their nickel shells and compared
them in [9.192] with those measured by Okubo ct al. on their Mylar shells of (RI f) = 371
[9.193] and those measured by Arbocz and Babcock on their electroformed copper shells of
(RI t) = 800 [9.194]. After all the results were adjusted to the same radius for the comparison,
the reduction of initial deviations due to the tapered end fixtures in the nickel shells was
shown to be evident.
In Figure 9.37 the ratio of the experimental buckling load under axial compression to the
classical one, the knock-down factor, of the nickel shells tested in the Stanford test rig are
compared to other near-perfect shells: the Mylar shells of [9.193], two series of Caltech
electroformed copper shells, those of [9.194] mentioned above and those of (Rit) = 900 of
[9.86], and the spin-cast epoxy shells of (Rit) = 219 tested by Tennyson and Muggeridge
[9.195]. As is clearly seen in the figure, the electroformed nickel shells tested in the Stanford
test rig exhibit consistency and a higher level of buckling stresses for their range of (Rit)
ratios, pointing out the advantage of this loading rig. Figure 9.37 also shows three curves of
computed knock-down factors based on Koitcr's theory, taken from [9.196]. These curves
represent earlier standards of shell manufacture. The ordinary shells were assumed there to
have ratios of amplitude of initial imperfections to thickness of the order of 10-' that of
their (R It) and the expertly made ones imperfection amplitude ratios of I 0 4 their (R It). For
(R It) = 1000 this means imperfection amplitudes 10 times the thickness for ordinary shells
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
666 Shell Buckling Experiments

1.0.-------------------------------,
EXPERIMENTAL POINTS
0 SENDELBECK, HOFF l NICKEL).
OOKUBO, WILSON, WHITTIER !MYLAR) [9.193]
8 BABCOCK, SECHLER (COPPER) [9.86 j
VI TENNYSON MUGGERIDGE
(PHOTbELASTIC PLASTIC) (9.195)

rE
....
rl.t 0.4

0.2

0 800 1600 2400 3200


R/t

Figure 9.37 Plot of buckling load ratio (P"./ 1', 1) versus (R It) of circular cylindrical shells under axial
compression (from [9.192]; the curves are taken from [9.196])

and on the order of t for expertly made ones, with well-made ones in between. The com-
parison brings out the high quality of the near-perfect shell models and the improvement
brought about by the Stanford test rig.
It may be added that the use of the guide shaft did not perceptibly affect the buckling
load. In the tests, 4 shells were buckled with the guide shaft in position and II with it
removed. This was good evidence that the frictional forces between shaft and the end fixtures
remained small, as they would have otherwise increased the buckling loads, and that the
axial alignment remained accurate after removal of the shaft, since misalignment would have
reduced the buckling loads.

9.4.4 Typical Modern Test Systems for Cylindrical Shells


Now, having discussed some of the earlier test rigs for cylindrical shells (see for example
Figures 9.26 and 9.29), as well as some special-purpose ones (see Figures 9.31 and 9.36),
two larger modern test setups for cylindrical shells subjected to axial compression. repre-
senting the state of the art, are discussed m detail.
One, the RZIOO buckling test system of the DLR Institut ftir Strukturmechanik, Braun-
schweig, has already been mentioned in Section 9.2.1 in connection with high-speed pho-
tography studies of the buckling process. The test setup (see Figures 9.38 and 9.39) was
developed and improved over the last two decades. It is a large test rig, 5.1 m high, 2.3 m
wide (without the operators' balcony), with 2 m distance between cross-heads and can take
specimens up to 1.1 m in diameter. The rig has a I MN (~I 00 ton) capacity, though at
present only a 0.4 MN jack is provided. Its design was guided by the following requirements,
which generally apply to such test systems (see [9.197] and [9.198]):
1. The test rig should accommodate test specimens in a manner that does not give rise to
residual stresses and edge distortions.
2. The boundary conditions of the specimen should be well defined.
3. Exact axial alignment is essential, which means very close bearing fits and high stiffness
of the setup.
4. The tolerances of the moving parts should be limited to a few microns.
The test facility is shown schematically in Figure 9.39. The top cross-head rests on three
stiff screw columns. With its driveshafts, an electric motor on the top cross-head rotates three
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Test Setups for Cylindrical Shells under Axial Compression 667

Figure 9.38 RZ 100 shell buckling test setup at DLR Braunschweig (courtesy of Dr. B. Geier)

large nuts on these columns, wh ich causes the top cross-head to move vertically. The close
tolerances on this drive system limit the wobbli ng of the top cross-head to less than 20
J.Li m. T he purpose of the electric dri ve is the vertical location of the test specimen and
measuring system in the test facilit y. whereas during the test the motor is inoperative and
the hydraulic jack takes over entirely. In order to cancel the beari ng clearances, which are
necessary for the vertical movement of the top cross-head but would disturb the test. the
screws are then locked to the top cross-head. Th is locki ng is achieved by six small hydraulic

_ _IESL _
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

SPECIMEN

Figure 9.39 Schematic diagram of the large DLR RZ 100 buckling test setup ( from [9.197], courtesy
of Dr. B. Geier)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
668 Shell Buckling Experiments

spanners for each screw column, which are activated automatically when the oil flow to the
main jack initiates. The top cross-head is then positively locked in its position during the
test. After unloading, these spanners also automatically release the top cross-head to permit
its vertical motion.
The movable cross-head has a very accurate parallel guidance consisting of 16 individually
adjustable no-play supporting rollers that act on a stiff box structure, ensuring axial alignment
of the cross-head. Instead of by the hydraulic jack, the movable cross-head can also be raised
by an electric motor that drives two additional built-in screw jacks. Their primary purpose
is a slight preloading during the mounting of the specimen and casting it with resin in the
grooves of the end plates. If the hydraulic system were used for this preloading, it would
have to operate during the entire resin ctlfing period, whereas the auxiliary electric drive
retains the position and preloading after it is switched off.
The load and bending moment (which is nominally zero) are measured by three load cells.
For cylindrical shells under axial compression, these three concentrated loads have to be
converted to a uniformly distributed loading. This is achieved by a very stiff load distributor,
shown schematically in Figure 9.40.
The method of equalization of the load is also shown in this figure. The end plates of the
specimens have grooves filled with resin (which contains a filler for greater stiffness). The
shell is first set into the resin of one end plate and the resin is cured. Then the second end
plate is attached in a similar manner, which gives rise only to negligible residual stresses in
the shell. The experimental boundary conditions are close to fully clamped ones, on account
of high stiffness of the filler reinforced re~.in, as will be further discussed in Chapter ll. The
specimen, together with its end plates, is then mounted between the load distributor and the
top cross-head, with an equalizing layer, consisting also of resin with filler, introduced at
each end (see Figure 9.40). These resin layers lead to very good load uniformity.
The control system of the test setup (which is typical of such test machines) is shown
schematically in Figure 9.41 for a cylindrical shell under axial compression. The output of
the controller is an analog voltage that controls the movement of the hydraulic jack. The
servo valve at the jack converts this analog voltage into piston movement by opening the
oil flow above or below the piston. Thus the specimen is loaded or unloaded. The displace-
ment (the axial shortening) is measured by displacement pickups and the load by load cells.
Their output signals are also analog voltages, which present one of the input signals to the
controller, the control point. The other input signal, the set point, is an analog voltage set
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

by the user.
If the displacement is connected as the control point, the set point is the desired axial
shortening. The test setup is then displacement controlled, which means that the displacement
at any time depends only on the drive of the test machine and not on the behavior of the

SUPPORT

- CYLINDRICAL
stiE:l._C___ _

DISTRIBUTED
LOAD
EQUALIZING
LAYER

LOAD
DISTRIBUTION

SINGLE
FORCES

Figure 9.40 Loading system of RZ 100 test setup-schematic (from [9.197], courtesy of Dr. B. Geier)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Test Setups for Cylindrical Shells under Axial Compression 669

DISPLACEMENT
PICKUP

Figure 9.41 Control systems of RZ 100 test setup-schematic (from [9.197], courtesy of Dr. B. Geier)

specimen. If the load is connected as the control point, the set point is the desired axial
loading. The test rig is then load controlled. The comparison of the load path with displace-
ment control and that with load control, shown in Figure 9.42, emphasizes the significantly
different postbuckling load paths and behavior. Up to buckling, the mode of loading does
not affect the behavior of the specimen and the buckling load is identical for displacement
control and load control. It may be recalled that in Subsection 9.2.2 it was pointed out that
the buckling load is entirely independent of the stiffness of the loading device, which sup-
ports the present statement that the buckling load is independent of the mode of loading.
At buckling, the specimen jumps to a new state of equilibrium. For displacement control,
the axial shortening remains constant and the new state of equilibrium is very rapidly formed
at a lower level of load (see Figure 9.42), with shallow buckles and elastic strains. It may
be recalled from Subsection 9.2.1 that the time of initiation of buckling is less than 1 ms.

THEORETICAL
BLICK LING
LOAD OF
LOAD

l
I_[ ____..
PERFECT
SHELL
LOAD CONTROL
EXPERIMENTAL} NO NEW STATE
BUCKLING DF EQUILIBRIUM
LOAD BEFORE FAILURE
II
/1, DISPLACEMENT
/ I CONTROL
/ I DROP IN LOAD AT CONSTANT
{ + AXIAL SHORTEN lNG

l,,_L/~
NEW STATE OF EQUILIBRIUM

AXIAL SHORTENING

Figure 9.42 Comparison of load paths for displacement control and load control in a typical cylindrical
shell under axial compression
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
670 Shell Buckling Experiments

After unloading, the shell is therefore undamaged and can be retested. For example, at DLR
Braunschweig carbon fiber thin cylindrical shells with (R/t) = 250 have buckled 20,000
times without any change in behavior. For load control, on the other hand, the test system
jumps along a constant load line and tries to find a new state of equilibrium there (see Figure
9.42), but the shells fail much before that Hence, for repeated tests, displacement control is
evidently preferable.
The hydraulic control loop of the test system consists of the jack with the regulating valve,
the control point (the measured value after amplification) and the control amplifier. The
amplified measured displacement, or load, and the corresponding set point are fed into the
controller. The difference between the two analog voltages is amplified proportionally and
in addition their integral and differential are produced. The sum of these three signals results
in the command to the regulating valve. Since the controller includes three functions, Pro-
portional, Integral and Differential, it is called a PID controller.
The maximal actuating velocity practically depends only on flow cross-sections of the
pipes and the regulating valve. In the RZ 100 system a valve with a small cross-section was
chosen in order to reduce the sensitivity to disturbances by utilizing the integrating effect of
the resulting slow control. In spite of the small flow cross-sections, the control response is
rather rapid and very stable, provided the control parameters are adjusted appropriately.
Hence the system can easily hold the movable cross-head in a set position with an accuracy
of less than l mm, while the control loop displays a nearly ideal behavior.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
The high efficiency of the system results, however, in an increased sensitivity to operator
errors and malfunctions. An electric power cut or a malfunction of instruments in the control
circuit may cause a system instability, which can only be eliminated by a complete drop of
oil pressure, ensured by automatic switchgear. Experience has shown that these operational
troubles can be kept in bounds, provided the operating personnel keeps highly alert. With
the right care, however, the RZlOO system presents a well-adjusted test setup, practically
free of errors, for both static and dynamic shell buckling experiments.
In a typical cylindrical shell test the strain distribution is measured by about 100 strain
gages. They have two primary tasks: (l) verification of uniformity of loading, and therefore
many strain gages are located near both and plates (usually in pairs, on the inside and outside
of the shell); (2) determination of the stiffness of the shell, for which strain rosettes are
employed on the inside and outside, at six locations over the circumference.
For measurement of the contour of the shell surface, and thus its shape imperfections, a
scanning system is provided, which rides on a toothed ring mounted on the movable cross-
head that can been seen in Figure 9.38. The scanner, consisting of a carriage with a non-
contact probe, moves up and down the shell at regular circumferential positions, charting a
quarter of the reference cylinder, while the probe records the deviation of the shell surface
from the reference surface. The system has four such scanners and the scanning is actuated
by electric motors. It is typical of modern imperfection measurement systems discussed in
detail in Chapter 10.
It may be mentioned that though the main task of this test system is buckling of shells
under axial compression, it is also used for experiments on shells under external pressure
and, with the addition of a clamping frame and a shear loading device, for tests on panels
under combined axial compression and shear (see [9.197]).
The other one is the series of test systems that have been developed since the eighties by
Jullien and his students at the shell buckling research center of the French Institut National
des Sciences Appliques (INSA) in Lyon., where great strides in electroforming were also
made, as described earlier in Section 9.3. Two test systems of the series, referred to as A
and B, are presented. The earlier one, test system A, is smaller in capacity than the DLR
one but has a remarkably wide range of measurement capabilities. It was designed specifi-
cally for buckling studies on cylindrical shells, after a comprehensive review of the test rigs
and measurement systems used in other laboratories (see [9.99]-[9.101]).
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Test Setups for Cylindrical Shells under Axial Compression 671

A schematic view of test system A is shown in Figure 9.43. The test setup cons ists of an
automatically regulated loading system, an imperfection scanning dev ice, a series of force
and displacement sensors connected to an automatic data acquisitio n system and an optical
projected fringe system, located above the test section.
The loading jack ( I) and its hydraulic pump are located below the laboratory test floor.
The compressive axial load is applied to the test shell by pulling the central shaft (2). which
is attached lO the upper j aw (3). towards the lower jaw (4). Tn mai ntai n axial ity, the central
shaft is guided by two ball bearings, o ne (5) above the test flow and one (6) just above the
fixed support plate (7), which is attached to the top of the heavy barrel shape spacer (8)
(clearly seen in Figure 9.43), wh ich rests on the test noor. Axisymmetry of loading is ob-
tained by adjustment of three ball joi nt screws. according to the o utput from three peripheral
load cells (9), also located o n the spacer, which trims the top plate (10). Axisymmetry is
first adjusted in the same manner under an in itial prelnad.
There is a central inductive displacement sensor (I I). connected by an lnvar rod (of
negligible coefficient of ex pansion) to the top jaw, inside the central shaft for measurement
of the compressive shortening. However, to eli minate any '' parasitic " displacements, three
peripheral inductive displacement sensors ( 12) with lnvar rods, also inside the barrel-shaped
spacer. measure the di stance between the lower jaw and the lower support plate ( 13), and
their measured di splacement is deducted from that measured by the central displacement
sensor ( I I).
The compressive axial load is applied graduall y to the shell ( 14) till it is subjected to a
set compressive axial displacement. This loadi ng is achieved with the aid of a hydraulic
servo-vah•e (1 5). operated automatically by the axial displacement regulator (16) with a
program that integrates the measurements of the central load cell situated o n top of the upper
jaw and those of the central displacement sensor ( I I). The axial displacement is thus imposed
with an accuracy of ± I !-'111 and wi th a cnntrolled velocity of 25- 90 !Lm l min. T he axial
displacement regu lator ( 16) also monitors the axisymmetry of the load du ring the test and

Figure 9.43 INSA Lyon buckling shell tC~I system A- schematic (from [9.100]. counesy of Professor
J.F. Jullien)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
672 Shell Buckling Experiments

provides an on-line corrected load-displacement curve. The buckling load can be determined
from this curve.
The test system can also apply external pres~ure to the shells tested. by meam. of a vacuum
pump (not shown in Figure 9.43). If only lateral external pressure is desired. the regulator
( 16) cancels the resulting axial compression. S imilarly. internal pressure can be applied with
compressed air through pons in the bonom jaw.
An overall view of the more recent test system B is shown in Figure 9.44. This system
is of simi lar design to the earlier one hut can accommodate larger cylindrical shells (up to
a diameter of 30 em) under combi nations of axin l compression, horizontal shear and bendi ng
loads, and external or internal pressure (sec 19.199)). There is no optical (moire) measurement
system in test system B, but add itional vertical and horizomal loading jacks that have been
added above the specimen for axial comprc:.sion. shear and bending loads can be seen in
Figure 9.4-1.
Test system B was developed for buckling investigations of stiffened and unstiffened
cylindrical shells under combined loading. re lated to the cryogenic main stage tanks of the
Ariane 5 European Launcher (designed by CYROSPACE. Les Mureaux. near Paris).
An automatic ~canning system is provided on each test system for measurement of the
shell geometry. This system (see Figure 9.45) consbts of a carriage. holding the radial
displacement sensor. that moves venically on two guide rods, which rotate around the shell.
The vertica l translation and the rotation are propelled by electric motors and are rnea~ured
by an inducti ve sensor and potentiometer, re~pccti vcly. The position and radial di splacement
arc automnticull y recorded, resulling in an automatic scan (of equally spaced circumfcreru ial
or vert ical scans averagi ng about 60.000 poi nts) of the shell geometry. The ini tial geometri c

Fi~ure 9.44 INSA Lyon test system 8 ---{),erull view of tc\t for the cryogenic tank of Arianc 5. " In
the frame of ARlANE 5 European launcher development, S<'>me buckling te;t< hove been
ordered by CRYOSPACE to INSA Laboratoirc B~t(mS ct Structures on stiffened unci
unstiffened thin cylindrical shells (II t35 nun, t - 0.180 mm). under combined loading.
CRYOSPACE. joim-vemure of AEROSPATtAl.F. nnd L' AIR LIQUIDE. plnccd in Lcs
Murcaux near Paris. is in charge of the dc>ign and mnnufncturing of the wnks of ARIANE
5 Cryogenic Main State (5.4 rn diameter· LOX/LH2 tanks)."
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Test Setups for Cylindrical Shells under Axial Compression 673

Fit,'llre 9.45 INSA Lyon test system B- buckled shell in position with imperfection mca.1urcment
device (counesy of Professor J.F. Jullien)

imperfections are thus obtained and decomposed into Fourier series (in the usual manner, as
discussed in detai l in Chapter I 0). The system can be considered typical of modem imper-
fection scanning systems. The development of the shell geometry with load. or in other
words the growth of the imperfections. as well as the postbuckling shape, can also be sim-
ilarly recorded and analyzed.
The optical instruments of test system A are placed on the test system by means of a
special structure, (17) in Figure 9.43, which is rigidly attached to the top plate (10). Hence
the adjustment of the axisymmetry of loading does not affect the reference plane of the
optical system. The system consists of eight fringe projectors ( 18), which each project a grid
of vertical li nes onto the shell surface, covering the whole circumference. The projected
image of these grids is renected by four mirrors and focused in the image plane of the
photographic chamber (19). A negative image of the fringes on the initial unloaded shell is
placed on the image plane of the camera. in the exact place it occupied at the moment of
exposure, as a reference negative image (20). If the shell is observed through this reference
image. in its initial unloaded state, the image of the observed fringes coincides exactly with
the reference image. However. as the shell deforms under load , the observed fringes move
and the interference of their image with the reference one yields the characteristic moire
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

pattcms that represent the normal displacements of the shell surface. These patterns can be
recorded, through the semi-transparent mirror (21), with the camera (22) situated above, and
are recorded continuously with a video camera (23), viewing the reflection from the semi-
transparent mirror (2 1). The moire pattems are obtai ned for the entire shell (see for example
[9.99] or [9. I 00]). and facilitate monitoring and qualitat ive study of the prebuckling, buckling
and postbuckling behavior of the shell tested.
In some of the tests. an internal mandrel that restricts the buckle depth was also employed
here (see [9. 100)), resulting in a regular moire pattern similar to the buckle pattern observed
directly on the shells with intemal mandrels by Horton (see Figure 9.13).
Extensive strain gage measurements are also taken during some of the tests, and data are
acquired and reduced in a data logger that acts as a data acquisition center.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
674 Shell Buckling Experiments

In summary, the following are the principal parameters simultaneously measured and re-
corded during the entire buckling experiment [9.200]:
1. The axial load P effectively applied to the shell, measured by the load cell located on
top of the upper jaw
2. The three peripheral reactions, measured by three load cells, (9) in Figure 9.43
3. The total axial shortening t:..l, measured by an inductive central displacement sensor (l)
4. The local axial shortening t.l' at four generators, measured by additional inductive dis-
placement sensors inside the shell
5. The radial displacement w of one point of the shell surface, measured by a noncontacting
capacitance probe
In some experiments the following additional parameters are measured for better under-
standing of the behavior of the shell:
1. The exact axial deformations obtained from the mean of four to eight electric strain
gages bonded on the inside of the shell around one horizontal circle.
2. The relative rotation between the two jaws, measured by two inclinometers attached to
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

the upper jaw.


If the experiment is carried out with the optical system, as in test system A, the moire
patterns are also recorded simultaneously.
Before closing this section, it should be mentioned that some additional large test rigs that
serve primarily for stiffened shells are di,.cussed later in Chapter 13.

9.5 Recording of Buckling and Postbuckling Behavior


As was pointed out in Section 9.2, buckling is a very short dynamic occurrence (for axial
compression, 5-30 ms from initiation to the stable postbuckling state) and therefore the
measurement of its initiation and of the transition to the postbuckling state is not easy. One
should remember that the behavior of a cylindrical shell at buckling ditiers for different
loadings. For axial compression, as described in Subsection 9.2.1 (see Figure 9.3), there is
an abrupt and large drop in load at buckling. On the other hand, for external pressure,
discussed in Subsection 9.7 .1 (see Figure 9.69), buckling is less violent and the shell con-
tinues to carry a load fairly close to that at buckling, exhibiting a nearly stable postbuckling
behavior. Buckling under torsion, also discussed in Subsection 9.7.3 (see Figure 9.85), rep-
resents an intermediate case with a significant but small drop in load. The time interval of
the buckling process differs too, being the shortest for axial compression.

9.5. 1 Determination of Onset of fJuckling


The determination of the instant of buckling was the subject of many discussions among
experimenters in the sixties and seventies. Though detection of overall buckling is usually
not much of a problem when the structure has a large drop-off of load, since then buckling
is accompanied by a loud report and large radial displacements, the exact initiation of buck-
ling is more difficult to define. The determination of the initiation of buckling is important,
however, not only for the exact buckling load but also for arresting the loading in order to
prevent damage to the specimen and thus permit its reuse for repeated buckling tests. As
already pointed out by Babcock [9.58], if the structure has increased postbuckled load-
carrying capability or the postbuckled load is constant, initiation or the exact point of buck-
ling is more difficult to detect because the shell deflection (say, radial deflection under
hydrostatic pressure) is a continuous function of load and only the deformation rate increases
near the buckling load of the perfect shell. For imperfect shells, much of this behavior is

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Recording of Buckling and Postbuckling Behavior 675

masked by the imperfections, as in the case of flat plates discussed in Chapter 8, making
the identification of buckling more problematic.
Many investigators considered reliance on the "puck" sound at buckling and on the re-
corded load drop inadequate. Different devices were therefore tried for all types of loading,
and usually the onset of buckling was determined simultaneously by more than one method.
For example, in 1962 Technion experiments on conical shells under external pressure
[9.201], a special electrical contact visual indication system was installed. It consisted of 24
contact points (fixed to a perspex disc that was lowered into the cone), which were adjusted
to clear the cone by about one thickness and were connected in series (see Figure 9.97).
Onset of waves caused contact at some point, which was shown by an indicating light. The
visual indicator would give an accurate signal of onset of buckling if the shells were perfect,
or nearly so. The specimens, however, had initial imperfections, which grew steadily as the
pressure rose until they actuated the indicator. But by adjustment of the contact points during
the prebuckling part of the test, the system became suitable for indication of the onset of
buckling and was used for this purpose in the tests. After the onset of buckling the indicator
was removed.
A sensitive Statham pressure gage (which records via an unbonded strain gage) provided
a second means for determining the onset of buckling. As shown in Figure 9.97, a mechanical
pressure gage was fitted in parallel to allow continuous recalibration of the more sensitive
Statham gage, with which the initial buckling pressure was measured.
The audial indication (the "puck") was effective in the case of air loading (for the thinner
shells when the pressure was applied via air) but was nearly completely obliterated for oil
loading by the damping of the oil.
In another example, the investigation of local buckling of ring-stiffened cylindrical shells
subjected to axial compression by Tenerelli and Horton in 1969 [9.202], the instant of buck-
ling was again determined by three concurrent methods. A concentric mandrel restricting
the buckle depth to 3 I 4 of the shell thickness was fitted inside the shells. The first method
was a buzzer that was actuated when contact was made between the shell and the mandrel.
The second was a ring-type load cell above the top loading plate, which detected buckling
by the drop-off in load once it was more than 4 lb (18 N), i.e., 0.3-3 percent of the buckling
load. The third was a non-contact photonic sensor measuring the abrupt lateral displacements
at buckling by means of light rays. In a photonic sensor, light from a constant intensity
source is brought to the surface of a body to be studied (here the shell) via a fiber optic. It
is reflected from the surface, collected in a fiber system and retransmitted to the instrument,
where its intensity is compared with the initial level. Variations in the gap, caused by radial
deflection of the shell, are measured by changes in the amount of reflected light collected
by the probe. The resolution of a photonic sensor is 0.025 J.Lm but the center of a buckle
had to be located by a previous test, repeated buckling being feasible with an internal man-
drel.

9.5.2 Buckling Behavior of Oval Cylindrical Shells


--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

An interesting case of a type of shell which exhibits an increased postbuckled load carrying
capability even under axial compression is an oval shell. Feinstein et al. tested and studied
Mylar oval cylindrical shells under axial compression [9.203]. The geometry of the shells
was defined by the expression

(r/ ro) = l + t; cos 41TS (9.5)

where r is the local radius of curvature; r0 is the radius of a circle of a circumference equal
to that of the oval; 0 :S g :S 1 is a measure of ovality (g = 0 represents a circle, while for
g = I the minimum curvature is 0 at the narrow part of the shell cross-section); and S =

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
676 Shell Buckling Experiments

(s/ L 0 ), L0 being the circumferential length of the median line of the oval cross-section and
s the circumferential coordinate (see Figure 9.46). Three shell configurations:

g=O circular b/a = I


g = 0.5 moderate oval b/a = 1.4
g = 1.0 severe oval b/a = 2.1

with five thicknesses each, yielding (r0 / t) = 268-938 were tested.


The Mylar shells were formed by wrapping around mandrels, made from mild-steel tubing
(pressed or forged into shape for the oval shells). The shells were heat treated to relieve the
bending stresses, then anchored with CeJTobend in heavy end plates and finally loaded in
axial compression in a displacement-controlled Instron testing machine.
The difference in buckling behavior between the circular and oval cylindrical shells can
be seen in the experimental load versus end-shortening curves (Figure 9.46). For the circular
cylindrical shell, buckling occurs simultaneously over the entire circumference and the cat-
astrophic drop in load noted in Figure 9.3 is again observed. Indeed, the initial buckling
load and collapse load are identical. For the oval shells, however, the initial buckling occurs
much earlier in regions of minimum curvature of the circumference and is accompanied by
relatively little or no drop in load-carrying capacity. The collapse load, with its large drop
in load, is reached only when the buckles have propagated through the regions of maximum
curvature of the circumference, and it is significantly higher than the initial buckling load.
Initial buckling was defined here as the appearance of a visible buckle on the shell. It may
be noted in Figure 9.46 that for the severe oval (g = 1.0), there is no drop in load at initial
buckling, only a change of shape, which is therefore taken as evidence of the onset of
buckling.
It may be pointed out that oval shells, though usually exhibiting an initial unstable post-
buckling behavior similar to that for circular cylindrical shells (discussed in Chapter 3), as
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

in the moderate oval shell (g = 0.5) in Figure 9.46, they very quickly change their postbuck-
ling behavior to stable and their collapse load exceeds their initial buckling load (see also
[9.204] and [9.205]). They therefore reveal that there are types of shells for which the
imperfection sensitive initial postbuckling behavior presents an unrealistic and too severe
picture that is not borne out by experiment and analyses of the far postbuckling region.
The suddenness and rapidity of the buckling process present a triggering problem for the
recording equipment. An early elementary technique of triggered photography was attempted

( /
A'I
0.00 - - - - CIRCULAR / I
0.50 - - MODERATE CM!.l // I
1.00 - · - · SEVERE OVAL / I
/ I
/ I
D. INITIAL BUCKLING LOAD // I
o 0 COLLAPSE LOAD /
<t /
0
_J

.......

ENO SHORTENING

Figure 9.46 Experimental load versus end shortening curves for oval shells under axial compression
(from [9.203])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Recording of Buckling and Postbuckling Behavior 677

at Technion in 1967 during tests on ring-stiffened cylindrical shells under axial compression
[9.206]. The tests were carried out at night, and at higher loads the shells were not illumi-
nated. In order to eliminate the delay of shutter opening, the shutters of two cameras aimed
at the shell were kept open during each loading step (about 60 seconds). Flashlights were
then triggered by contacts set at predetermined distances to produce the photographs. The
results showed some initial buckling patterns but the records were not definite enough. More
sophisticated triggering, based for example on a number of fiberoptic sensors spread around
the shell, could perhaps yield better results, but none have yet been reported.

9.5.3 High-Speed Photography


A more promising approach is high-speed photography, discussed earlier in Subsection 9.2.1,
which was employed by some investigators in the sixties to study the unstable initial buckling
behavior of circular cylindrical shells under axial compression (see [9.67]-[9.73]). But there,
too, it will be remembered, timing of the camera, or triggering, was a problem since due to
its rapidity, the buckling process would be completed before the camera had accelerated to
its speed. Repeated tests therefore had to be employed to yield the expected time and location
of onset of buckling with which the camera could be synchronized. However, once this
synchronization was accomplished, very revealing records of the process resulted (see Fig-
ures 9.8-9.11).
Other recent forms of high-speed photography, such as rotating mirror-drum cameras,

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
continuous access rotating mirror cameras, and electronic cine cameras (see for example
[9.207]), could perhaps partly overcome these problems, though no successful application to
the buckling process has yet been published.
It must be noted here that triggering presents a difficult problem for shells whose initial
buckling pattern is not unique but depends on the imperfections, as in the case of isotropic
shells under axial compression. Since the location of the initiation of buckling is then not
known, many sensors are needed to realize the triggering by it. Whereas for shells with
unique buckling patterns, such as isotropic cylindrical shells under external pressure or
closely stringer-stiffened cylindrical shells under axial compression, the location of sensors,
and therefore triggering, is considerably easier.
Most of the high-speed photography records of the onset of buckling in shells were taken
at film speeds of about 1500-8000 frames per second. Even buckling of circular cylindrical
shells under transient axial impulsive load was recorded (with a 16 mm Hycam camera
recording the change in photoclastic isoclinic patterns that characterize the buckling modes
of the shells, as discussed in Subsection 9.2.1) at speeds of 1500-6500 frames per second
(see [9.208] and [9.209]). Modern high-speed cameras (such as Hycam and Fastax) can
usually operate at higher speeds, up to about 10,000 frames per second or more, but then
the total time span recorded is very small (about I second).
In dynamic buckling under impact, the triggering problem can be solved by activating the
impact when the recording camera is already at the required speed. Lindberg and Herbert
[9.210] accelerated their impacting mass explosively in a manner that ensured control of
impact time to within about 2 JLSec. This allowed them to use, for their cylindrical shells
buckling under axial impact, a Beckman-Whitely framing camera running at a speed of
240,000 frames/sec (with exposure time of about 1.4 JLSec), fast enough to observe the
details of the initial wave formation. These appear to show initiation of buckling correspond-
ing to small deflection theory but rapidly changing (after about 20 JLSec) to a large deflection
diamond pattern with non-linear effects predominating.
Recording of the postbuckling behavior is less problematic, as the process is considerably
slower than that of onset of buckling and essentially stable, except for the drops in load at
points of secondary buckling. Hence the same contact or noncontact sensors used for re-
cording of initial geometrical imperfections can be employed for measurement of the de-
formed shapes of the postbuckling domain. For example, in the seventies Yamaki and his
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
678 Shell Buckling Experiments

co-workers successfully used a very simple deflection-measuring contact lever, that activated
strain gages (see Figure 9.31) to obtain rather complete postbuckling contours of their cy-
lindrical shells. And in the early eighties Yamada and Yamada [9.211] employed a linear
displacement transducer, movable in the meridional and circumferential directions, to scan
their shallow spherical caps for the initial shape as well as the deformed postbuckling shapes
(see Figure 9.47). The analog measurements of the normal deflection w were recorded on
an X-Y recorder and then obtained numerically or graphically from the X-Y plot by means
of a graphic reading system and a digital computer. The measurements of w were to an
accuracy of 0.01 mm (0.3-1 percent of the shell thickness t, and more precise than t itself,
which was within ± 1.5 percent). The specimens in these experiments were formed by ther-
movacuum molding of rigid polyvinychloride sheets (similar to the processes discussed in
Section 9.3), and the resulting postbuckling curves (see Figure 9.48) show very good agree-
ment between experiment and theory.

9.5.4 Strain Gages for Detection of Incipient Buckling


Strain gages can also be employed for mea.surement of onset of buckling. One circumferential
row of five to six pairs of strain gages or more is usually bonded to the cylindrical shells
tested to check the symmetry and uniformity of loading. If the buckling pattern is more or
less uique, and the location of the buckles can therefore be reasonably assessed, this one
row will in general also suffice for detection of the onset of buckling. If, however, initial
buckling is more of a local nature and the location of the buckles strongly depends on the
geometric imperfections, as in the classic case of an isotropic shell under axial compression,
a very large number of strain gages is necessary to make detection of the onset of buckling
likely. But even a large number of strain gages does not ensure detection in such cases of
extremely unstable initial postbuckling behavior. Strain gages can be relied on primarily
when the initial postbuckling behavior is milder.
For example, in experiments on closely ring-stiffened cylindrical shells under axial corn-
pression carried out at Technion in the late sixties (see [9.206] or [9.212]), most of the
specimens were "covered" with 38-48 strain gages each (6-12 gages around the circum-
ference at each vertical location) in order to assist in detection of incipient buckling. The
strain gages proved to be excellent indicators of onset of buckling. In spite of the suddenness
of the actual buckling, most of the gages showed signs of near-buckling. Hence, buckling
could sometimes be predicted from the gages during the test to within 5 percent of the load.
An interesting result of the use of many strain gages was the spread over the shell of the
indications of incipient buckling. Instead of isolated local indications of near-buckling, the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

I. SPECIMEN 7. REGULATING VALVE


2. CIRCULAR GROOVE 8. PRESSURE TRANSDUCER
3. CHAMBER 9. DISPLACEMENT TRANSDUCER
4. VACUUM PUMP 10. SLIDING HOLDER
5. TANK II. RESISTANCE WIRE
6. EXHAUST VALVE 12. REVOLVING RING

Figure 9.47 Yamada and Yamada's experimental apparatus for clamped spherical caps under external
Copyright Wiley pressure-schematic (from [9.211])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Southwell's Method for Shells 679

Q-6., CURVES : -THEORY Q-V<: CURVES : ~ THEORY


- --- EXPERIMENT o EXPERfMENT
( •: EXPERIMENTAL BUCKLING PRESSURE)
..

54
·'

8 Vc I() I.Z~

-·~r
1
56 2
~4 \A' _ __ /
B

: : "1d'--,. ~. '

0 2 8 Vc 10 12

,,
a
.. 57
..
a ,.'
57

..
•2

Figure 9.48 Postbuckling behavior of clamped spherical caps under external pressure: pressure Q
versus the deflection at the apex l1 0 and versus volume change V, (from [9.211])

gages became lively at many locations simultaneously. As a matter of fact, the direction of
deviation of the strain gages in each horizontal row (which extended over complete circum-
ferences) was the same, with a few exceptions, indicating axisymmetric deformation. With
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

some stretching of the imagination, one could see in these indications of incipient buckling
the initial axisymmetric pattern (predicted by small deflection theory) that was missing, or
at least confirmation that initial buckling has a complete periodic pattern that differs from
the usual two-tier diamond postbuckling pattern. This evidence supports the conclusion of
Subsection 9.2.1 that the initial unstable buckling pattern differs significantly from the stable
postbuckling one.
In the fifties and sixties, extensive strain gaging, which appears to have advantages, re-
quired considerable recording and data reduction efforts and was therefore not popular. Now-
adays, however, data acquisition and reduction has become more automatic and much more
convenient with modern data loggers and PCs, making extensive strain gaging an option
worth considering in shell buckling experiments. Extensive strain gaging has the additional
advantage that it makes the use of Southwell plots feasible.

9.6 Southwell's Method for Shells


In Chapter 4 the Southwell method was derived, its usefulness and extension to many types
of columns, beam columns and frames discussed in detail and the general limitations of its
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
680 Shell Buckling Experiments

validity elucidated. The extension of the method to plates and its limitations there were
discussed in Chapter 8. Also for shells, the Southwell method has considerable potential for
smoothing data in parametric studies and as a nondestructive technique for determination of
buckling loads and hence warrants a detailed study. The discussion will again focus on
cylindrical shells, but applications to other shells, as well as stiffened shells, will also be
included.
Donnell in 1938 [4.26] was the first to consider application of the Southwell method to
shells, and he applied it directly to one of the most difficult cases, that of a cylindrical shell
under axial compression. Employing his 1934 approximate finite deflection solution for an
imperfect cylinder [9.54], in which "square" waves and an initial imperfection of the same
shape as the radial deflection were assumed, he obtained from energy considerations the
following expression for P:

1+ ( ~)(W + 2W,)(W + W,) }


(9.6)
+ ( ~~}(w + 2W,f(2W + W;)/(W- W,)]

where W is the amplitude of the radial deflection and W; that of the initial imperfections, R
the radius of the shell, L the length of the half-waves of the main buckling deformation, and
Pe, is the critical load for the perfect shell (Per is the classical buckling load Pe 1 = 27TRtucl,
with uc 1 being given by Eq. [9.4]). When the expression in the large brackets in Eq. (9.6) is
approximately 1, Eq. (9.6) is the usual hyperbolic relation that indicates applicability of the
Southwell representation of test data. The conditions of applicability are therefore either
that W and W; are small compared to (21} I 7T 2R), which according to Donnell means small
compared to about five times the thickness of the shell, or that W, is very small compared
to W. Donnell calculated (from Eq. [9.6] and with data from [9.54]) and plotted W versus
(WI P) for a number of cases. These Southwell plots were indeed straight lines, but their
slopes were 15-30 percent less than the classical buckling load. This could be explained, as
suggested by Horton et al. [9.213], by regarding the expression contained in { } of Eq. (9.6),
which is always less that unity, as an attenuation factor on Per.

9.6. 1 Application of the Method in the Case of External Pressure


The first practical application of the Southwell method to shells was carried out by Galletly
and Reynolds at the U.S. Navy David Taylor Model Basin in the mid-fifties (see [9.214]),
in connection with the extensive studies on the buckling of ring-stiffened shells under ex-
ternal pressure in progress there at the time. They began the analysis by pointing out that
for the problem of the general instability of stiffened cylindrical shells subject to hydrostatic
pressure, linear theory apparently sufficed and therefore Southwell's method could be applied
successfully. Then they derived a relation between the circumferential bending strain sh of
the imperfect shell and the applied pressure p:
p
+ ... + am,/ (9.7)
Pmn- P
where s 0 is the axisymmetric circumferential strain of the perfect shell at pressure p and p,,
is the mn-th buckling pressure, i.e., that associated with the mn-th buckling mode fjx)g,(¢),
which for simply supported ends is sin (m7Txl L) sin n¢, and
a,, = a,,j,(x)g,(¢)
where a," is the amplitude of the mn-th component of the initial imperfection. Two typical
circumferential strain versus pressure plots, from measurements on one of a series of ring-
stiffened cylindrical shells tested hydrostatically at the David Taylor Model Basin, are shown
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Southwell's Method for Shells 681

in Figure 9.49. These shells were 8.12 in. (206 mm) in diameter and 18.55 in. (471 mm)
long and had 14 closely spaced machined external rings, ensuring failure by general insta-
bility. One may observe that the plot of s, versus pis nonlinear and note from Eq. (9.7) that
s, will first become infinite at p = Per the minimum elastic buckling pressure Pmn· As Per is
approached, Eq. (9.7) reduces to
p p
Eh = --- amn = - - - acr (9.8)
p,,, - P Per - P
since the other terms on the right-hand side become negligible compared to that associated
with Per· This is the typical hyperbola representation, which can be rewritten in the Southwell
form as
(9.9)
similar to that for a column, Eq. (4.22) in Chapter 4. Sturm [9.38] had already, in 1941,
derived a hyperbola-type relation similar to Eq. (9.8) for the deflection amplitudes of slightly
imperfect unstitiened cylindrical shells under hydrostatic pressure, but had not applied the
Southwell method.
As can be seen in Figure 9.49, the definition of the bending strain s,, there relates to the
circumferential strain of the perfect shell s 0 , which has to be calculated from theory. An
alternative procedure was therefore proposed by Galletly and Reynolds for calculating Per·
Noting that typical measured strain-pressure plots, as in Figure 9.49, are linearly related to
p for a considerable part of the range, one can write

(9.10)
where C is a constant and s* is the nonlinear portion of the total circumferential strain,
which can be determined directly from the measured strain-pressure plots, without theoretical
calculations. Hence from Eqs. (9.7) and (9.10),
- p
s* = -Cp +---all+··· (9.11)
Ptt- P
and, as before, s* first becomes infinite at p = Per· Also, as Po is approached, the magnifi-
cation of the term with Pee in its denominator will make it dominant, reducing Eq. (9.11) to

I
I
I
I
/ GAGE._ 18AT 255 111 {IN'IIER FIBER AT RING No 7)

THEORETICAL CIRCUMFERENTIAL STRAIN


(SALERNO · PULOS)

RUN No.
0 ~ I
• ~ 2
"' ~ 3
• ~ 4
D ~ 5
• - 6

-~~~----~~~--~~-~-2~00~---~~0~---~600~-~-~~--~
STRAIN IN JL- in/in

Figure 9.49 Two typical circumferential strain versus pressure plots for a David Taylor Model Basin
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

ring-stiffened cylindrical shell under hydrostatic pressure (from [9.214))


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
682 Shell Buckling Experiments

p
e* = - - - acr (9.12)
Pee-P
which can be rewritten in the Southwell form, or as
(9.13)
The plot of (e*lp) versus e* should therefore be a straight line, whose cotangent is the
buckling pressure Pee· These Southwell plots (called the slope method by Galletly and Reyn-
olds) for the strain-press plots of the two strain gages of Figure 9.49 are shown in Figure
9.50. Galletly and Reynolds also proposed a second approach, plotting (1/e;,) or (1/e*)
versus p, as in Figure 9.51, the intercept of these curves yielding the buckling pressure Pee·
The intercept method, as they called it, results from the fact that e" or e* becomes infinite
at P = Pc,·
Table 9.1, also reproduced from [9.214], shows that the buckling pressures determined by
slope and intercept methods for all the gages (except those with insufficient data) agree very
well with the experimentally observed failure pressure and are practically all slightly above
it. These 24 strain gages were placed at 15° intervals around the inside surface of the shell
under one of the two middle rings, where roughly the largest deflections of the multilobed
general instability mode should occur. Under external pressure loading this mode has usually
one longitudinal half-wave and a number of circumferential waves 11, depending on the shell
geometry, here 11 = 4. The number of strain gages per circumferential wave, six, is therefore
large enough to yield good results for practically all gages, as seen in Table 9.1, irrespective
of their circumferential location.
In [9.214] the Southwell method was applied to four other models of the series, which
were identical to the one discussed except for varying depth of the stiffening rings. The ratio
of the average of the critical pressures determined by the Southwell method from the mea-
sured strains, using the two approaches to the experimental collapse pressure is presented in
Table 9.2. Again, very good agreement between the Southwell predictions and the experi-
mental failure pressures was obtained for all the shells, the Southwell ones, representing the
corresponding perfect shells, being slightly above the experimental pressures, as expected.
Galletly and Reynolds emphasized that for a successful application of the Southwell
method, sufficient experimental data must be obtained in the neighborhood of the failure
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

load. For general instability of a stiffened shell under external pressure this was not difficult,
since the buckling mode could be predicted, but for other cases problems might arise. Fli.igge

.~~ 2.0
::!-
l~
\1,1 1.0

- 500 -400 -300 - 200 200 300 400 500

GAGE NO. 18

-2.0

-3.0

-4.0

Figure 9.50 Determination of p" by the slope method (Southwell plot) for the strain-pressure plots of
Figure 9.49 (from [9.214])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Southwell's Method for Shells 683

200

160
GAGE NO. 3 AT 30"

80

"'Q 40

:.
z
~ 110 120 130 180
a:
..... PRESSURE psi
"'
:::,.
-40

-80

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
-120
GAGE NO. 18 AT 295 •

-160

-200

Figure 9.51 Determination of Pu by the intercept method for the strain-pressure plots of Figure 9.49
(from [9.214])

had already remarked earlier (see [2.55], p. 493) that for shells under axial compression it
was difficult to determine suitable locations for displacement measurements since there were
many different possible buckling modes with buckling loads rather close to each other. He
added that if care was taken to measure a displacement "which is large for the lowest
buckling mode but small for all competitive modes," the Southwell method "is well appli-
cable." He substantiated this by reference to unpublished experiments by Kromm and Fltigge
in the early forties on cylinders under combined axial compression and torsion. Professor
Fltigge, however, mentioned privately to Professor Horton (see [9.215]) that Southwell plots
were obtained only when torsion was the predominant loading, again a case where the
buckling mode can be reasonably well predetermined.
Galletly and Reynolds' extension of the Southwell method was very successful since they
applied it to closely stiffened cylindrical shells under hydrostatic pressure, for which the
postbuckling behavior is nearly neutrally stable (i.e. imperfection insensitive) and for which
the bifurcation buckling modes are well separated (a behavior that somewhat resembles that
of a column). Also, unstiffened cylindrical shells under external pressure, and to a lesser
extent under torsion, exhibit similar behavior and hence can yield reliable Southwell plots.
Horton and Cundari [9.215] showed this for two steel shells under external pressure tested
by Sturm [9.38] in the late twenties (see Figure 9.52b and Figure 2 in [9.215]).
In Figure 9.52a (reproduced from [9.38]) the measured radial deflection curve and the
computed one for a long, 18 in. diameter, steel shell are shown, and in Figure 9.52b the
corresponding Southwell plot (reproduced from [9.215]) is presented. The buckling pressure
predicted by the Southwell method for the equivalent perfect shell, Psourh = 4.90 psi, is 15
percent above the experimental collapse pressure, Pexp = 4.25 psi, indicating that the shell
had significant imperfections, though the measured out-of-roundness was only 1.35 t. The
computed buckling pressure, Per = 4.70 psi, was close to the Southwell prediction, about 4
percent below it.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
684 Shell Buckling Experiments

Table 9.1 Critical pressures (psi) determined from Southwell plots for ring-stiffened cylindrical
shell under hydrostatic pressure

Gage Per Gage Per


no. no.
Intercept Slope Intercept Slope

f:;, e* e* f:;, s* .s*

1* - 187 172 13*** - - -


2 179 179 171 14** - - -
3 173 173 173 15 172 172 172
4* - 172 171 16*** - - -
5 182 182 173 17 178 178 174
6 168 168 169 18 176 176 173
7* - 171 171 19 174 174 169
8 172 172 178 20 166 166 168
9 175 175 174 21 177 177 174
10 183 183 173 22* - 171 174
11 171 171 174 23 179 179 168
12* - 176 171 24 180 180 174
Average Pu Observed
failure
Intercept Slope
pressure
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

f:;, e* e*

175 175 172 168


* e intersects eO'
**gage inoperative.
***insufficient data.

Another example for cylindrical shells under external pressure is a long. 6 in. diameter,
extruded aluminum alloy tube, from the second series of shells tested by Sturm in the early
thirties [9.38]. The pressure versus measured and computed radial deflection W is shown in
Figure 9.53a (reproduced in part from [9.38)), and the corresponding Southwell plolt is
presented in Figure 9.53b. Here the buckling pressure predicted by the Southwell method,
Psnuth = 8.64 psi, is only 7 percent above the experimental collapse pressure, Pexp = 8.11

psi, since the shell had apparently fewer imperfections. The computed critical pressure for
fixed ends (appropriate on account of the heavy fitted-end bulkheads) Po = 8.25 psi, was
again close to the Southwell prediction, 5 percent below it.

9.6.2 Stanford University and Georgia Tech Comprehensive Studies


As an example of torsional buckling, Figure 9.54 (reproduced from [9.94)) shows a Southwell
plot for a steel shell of a series tested in torsion and combined torsion and axial compression

Table 9.2 Southwell plots applied to ring-stiffened cylindrical shells under


hydrostatic pressure: Pulp1 as given by two methods (p 1 is the
experimental failure pressure)
Model Southwell
no.
Intercept Slope
1.040 1.023
2 1.010 1.011
3 1.040 1.033
4 1.034 1.033
Copyright Wiley
5 1.020 1.023
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Southwell's Method for Shells 685

LAP JOINTS RIVETED AND SOLDERED


6 D=IBin t=0.063in. L•19ft.-8in
DISTANCE BETWEEN JOINTS•3fl- II in
w 0 ·0~ in.

5 PeR INCLUDING SEAMS • 4.70 psi


fiNAL COLLAPSE

·~ 4

- - COMPUTED, SEAMS INCLUDED.

RADIAL DEFLECTION, WI ;nchesl

(a)

10

a R!t • 143

'1-
Q 6
"
.0 0 TEST PExP =4.25 psi
4
"''c 0
0
0
GRAPHICAL PsourH •490 !SOUTHWELL)
THEORETICAL PeR •4. 70
Q. 2
PsourH/ PCR • 1.04
~ PSOJT I PExP• I. 15
0
0 5 10 15 20 25 30 35 40
RADIAL DEFLECTION W (;n, o-2)

(b)

Figure 9.52 Southwell plot for a long 18-in. diameter cylindrical steel shell tested by Sturm under
external pressure: (a) radial deflection curve (from [9.38]), (b) Southwell plot from the
data of (a) (from [9.215])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

by Craig at Stanford University in the late stxttes. The shells of (Rit) ~ 790 were made
from shim stock by a simple wraparound and seam process, which, however, still yielded
good shell specimens whose measured initial imperfection magnitudes were less than half
the thickness. The clearly defined Southwell plot in the figure used deflections measured at
the center of a buckle. The predicted buckling load was about 1 percent above the observed
one. Reliable Southwell plots were also obtained for the other tests in this series. Another
example for torsional buckling is the Southwell plot for the orthotropic (single-layer glass-
epoxy) cylindrical shell tested by Bank [9.216], shown in Figure 6 of [9.215] or Figure 3.18
of [4.48].
In the late sixties and early seventies Horton and his co-workers at Stanford University
and Georgia Institute of Technology carried out comprehensive studies of the application of
Southwell's methods to shells (see f9.213], [9.215]-[9.219]). In addition to the primary task
of establishing the applicability and limitations of the method, these studies also aimed at
development of non-destructive test techniques for buckling of shells. The Southwell method
appeared to be a likely candidate for NDT uses, but it proved to have significant limitations.
In order to theoretically justify the use of the method for shells, Horton and Cundari
[9.215] applied linear Donnell equations, in Batdorf's modified form, to cylindrical shells
with imperfections of the same modal shape as the anticipated deformations. For axial com-
pression, external pressure, hydrostatic pressure and torsion, they obtained expressions of
Southwell type, similar to Eq. (9.8), which can be rewritten in the usual form to yield
Southwell plots. They also obtained a similar expression for spherical shells under external
pressure. Horton and Cundari therefore concluded, from this and from the extensive exper-
imental
Copyright Wiley
evidence, such as Figures 9.49-9.54 Licensee=McDermott
Provided by IHS Markit under license with WILEY
and others to be discussed below, that the
Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
686 Shell Buckling Experiments

UNOER UNIFORM EXTERNAL PRESSURE ON SIJES ONLY.


MATERIAL ~iS-H ALUMINUM ALLOY
DIAMETER 6.00 in. ,SliELL THICKNESS: 0.042 in
NET LENGTH 114 in.
10.0
~·38 ~· 72 Pc 1~ 8.25 psi - EDGES FIXED

.

MEASURED PRESSURE AT COLLAPSE--=-t


8.0
·;;
<>. "l;DGES FIXED
!!::
... 60
~

~
Q.
_J - - MEASURED DEFLECTION,
<(
- - - - COMPUTED DEFLECTION.
~ 2.0
...
1-

~ 0~--~~~~--~~~~--~~--~~--
0 0.2 04 0.6 08 10 1.2
RADIAL OEFLI'CTION IN INCHES W
(a)

0.10
Wtp
.08

.06
Rlt = 72

/ PEXP

PsouTH
=8.11
=8.64
psi
psi
PeR =8.25 psi THEORETICAL
P souTH I PeR = 1.05
p SOUTH I pEXP • l.07

0 .2 .li .8 .10 INCH


w
(b)

Figure 9.53 Southwell plot for an extended aluminum alloy tube tested by Sturm under external pres-
sure: (a) radial deflection versus pressure (from [9.38]), (b) Southwell plot from the mea-
sured data of (a)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

8.-------------------------------~

6 Tcrit = 26.83 l't-lbs.

·e 4

SOUTHWELL PLOT

2 '"""" ·~· ~"

/
0 0.1 0.2 0.3
8n {milstft-lbl

Figure 9.54 Southwell plot, for a thin steel cylindrical shell under torsion, made at the center of a
buckle (from [9.94])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Southwell's Method for Shells 687

Southwell method is "of absolute generality." This statement is obviously too strong since
the use of Donnell's linear equations implies that the linearized approximation fairly correctly
describes the behavior of the shell, which occurs only for those types of loading and shells
whose postbuckling behavior is nearly neutrally stable.
Unstiffened cylindrical shells under external pressure, hydrostatic pressure or torsion, as
well as closely stiffened shells also under axial compression, come into this category and
hence lend themselves to a reliable application of Southwell's method. For other cases,
Roorda's advice of care in the interpretation, discussed in Chapter 4 (see also [4.47] and
[9.220]) should be heeded. However, careful experimentation often yields very good South-
well plots also in such cases, as shown for example in [9.215] for cylinders under axial
compression and spherical shells under external pressure and point loads.
Figure 9.55 shows a Lundquist plot (a modification of the Southwell plot, discussed in
Chapter 4, which uses a shifted origin in the load-displacement curves to eliminate the low
load non-linearities in the plot) for a well-made unstiffened aluminum cylindrical shell, of
(Rit) = 335, under axial compression. High-resolution non-contact Fotonic sensors were
used for the displacement measurements in the test. The buckling load predicted from the
plot agreed very well with the classical one, less than 2 percent difference, and was only a
few percent above the measured one. The high knock-down factor indicates that the specimen
was indeed well made and the loading rather uniform. Horton and Cundari showed also (see
Figure 5 of [9.215]) that the predicted load was not sensitive to the location of deflection
measurement, provided these points were chosen judiciously.
In the mid- and late sixties, Tenerelli and Horton studied the local buckling of ring-
stiffened cylindrical shells, with (R/ t = 400, under axial compression [9.202] and obtainerd
very consistent Southwell plots. The investigation included four magnesium and three stain-
less steel specimens, but since an internal mandrel, restricting the buckle depth, was used,
many repeated tests were possible. The lateral shell wall displacements prior to buckling
were measured precisely at random points with a non-contact Fotonic sensor. Since the tests
were repeated, the load deflection curves for the center of a buckle could be obtained by
placing the sensor at a point that appeared as the center of the buckle in the previous test.
The measured prebuckling load deflection curves showed that buckles initiated relatively
early in the loading process and that, after an initial outward displacement due to the Poisson
effect, the center of the buckle deflected inwards until failure occurred, as can be seen in a
typical load deflection curve for the steel shell S-3 shown in Figure 9.56 (reproduced from
[9.202]). The inward displacement of the buckle center resembles the lateral motion of a
column and hence lends itself to the Southwell representation, as shown in the figure.

3.0.---------------,

0
2.5

.'g 20
.
...::::
0
.E 1.5
-;E
IL '
:::::' l.O Pslopo • 240 lb.
0
"P P0 • 725 lb.
~
l:P • 965 lb.
0.5

0 L_~s~o-~t~oo~~,s~o-~2oo~~2W~~~~o
(8- l.lol In • 10" 1

Figure 9.55 Lundquist plot (a modification of the Southwell plot) for a thin aluminum cylindrical shell
under axial compression (from [9.215])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
688 Shell Buckling Experiments

11;oo

~-~ .. 1200
·!;
z
Q

~~
Li9
w
0
~ 400
oolo..

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
8 0 2 4 6 8
D,EFLEC/ION. Ticrons,
0 2 4 6 8 10 12 14 16
8, (microns), lor Southwell plot

Figure 9.56 Load versus lateral deflection and Southwell plot for center of a buckle of the initial
instability pattern, for a steel cylindrical shell S-3 under axial compression (from [9.202])

The Southwell plots were obtained by Tenerelli and Horton with an extension of the
method which they called "displaced Southwell." This extension (described in detail in
Appendix I of [9.202]) takes the reversal point of the radial shell wall displacement to inward
motion (A in Figure 9.56) as the new zero deflection point, keeping the true load value. The
displaced Southwell thus disregards the initial outward displacement, which is not due to
the buckle formation. This extension is similar to Lundquist's modification of the Southwell
plot, except that there the origin of both displacement and load are shifted. Indeed, if one
plots here Lundquist plots instead of the displaced Southwell ones, one obtains very similar
predictions, within 2 percent.
For this shell, two other buckles were similarly observed, and the resulting Southwell
(displaced Southwell) plots yielded predicted critical loads that were within about one percent
of each other. On one of the buckles a point shifted circumferentially from the buckle center
was also monitored. It showed smaller inward displacements, but the corresponding South-
well plot yielded again a predicted load within less than 2 percent of the other values. Hence
again the location of the measurement point was not very critical, provided it was in the
vicinity of the center of a buckle. Displacement measurements at the nodes did not show
any reversal to inward motion but continued to deflect radially outward, the rate of deflection
increasing as collapse was approached. They were therefore not suitable for a Southwell
representation. Thus, in the absence of a mandrel permitting repeated buckling, which reveals
the positions of buckle fonnation, the choice of appropriate measurement points remains a
problem, indicating the necessity of a large number of such points.
Harmonic analysis, as originally proposed by Donnell [4.26] was further developed by
Horton and Craig [9.94] to overcome the difficulties of data reduction from a large number
of measuring points. They showed, by examining the data presented by Arbocz and Babcock
[9.221], who performed an harmonic analysis of the displacement measurements for their
axially compressed shells, and from an harmonic analysis of their own tests under torsion,
that consistent Southwell plots could also be constructed for the predominant higher har-
monics. The slopes of these plots for several harmonics were found to be in fair agreement
in both cases, and it was usually not difficult to choose the plot with the highest slope (i.e.
lowest Psouth) as that yielding the critical load. Hence harmonic analysis apparently facilitates
the data reduction for application of the Southwell method. These critical loads were about
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Southwell's Method for Shells 689

10 percent below the classical ones for axial compression, which they were supposed to
predict. Horton and Craig attributed this difference to load non-uniformity. The experimental
buckling loads under axial compression for these well-made, but still imperfect, shells were
60-70 percent of the classical ones. The fact that the Southwell predictions were often nearer
to the experimental buckling loads than to the classical ones was also observed in other
studies (for example in [9.222] or [9.229]). However, even with harmonic analysis a large
amount of data reduction has to be carried out, which obviously detracts from the usefulness
of the Southwell method.

9.6.3 Application to Spherical Shells


Horton and Cundari also demonstrated the applicability of the Southwell method to different
experimental studies of spherical shells and spherical caps subjected to external pressure or
to a concentrated load (see Figures 7-9, 11 and 12 of [9.215]). Lebouc [9.218] presented
more details and many additionhal Southwell plots. Figure 9.57 (reproduced from [9.218])
presents a typical Southwell plot for one of the hot-spun magnesium clamped spherical caps,
with (Rit) = 210 and geometry parameter A = 4.94, tested under external pressure by Kaplan
and Fung at Caltech in 1954 [9.223]. The Southwell plot yielded here Psouth = 121 psi,
whereas the experimental buckling pressure was Pexr = 72.5 psi and the classical buckling
pressure for the corresponding perfect complete sphere Pc1 = 179 psi. Hence Psouth was about
32 percent below Pe~ and about 40 percent above the measured Pexr• i.e. Psouth was significantly
nearer to the experimental buckling pressure than was the classical one. Similar results were
obtained from Southwell plots of five additional spherical caps tested by Kaplan and Fung,
the critical pressures yielded by the plots falling between the classical prediction for a perfect
sphere and the experimental value. Table 9.3 shows the relative position of the Southwell
critical pressure for the six caps analyzed by Lebouc [9.218].
It is apparent that the Southwell pressures are 19-40 percent above the experimentally
observed ones but 26-45 percent below the classical buckling pressure for the corresponding
perfect complete sphere. The Southwell procedure does not converge onto Pe~• as it should,
but to a critical pressure about 25-45 percent below it. Comparison with symmetric and
asymmetric nonlinear theories (see for example Kaplan's review, [9.224]) shows that the
Southwell pressures happen to be much nearer to the nonlinear predictions (8 percent above
to 19 percent below the nonlinear theoretical value). This may be encouraging to the exper-
imenter, but needs further analytical justification.
The Southwell technique has also been applied to complete spherical shells (see for ex-
ample Figure 9.58 and Figure 42 of [9.218], or Figures 11 and 12 of [9.215]). However, for

06

0.5

). • 4.94
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

0.4
Psouth ~:: 121 psi

Pel 179 psi


O.J
P exp • 72.5poi

02
· .019 .020 .021 .022 .023 .024 .025 .026 .027 .028
2
(f)/ P('E" )(ft
Figure 9.57 Southwell plot for one of the hot-spun magnesium spherical caps tested under external
pressure by Kaplan and Fung [9.223], (from [9.218])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
690 Shell Buckling Experiments

Table 9.3 Relation of Southwell plot critical pressure to the classical theoretical and experimental
buckling pressures for typical spherical caps from the series tested by Kaplan and Fung in
1954 (NACA 3212)
Specimen No. 3 4 5 7 10 16

above Pcxp} 22 38 40 25 19 39
Ps"'"" (percent)
below Pc~} 45 42 32 26 36 38
Psooch (percent)
below p,,} 40 38 37 24 24 23
p""""'""' (percent)

A 4.6 4.8 4.94 5.57 7.22 8.98


Symmetric theory Asymmetric theory

complete spherical shells the Southwell plots predicted approximately the classical buckling
pressures for perfect spheres, provided the local reductions in modulus, observed in the initial
linear pressure-deflection plots of these shells, were accounted for. For example, for Thomp-
son's polyvinyl chloride shells, with (RI t) = 21 [9.225], the Southwell plot for one of them
shown in Figure 9.58 yielded, employing the locally observed modulus of elasticity, (Psou<h
I Pc~) = 1.00. Similarly, the Southwell plot for Lebouc's electro formed nickel shell, with (R
It) = 2000 [9.218], gave (Psouthlpci) = 0.92 when the local modulus was used.
Application of the Southwell technique to spherical caps subjected to a point load at the
apex, or its combination with external pressure, also produced very consistent plots (see for
example Figures 7 and 8 of [9.215] or Figures 27-29, 33-35 and 38 of [9.218]), that yielded
buckling loads somewhat above the experimentally observed ones. For example, for one of
Ashwell's aluminum alloy shallow spherical shells (Figure 12 of [9.226]), with (Rit) ~ 320,
Psouth was about 25 percent above the experimental buckling load, but also about 17 percent
above the large-deflection prediction for a perfect shell. Or, for typical plastic shells of the
Evan-Iwanowski, Cheng and Loo test series [9.227] Psouth (see Figures 33. 34 and 35 of
[9.218]) was 23-42 percent above the experimental value.

9.6.4 Application to Stiffened Cylindrical and Conical Shells


The Southwell method was also extensively applied in the experimental investigations on
buckling of closely stiffened cylindrical and conical shells carried out at the Technion Aircraft
Structures Laboratory in Haifa in the sixties and beginning of the seventies [9.149], [9.206],
[9.212], [9.222], [9.228], [9.229] and [9.298]. These experiments are examined in Chapter
13, whereas here only the application of Southwell plots is discussed.
The earlier series of tests [9.206], [9.212] studied closely integrally ring-stitTened cylin-
drical shells made of alloy steel, with (Rit) = 620-740, subject to axial compression. As
mentioned in Section 9.5, these shells were well "covered" with strain gages, which became
lively near buckling simultaneously at many locations and hence presented the data required
for Southwell plots.
For most specimens these plots could be readily constructed at various locations. The
buckling loads were predicted from these Southwell plots, using the slope and intercept
methods as in [9.124]. They were between 3 and 19 percent above the experimental values
(except equal in one shell), and between 4 percent above and 26 percent below the linear
theory general instability loads for perfect shells. For the more near-perfect shells, whose
experimental buckling loads approached the linear theory predictions PG,, Psou<h was close to
PG, (and sometimes slightly above it) while only slightly above Pexp· This can be seen in
Table 9.4, which shows these ratios for some typical shells (taken from Table 2 of [9.212]).
Comparison of the Southwell loads obtained from strain gages at various locations showed
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Southwell's Method for Shells 691

.37

.33

.31

.29

.27 [Psouth/ Pel (wilh locolly observed E!] = 1.00


.2~

0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4


NORMAL DISPLACEMENT 8
Figure 9.58 Southwell plot for Thompson's experimental data on a complete spherical shell under
external pressure (from [9.218])

that, with few exceptions, those obtained at locations far away from the final buckling pattern
location were similar to those near it.
The Southwell method was also applied to experiments on steel ring-stiffened conical
shells under axial compression [9.229]. Both the slope and intercept method were applied,
but whereas the slope method could always be used, the intercept method failed in some
cases. Since the strain gages exhibited almost linear behavior up to buckling in these tests,
there were insufficient points near the buckling load for very good Southwell results. The
intercept method resulted in less scatter and higher critical loads than the slope method,
whereas the slope method always yielded buckling loads very close to the experimental ones
(see for example Figures 9.59 and 9.60, showing the results for one shell). But the loads
obtained by both methods were always smaller than linear theory predictions, similar to the
Southwell loads in unstiffened and ring-stiffened cylindrical shells discussed earlier.
More recently, the Southwell method has been applied by Foster and Tennyson at the
University of Toronto to epoxy stringer-stiffened cylindrical shells [9.230]. Their shells, of
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(Rit) = 68-118, were subjected to axial compression, external pressure and their combi-
nation. Only one strain gage was used in each shell. In a manner similar to that proposed
by Galletly and Reynolds in the fifties (see Figure 9.49), the bending strain was taken to be
the difference between the asymptote to the load deformation curve and the curve itself. This
apparent bending strain was then used as the deformation parameter in the Southwell tech-
nique. A typical load deformation curve (for cylinder no. 1, under combined loading, with

Table 9.4 Relation of Southwell plot predictions to classical theoretical


and experimental buckling loads for Technion ring-stiffened
cylindrical shells under axial compression (from [9.212])
Shell no. Psouth/ Pe . . p P'c.nuth/pc;" P = P"r/ Pc;,
MZ3 1.10 0.90 .82
MZ5 1.05 0.96 .88
MZ8 1.19 0.89 .75
MZIO !.12 1.03 .93
MZI2 1.07 0.78 .72
MZ13 1.00 1.04 .95
MZ18 1.09 0.75 .69
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
692 Shell Buckling Experiments

M 3-5A

:~:::~~~
_,w 0.02 'l( v .

0
1800 2600 3400 42.00 5000 5800
P(kgl
v GAGE NO.9 f'sou~.~5400kg Psoutn.~ 5510 kg
o GAGE NO.I I Pso..llt•5660kg lA:rlth. •7340kg
A GAGE NO. 12. f's0UIIr5660kg U'!:rlup~5018 kg
0 GAGE NO.I4 Psout,,~5400t,.
o GAGE NO. 17 f'soullr5440kg
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 9.59 Critical (Southwell) loads for ring-stiffened conical shell M3-5A obtained by the intercept
method (from [9.229])

collapse occurring at an external pressure of 40.6 kPa and an axial load of 37.0 kN) is shown
in Figure 9.6la and the corresponding Southwell plot in Figure 9.61b. Though the straight
lines were fitted by eye, a repeatability within 1.25 percent of the critical load was obtained.
In this case only the last few points before buckling were useful in the construction of the
Southwell plot, but the predicted critical load was very close to the measured load. Most
Southwell plots for various load combinations and cylinders nos. 1-3, which buckled glob-
ally (by general instability), were like those in Figure 9.61 b. In cylinder no. 4, the thinnest
of the series, a tendency to buckle locally was observed. When local panel buckling occurred,
the load deformation curves were like tho:;e in Figure 9.62a. The fitting of a straight line to
the corresponding Southwell plot (Figure 9.62b) then presented a problem. If, however, one
remembered that the Southwell technique applies to small deformations only, and that there-
fore only the lower strain values had to be considered (as suggested by Foster and Tennyson),

0.06-,--------, 0.06,--------,
17

0 o'-'---'--:,~2o::-'---'--='240~~360
E•x 106
M3-5A

v GAGE N0.9 PsOU!h • 5550 kg


0 GA<iiENO.II Ps.ulh • 5550 kg
A GAGE NO.I2 PSOU!n • 5550 kg
0 GA~IE NO.I4 PSOU!n ~ 5000 kg
0 GAGE NO.I7 Ps.u1n•5150 kg
Ps.uth "'"5360 kg
(Paltn •7340 kg
(Pcrl., 0 ~ 5018 kg

Figure 9.60 Critical (Southwell) loads for ring-stiffened conical shell M3-5A obtained by the slope
(modified Southwell) method (from [9.229])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Southwell's Method for Shells 693

30

z
~
20

g
10

0o~--~200~~0--~4~0~00~----~oo~o~o--~oo~oo

STRAIN READING (f-Ltii)

-004

~z
.
i: -003
0::
>-0
"'
~3
.
0
z
"'"'
(b)

-!50 - 100 -150


BENDING STRAIN (f-Ltii)

Figure 9.61 Southwell plot for stringer-stiffened spin-cast cylindrical shell no. l, under external pres-
sure of 40.6 kPa and axial compression of 37.0 kN (from [9.230]): (a) load-deformation
curve, (b) Southwell plot from the data of (a)

a fair estimate of the initial load was obtained, about 13 percent above the initiation of
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

buckling (detected visually). The suggested restriction to lower strain values is a remedy to
the curving of the Southwell plot, which may appear when large deformations precede buck-
ling, and as a matter of fact follows Roorda's advice mentioned in Chapter 4 (see also [4.47]
and [9.220]). Foster and Tennyson also presented interaction curves for measured collapse
conditions and those obtained by the Southwell method (see for example Figure 9.63 for
cylinder no. 2), which demonstrated good agreement. When local panel buckling occurred,
care had to be taken to separate it from global buckling, and then good results were obtained.
In practically all the cases, the Southwell predictions were a little higher than the measured
loads (within 10 percent), as was to be expected, since they represent the critical loads of
the corresponding perfect shells.

9.6.5 On the Applicability of Southwell's Method to Shells


In recent years the Southwell technique has been applied successfully to many types of
shells, as for example to the buckling behavior of a drop-shape underwater storage vessel
(see [9.231]), non-destructive estimation of the buckling loads of buried flexible pipes [9.232]
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
694 Shell Buckling Experiments

0
g

2000 4000 6000 8000


STFIAIN READING (fL#.)
-20

"
"
-15
"
"
"
1Jz-10 {;

z
<i
{;

a:
1-0
<flo(

"'9
~
z -05 (b)
Ill

2000 -4000 6000


BEtiDING STRAIN (fL~)

Figure 9.62 Southwell plot for stringer-stiffened spin-cast cylindrical shell no. 4, under axial com-
pression only, of 33.0 kN (from 19.230]): (a) load-deformation curve, (b) Southwell plot
from the data of (a), considering only the lower strain values

or tests of graphite-epoxy curved stiffened cylindrical panels loaded in axial compression


[13.134].
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Summarizing this rather comprehensive discussion of the applicability of Southwell's


method to shells, it can be concluded that, contrary to the warning of Roorda [4.47] and the
pessimistic evaluation of Bushnell [4.48], the Southwell method has a wide range of appli-
cability in shell buckling. For the types of loading and shells whose postbuckling behavior
is not far from neutrally stable, the method works very well, and it can also be employed
for non-destructive prediction of buckling loads. But even when the postbuckling behavior
is unstable, it can be successfully applied if the unstable postbuckling behavior is relatively
mild, as for example in closely stiffened shells, provided sufficient data has been acquired.
Though here the buckling loads predicted by the Southwell technique are usually nearer to
the experimental values than to the classical ones for the corresponding perfect shell, use of
the technique as a tool for non-destructive prediction calls for caution in these cases.
The other important use of the Southwell method, for smoothing of data in parametric
studies,
Copyright Wiley discussed already in Chapter 4, has found
Provided by IHS Markit under license with WILEY
widespread application to shells in recent
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Cylindrical Shells under External Pressure, Bending or Torsion 695

"
~30
LIJ
a:
::>
tf)
tf)
::! 20
Q.
..J
<(

~ A MEASURED
a: 10 0 SOUTHWELL
METHOD

0 L __ __ i_ _ _ _~~~~~~~~~
0 0.000 20.000 30.000 40.000 50.000
AXIAL LOAD (N)

Figure 9.63 Interaction curve for stringer-stiffened spin-cast cylindrical shell no. 2 under axial com-
pression and external pressure, for measured collapse conditions and those obtained by
the Southwell method (from [9.230])

years. The method has even been successfully employed for data smoothing in plastic buck-
ling (see for example [16.146]), as will be discussed in Chapter 16.

9. 7 Cylindrical Shells under External Pressure, Bending or


Torsion
9. 7. 1 External Pressure Loading
As was seen in the historical introduction of this chapter, external pressure loading was the
primary source of instability in pipes and shells in the 19th century and in the first decades
of the 20th, and is even today one of the prime causes of buckling. It has therefore been the
subject of extensive experimental and theoretical studies, in particular the classical experi-
ments carried out in the twenties and thirties by Saunders, Windenburg and Trilling at the
U.S. Navy David Taylor Model Basin, Tokugawa in Japan and Sturm at the University of
Illinois [9.33]-[9.38], which were discussed earlier.
In Section 9.6 it was noted that buckling under external pressure loading exhibits the least
violent behavior in cylindrical shells and has a nearly neutrally stable postbuckling behavior.
There is therefore fairly good agreement between theory and experiments (see for example
Figure 9.67), and the Southwell method also works very well. Recall that there are two
common cases of external pressure loading: lateral (or radial) pressure and hydrostatic pres-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

sure which includes an axial component. For short cylindrical shells the critical hydrostatic
pressure is lower than the lateral one, but for medium length and long shells (Z =
'\11=7 (U!Rt) > 100) the difference is negligible. The comparison in Figure 9.67 is for
hydrostatic pressure.
For buckling under external pressure, the boundary conditions considerably influence the
results even for long shells. This led to extensive theoretical studies on the effect of boundary
conditions on the stability of cylindrical shells in the sixties, which were also motivated by
the large scatter of buckling loads under axial compression. These investigations focused on
the influence of the in-plane boundary conditions. An earlier study by Singer [9.233] already
indicated that elastic axial restraint of the edges significantly increased the buckling pressure.
Then in 1964 Sobel [9.234], Thielemann and Esslinger [9.235] and others showed that
whereas rotational restraint (clamping) did not significantly increase the critical pressure,
unless the cylinders were quite short, axial restraint at the edges (prevention of warping)
notably increased it for intermediate-length cylinders, by up to about 50 percent. This per-
mitted better interpretation of test results and presented an important guideline to designers
in the layout of their end rings or covers.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
696 Shell Buckling Experiments

Since cylindrical shells subjected to external pressure can usually support loadings con-
siderably higher than the buckling pressure, their postbuckling behavior has been widely
studied experimentally and theoretically. A good example are the experiments of Yamaki
and Otomo [9.1 09], which preceded their experiments on cylindrical shells under axial com-
pression discussed in Section 9.4.
Since the external pressure required for buckling of thin shells, made of Mylar-type Diafoil
polyester film, is much less than atmospheric pressure, the tests were carried out by evacu-
ating the air from the inside of the cylinder. This also allowed visual observation of the
formation of buckling waves. A schematic diagram of the test setup is shown in Figure 9.64.
The test cylinder (3) is again placed on the base plate (2) of a tension test machine, as in
Figure 9.31 for axial compression, but the loading is now applied by removing the air from
the inside of the cylinder via a vinyl tube (13 ), a pressure transducer ( 14) and an air reservoir
(15) with a volume of 25 L to reduce pressure fluctuations. The pressure in the reservoir
(15) is controlled by a hand pump (16) and a vacuum pump (17). The pressure transducer
(14) has a capacity of 30 kPa ('=0.3 kg/cm2 ) and is calibrated with a manometer (18).

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
The deflection-measurement apparatus (5)-(9) is the same as that employed for axial
compression in Figure 9.31 [elements (15)-(20) there], permitting measurement of displace-
ments up to 8 mm while exerting negligible forces on the surface of the shell. The edge
shortening of the shell is measured here by a differential-transformer type displacement
transducer (10), with reference to the cross-head of the test machine, and a dial gage (12)
is used for calibration. Vertical displacements of ± 2.5 mm can be measured with an accuracy
of I fLm. A six-element dynamic strain meter is also used here to measure the signals from
the transducers.
The test setup has since been modified (see [2.44], pp. 182-198), mainly by increasing
the volume of the reservoir to 230 L and adding a stitiening steel top plate with appropriate
counterbalance weights; but the more recent tests and results are very similar to the earlier
ones and all can therefore be considered as belonging to the same group.
The relation between the applied hydrostatic pressure p and the end shortening 8 is re-
corded and the path for a typical shell no. 1000-5 (of the more recent test series, [2.44],
with (Rit) = 405 and Z = 1000) is shown in Figure 9.65a. With increase in pressure, 8
increases linearly up to the critical pressure Poxp = 1.21 kPa ('= 0.176 psi) point (A), where
the shell buckles and snaps into the stable postbuckling state (B) with the circumferential
wave number N = 10. With further increase in p, 8 again increases almost linearly up to
the maximum pressure Pmax = 1.7 kPa, point (C). If p is now decreased, the shell traces
back the original path but continues past point (B) till it reaches point (D), the lowest
postbuckling pressure, and then snaps back to (E) and from there retraces the prebuckling

Figure 9.64 Schematic diagram of Yamaki's te·st setup for hydrostatic pressure (from [9.109], courtesy
of Professor N. Yamaki)

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Cylindrical Shells under External Pressure, Bending or Torsion 697

c
1.!5
p
(kPa) -~
--/
1.0
\-
-
P0 = 1.40 kPo N = 10
0.5 lPoloxp= 1.21 kPo Q5
=0.86 Pc

0
0 50 8 (,u.m) 100 -I 0 2 wmlmm)4

(O) (b)

Figure 9.65 Postbuckling end shortening and normal deflections versus pressure for a typical Diafoil
cylindrical shell no. 1000-5, with Z = 1000 (from [2.44], courtesy of Professor N. Ya-
maki): (a) p versus end shortening 8, (b) p versus maximum inward and outward deflec-
tions W,

path. A similar snap-through at buckling and unsnapping at the lowest postbuckling pressure

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
was already observed in 1956, in tests on thin cylindrical shells made of spring steel, at the
U.S. Navy David Taylor Model Basin [9.236].
If the pressure is increased beyond (C), a secondary snap-through buckling is observed
(not shown in Figure 9.65a) with strong torsional deformations. Similar failures with torsional
patterns were also observed in hydrostatic pressure tests on thin Mylar shells at the DLR
Braunschweig (see [9.74] or [9.115]). One should note that the shell can sustain a pressure
40 percent above the buckling pressure.
In Figure 9.65b the maximum inward and outward deflections are plotted versus the ap-
plied pressure. The large deflections occurring at buckling are clearly seen. With further
increase in pressure p the inward deflection continues to grow, while the outward deflection
(to the left in the figure) soon attains its maximum and then even slightly decreases. When
the pressure is decreased the paths are retraced, except for the deviation to the lowest post-
buckling pressure, as in Figure 9.65a.
Figure 9.66a shows the circumferential distribution of the deflection along the shell mid-
section for the same specimen no. 1000-5 and for two postbuckling pressures p = !.23 kPa
('= 0.178 psi) and p = !.65 kPa ('= 0.239 psi). Contour representations for the postbuckling
waveform corresponding to these two pressures are shown in Figure 9.66b and c.
It may be noted that the number of buckling waves remains constant in the postbuckling
region (here N = I 0) until the torsional failure pattern appears at very high hydrostatic
pressure. As pointed out in Section 9.4, this constant number of waves is typical for cylin-
drical shells under external pressure.
Yamaki and Otomo's experimental results for the critical-pressure parameter k,,
(= Pco RU!7r 2 D) are plotted in Figure 9.67 versus the Batdorf shell geometry parameter
Z (= VT- v 2 (U!Rt)), together with previous experimental results (of [9.35], [9.39] and
[9.237]). The linear theory predictions for classical simple supports (SS3: w = w," = N, =
v = 0) and for complete clamping (C4: w = w,, = u = v = 0), which is nearer to the test
boundary conditions, are also shown for comparison. As mentioned earlier, the agreement is
fairly good for practically all the experimental results, especially if compared to the theo-
retical values for simple support. Those of Yamaki and Otomo fall between 80 and 90 percent
of the fully clamped predictions. One may recall that the difference between the two theo-
retical curves is primarily due to the axial restraint imposed by the clamping (u = 0 instead
of N, = 0), whereas the rotational restraint affects only short shells with Z < 100. The
experimental wave numbers also agree well with the theoretically predicted ones, being 85

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
698 Shell Buckling Experiments

W(mm) r-----·-----~ 0 0
-~~ LP ~ - - ~ ~ ~ ~
~ ,
vv
4E IV VV'J\IVVV
1.23 10 596

k/V\JVVVVVV\7' 1.65 10 119


p (kPa) N 8(,.ml
o• 90° 180" 270• 8 360°
(OJ

(C)

Figure 9.66 Postbuckling characteristics of shell no. 1000-5 (from [2.44], courtesy of Professor N.
Yamaki): (a) circumferential distribution of the deflections along the shell mid-section,
(b) postbuckling waveform at p == 1.23 kPa, postbuckling waveform at p = 1.65 kPa

to 95 percent of the values for fully clamped edges (see Figure 3 of [9.109] and Figure 3.34
of [2.44]).
The scatter of the previous results (of [9.35], [9.39] and [9.237]) for Mylar, steel or
aluminum shells is of the same order and may probably be attributed to the effect of initial
imperfection, variations in axial boundary restraints and uncertainty in modulus of elasticity.
Yamaki and Otomo's results exhibit less scatter, indicating good quality control of specimens
and test conditions.
The comparison between postbuckling theory [9.190] and experiment for external pressure
is presented as pressure versus end shortening in Figure 9.68 for a typical shell with Z =
1000 (shell no. 1000-5 of the more recent test series, [2.44]), whose pressure-displacement
behavior has been shown in Figures 9.65 and 9.66. The correlation between experiment and
theory in the deep postbuckling region in Figure 9.68 is very good, better than in the case
of axial compression (Figure 9.35a). Of the theoretical curves computed by Yamaki, only
the curves for N = 10, the experimental] y observed wave number, and for N = 11, the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

200r----------------------,

100
W=W,_ •u.;:v•O-----~

50 wEw,..,=N ..=v:o·--------
kp

20

10 e YAMAKI AND OTOMO (Ref. 9109)


E>: WINOEN~G AND TRILLING (Ref. 9. 35)
0 8: STURM (Ref. 9 38)
o.· WEINGARTEN AND SEIDE (Ref.9.237)
0

Figure 9.67 Theoretical and experimental results for the buckling of cylindrical shells under hydrostatic
pressure (from [9.109], courtesy of Professor N. Yamaki)

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Cylindrical Shells under External Pressure, Bending or Torsion 699

kp

GO ----EXPERIMENT

kp. p tRL~,.2o) Figure 9.68 Comparison between postbuckling theory of


!• ll (RILl) [9.190] and experiment for Diafoil cylindrical
20L-~~--~~~~~~
shell no. 1000-5, with Z = 1000, under hydro-
0 0.1 0.2 8 03 static pressure (from [2.44], courtesy of Professor
N. Yamaki)

critical wave number for the perfect shell, have been plotted here. Except for the snap-
through at buckling and the "unsnapping," the experimental results correlate well with the
predictions, indicating again that in the deep postbuckling region the geometric imperfections
barely influence the behavior of the shell.
The test installation of a similar series of experimental investigations on thin Mylar cylin-
drical shells of (RI t) "= 170-800 under external pressure, carried out a few years later at
DLR Braunschweig [9.115], also employed partial evacuation of a large vessel by a vacuum
pump to provide external pressure loading on the shell. A large reservoir served again to
reduce the pressure fluctuations. A differential pressure transducer measured the external
pressure, and three circumferentially located displacement transducers recorded the axial
shortening, yielding a load-shortening curve on the X- Y recorder. A typical load-shortening
curve for a medium length shell, of (Rit) "= 400 and (LIR) = 2.25, is shown in Figure 9.69.
It is a typical curve for external pressure loading and closely resembles Figure 9.65a.
As the shell walls become stiffer, the external pressure loading has to be applied by water
or oil pressure in a heavy pressurized test chamber. A typical modern example is that em-
ployed for buckling experiments on steel cylindrical shells at the University of Essen [9.238],
shown in Figure 9.70. A schematic diagram of the test system is presented in Figure 9.71.
The test rig consists of the thick-walled test chamber whose top is lifted for the installation
of the test specimen (see Figure 9.70a); the end plates, which support the test shell; the
loading system (Figure 9.70b) and the different measurement systems, connected through
appropriate outlets.
The hydrostatic pressure applied to the test specimen is provided by a servo-hydraulic
loading system that supplies a controlled quantity of pressurized water to the chamber. In
this subsystem, the piston of the pressure supply is moved by a servo-piston whose movement
is electronically displacement controlled. Thus the piston in the pressure supply applies
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

bar

0.010

0.005

0
0 0.1 0.2 0.3 mm 0.4
~L

Figure 9.69 Typical load-shortening curve for a medium length cylindrical shell, of Rlt =400 and
Ll R = 2.25 under external pressure (from [9.115])

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
700 Shell Buckling Experiments

Figure 9.70 Test setup for buckling of steel cylindrical •hell\ under hy-
drostatic pressure at University of Essen (from [9.238). cour-
teS) of Profc"or H. Schmidt): (a) test chamber. (b) lo.'lding
(0 1 (b) S)Stcm (servo-pi~ton and pressure supply)

quasi-deformation controlled loading and unloading of the test shell. The change in the
quantity of pressurized water supplied is the control parameter. but it is also directly related
to volu me inside the test speci men. the contro l permits d ifferent loading rates.
The test shell was au ached to the end rings either as clamped (see Figure 9. 72a) or with
simple supports (Figure 9.72b). For clampi ng, the shell wa~ set into a conical groove and
cast with a hardening resin. An additional external rubber seal ensured watenightncs~. This
clamping closely approached fully clamped boundary conditions C4.
For simple suppons (Figure 9.72h). the specimen wa;, placed in a cylindrical groove. with
a tight-fitting steel ring on the inside of the shell preventing any inward displacement. It
should be noted that here the spec ified boundary condition II' = 0 inward is ensured. whereas
the outward w displacement is not completely prevented. as the external rubber sea l ensures
only watc11ightness and not zero d isplacement. Since. however, fo r external pressure loadi ng
the o utward w displacement is o nly of secondary im r o rtance, no external steel ring was
introduced. in order to permit as much rotational freedom as possible. But one should rc·
member that such end condition~ arc slightly softer than w = 0. Warping. on the other hand,
is free only away from the covcrplate, and restrained towards it. But the significant warping
displ;~ccment is that away from the coverplate. which i~ free. Hence one can conclude that
in the balance the end conditions of Figure 9.72b simulate simple suppons fairly well.
Schmidt and Stracke [9.2381 and 19.239] carried out a series of buckling and postbuckling
tc~ts on shon steel cylindrical shells in the Essen test setup. Their aim was tO investigate
the behavior of metal shells whose buckl ing is assoc iated with plasticity, not due to their
relat ive thickness as usual. but d ue 10 thei r sho rt ness. They tested 30 cylindrical shells, made
from longitudinally welded cold-rolled steel plate. with (R It) = 112- 204 and (L/ R) - 0.52-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Fi~ture 9.71 Essen test setup for buckling of cylindrical shell s under hydrostatic prc,;un: ·>ehcnuuic
(from [9.238]. counc<y of Professor H. Schmidt)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Cylindrical Shells under External Pressure, Bending or Torsion 701

(a J (b)

Figure 9.72 Essen test setup for cylindrical shells under external pressure-details of end rings (from
[9.238], courtesy of Professor H. Schmidt): (a) clamping, (b) simple supports

1.25, as well as 4 upper bound perfect reference specimens of similar dimensions but ma-
chined from thick steel tubes. The specimens had a radius of 200 mm and were mounted as
either "clamped" or "simply supported," in the sense of Figure 9.72 discussed above. Except
for 6 pilot test specimens, the initial geometrical imperfections of all the shells were mea-
sured and classified in three groups according to their amplitudes. Though the imperfections
were evaluated as plots of radial deviations from best fit circles (see [9.238]), their shape
was not decomposed into Fourier components and analyzed in the manner discussed in
Chapter 10.
The specimens were pressurized in the test chamber under quasi-deformation- controlled
loading. The prebuckling state was checked by strain measurements, and the growth of the
buckling and postbuckling pattern with loading was monitored with an internally rotating
displacement transducer (see Figure 9.71). A typical pressure-volume plot for one of the
shells, made from longitudinally welded cold-rolled steel plate with Rlt = 200 and LIR =
0.52, is shown in Figure 9.73. Note that the specimen is mounted with "clamped" ends,
designated in [9.239] "warping-restrained," as opposed to "warping-free" for "simple sup-
ports," to emphasize the dominant effect of axial restraint in the case of external pressure
buckling, discussed earlier in this section. As seen in the figure, prebuckling deformation
was mainly elastic, till the first buckle snapped in at the initial buckling pressure p 1 with a
slight drop in pressure. Upon unloading, the buckle (or the two neighboring initial buckles
that appeared instead in a few of the specimens) deformed plastically. It may be pointed out
that though the postbuckling stresses were definitely elastic, the nominal circumferential
stresses being a, = 0.2a, - 0.6a,, the buckle deformations were large enough to involve
plasticity.
When reloaded, the second buckle snapped in at pressure p 2 , 11 percent higher than p 1•
Similarly, after unloading and reloading, the third buckle snapped in at p 3 , 21 percent higher
than p 1 • This could be repeated several times, with increased local. yielding taking place.
The highest pressure sustained by this shell Pmax was 22 percent above the initial buckling
pressure. It occurred when the fourth, fifth and sixth buckle appeared with a subsequent
significant drop in pressure. If one disregards the unloadings and reloadings, the overall
elasto-plastic postbuckling behavior resembles the purely elastic ones of Figures 9.65 and
9.69, except that the ratio (PmaJp 1) is smaller here. A similar behavior with Pmax > p 1 was
observed on most of the clamped shells, whereas for the simply supported specimens usually
Pmax ~ PI•
The observed overall stable, or at least neutral, postbuckling behavior was found to be
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

less pronounced with increase in Ll R. The initial buckling pressure p 1 was found to be
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
702 Shell Buckling Experiments

0.8,------------

BUCKLE
0.7

'"H
E
~ 0.4
~
0. UNLOAOI~!Q_

0.3 RELOADII!Q.

0.2

0.1

0.5 1.5 6V ( %) 2.0 2.5

Figure 9.73 Pressure-volume curve of a short steel cylinder, with (Rit) 200, (LIR) 0.52 and
warping-restrained ends, under external pressure (from [9.239])

imperfection sensitive in relation to imperfection amplitude, whereas the maximum sustained


pressure Pmax was not significantly affected by imperfections. Comparison of stress-relieved
specimens with non-stress-relieved ones indicated that the longitudinal weld did not introduce
significant residual stresses in shells, but the geometric distortions due to the welds had a
detrimental effect on p 1, with the first buckle usually initiating at the weld.
Since most of the modern test rigs for cylindrical shells under external pressure have been
designed for stiffened shells, the discussion of additional test setups is deferred to Chapter
13.

9. 7.2 Bending
The earliest buckling experiments on shells subjected to bending moments were carried out
by Fairbairn and Hodgkinson around 1850 as part of their shell test program discussed in
Section 9.1. Extensive experimental studies, however, were initiated only in the early thirties
by Lundquist at NACA Langley on duralumin shells [9.240] and by Donnell at Caltech on
steel and brass shells [9.54]. In the thin-walled metal tubes with clamped edges tested in
these studies, failure always occurred as a result of buckling in small diamond-shaped waves
similar to those observed under axial compression.
For long cylindrical shells another collapse mechanism was identified by L.G. Brazier in
1927 [2.70] and further amplified by Chwalla in 1933 [9.241]. The phenomenon was dis-
cussed in Chapter 2, Subsection 2.1.13. A flattening of the cross-sections takes place during
bending due to the curvature produced by the bending. This ovalization increases with cur-
vature, and because the bending moment required for a certain curvature is smaller for an
oval tube than for a round one having the same circumference, the bending rigidity of the
shell progressively diminishes. This nonlinear effect leads to a limit load-type instability,
whereas the snap buckling into diamond-shaped waves is a bifurcation-type instability. In
thin shells the Brazier type limit load instability is usually preceded by a bifurcation-type
instability, but in thicker shells, with (R It) < 50 approximately, the influence of the inter-
action between the induced ovalization and plasticity often results in the dominance of the
Brazier-type instability (see [9.242], [9.243] and [2.71 ]).
Returning to the bifurcation-type buckling of cylindrical shells under pure bending, there
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

was an interesting discussion in the sixties and seventies on the ratio of critical stress in
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Cylindrical Shells under External Pressure, Bending or Torsion 703

bending to that in axial compression. As a result of a calculation by Fli.igge in 1932 [2.48]


for a shell with a particular R It ratio and a particular assumed longitudinal buckle half-
wavelength-radius, which yielded a ratio of (anhcndinglancnrnp•·""'"") = 1.3, which was quoted
by Timoshenko [2.1, p. 483] without qualification of the particular assumptions; the 30
percent increase for bending was used as a general rule for three decades. In 1961 Seide
and Weingarten [2.49] showed that when a linear membrane prebuckling state is assumed,
the predicted bifurcation buckling axial stress due to pure bending for finite length simply
supported cylindrical shells was for all practical purposes equal to that for uniform axial
compression. The theoretical critical stress ratio was therefore 1.0 instead of the 1.3 com-
monly accepted till then. Axelrad [9.244] and W.B. Stephens eta!. [2.52]later showed that
when nonlinear prebuckling effects and edge constraints were taken into account the same
result was obtained for short shells. For longer shells they found an interaction between the
bifurcation buckling mode and the Brazier effect, leading to lower buckling stresses, with
the Brazier ovalization dominating for very long shells.
Experimental results, however, showed significantly higher critical stresses for bending
than for axial compression. Already in the early tests of Lundquist [9.240] and Donnell
[9 .54], which showed a similar decrease in buckling stresses with increase in (R It) for
bending and axial compression as well as a similar scatter, the experimental values for
bending were respectively 30-80 percent and 40 percent higher than those for axial com-
pression. These tests and others [9.245]-[9.247] were summarized and analyzed statistically
by Suers et al. at North American in 1958 [9.185]. The buckling stress coefficients are
presented in Figure 9.74 and the 90 percent probability curves (not shown) indicated 20-60
percent higher buckling coefficients (decreasing with increase in Rlt) for bending compared
to those for axial compression.
More recently, Esslinger and Geier [9.248] carried out comparisons between experimental
bending buckling stresses and those obtained under axial compression on similar Mylar
cylindrical shells of Rl t ~ 430 and Ll R = 2.33, with both series tested by the same crew
on the same system at DLR Braunschweig. They found that the experimental bending buck-
ling stresses exceeded the compression ones by about 20 percent. Their tests were performed

~,-------------------------------~

THEORETICAL COMPRESSIVE
BUCKLING COEFFICIENT

0
6 F----------+----------------
• I o DONNELL (REln~-:54)
0
! • IMPERIAL (REF. 9.246)
! : ~ o LUNDQUIST (REF. 9.240)
51- I • " e~~~ 9. 245) a (REF.
· 1 • r:,,; oo. r• NAA(REF. 9. 185)

~
~_E~l'_ERS_ON_~F. 9. 247)
0
0 0
~ §

4 "
: .. 0o~i •
~ ...... C/00 f : 0
0
~ .3 • ~lloo t:. ~ 0
, .. B o
~ * 0 0 • " 0

0
.2 0
0 me
0 I
0
0 I•
0
,I

500 1000 1500 2000 r 2500 3000


I
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 9.74 Experimental bending


Copyright Wiley buckling stress coefficients as a function of (Rit) (from [9.185])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
704 Shell Buckling Experiments

in a loadi ng dev ice, which was a bending extension to the DLR RZ 100 system discussed
in Section 9.4 and shown in Figures 9.38 and 9.39. The bending moment was produced in
their test setup by two additional hydraulic cyl inders, acting via hi nges upon two levers
attached through pivot bearings to the compression plates of the axial loading system (see
Figure 17 of [9.248)).
The reason for the higher experimental buckling stresses in bending is primarily that this
case is less imperfection sensitive because in bending. buckling is initiated in the narrow
region of the greatest compressive stress and only imperfections in this region can affect it,
whereas under axial compression imperfections anywhere on the shell surface can trigger
buckling.
As pointed out in Section 9.4, extensive tests were carried out in the aeronautical industries
in the forties, fifties and sixties on the buckling of cylindrica l shells under different loadings.
The North American Aviation experimental studies [9.185), discussed in Section 9.4 as a
typical example of these test programs. also provided empirical data for bending. Figure
9.26b shows schematically the application of bendi ng loads in the test jig, as well as torsion
loads and thei r combinations with axial compression. The bending was applied by a hydraulic
bending load strut via cables and pulleys and a load cell to the lower head. which then
exerted the bending moment on the shell. Torsion was appl ied in a similar manner, and
desired load combinations were obtai ned by appropriate loading of the respective load struts.
The apparatus was simple, but not very rigid.
For bending experiments on larger and stiffer shells a more rigid test rig was necessary.
In the fifties and sixties test setups for large and relatively stiff cylindrical shells were de-
veloped at NASA Langley Research Center that are still in use today [9.247]. [9.249)-
[9.252). The earlier experiments were on ring-stiffened circular cylinders with heavy rings
[9.247] and investigated only panel buckling between rings, whereas the later ones also
studied general instabil ity, which is discussed in Chapter 13. 1\vo of the test setups for study
of panel instability are shown in Figure 9.75. Both utilize the rigid reinforced fl oor and rigid
support wall of the NASA Langley structural testing laboratory. The ring-stiffened 7075-T6
aluminum alloy cylinders are bounded by heavy stee l end rings. One end ring is attached to
the support wal l. The other is fixed to a very stiff circu lar loading plate that applies the
bending moment to the shel l. In each test rig, a heavy loading frame transmits the bending
loads, applied by the hydraulic jack of a testing machine, to the loading plate. The weight
of the end fixtures and that part of the test rig that would otherwise be supported by the test
specimen was counterbalanced near their centers of gravity by weights from the cantilever
beam fixed above the specimen to eliminate stray loads in the test specimens. For the same
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 9.75 NASA Langley bending test rigs for stiffened cylin-
drical shells: (a) loading frame for failing moments
below 0.34 MN-m (= 3000 in.-kips). (b) loading
frame for larger and stiffer cylinders. up to R "' t.O
m ( = 40.4 in.). courtesy of NASA Langley Research
Center
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Cylindrical Shells under External Pressure, Bending or Torsion 705

purpose rollers were employed between the loading frame and its floor supports (at the pivot),
as well as between the loading frame and the top of the hydraulic jack. Use of rollers allowed
the cylinders to shorten during application of the load and ensured that the loads at the roller
locations were normal loads. The rollers and the surfaces against which they reacted were
case hardened.
The hydraulic jacks applied the desired loads within an accuracy of 1 percent. However,
when loading frames as shown in Figure 9.75 are employed, the load actually applied to the
test specimen may be less than that indicated at the jack because of friction in the bearings
of the loading frame. To correct for this difference, strains were measured at a few locatins
on each cylindrical shell and compared with those predicted from the indicated load. For the
smaller frame of Figure 9.75a, errors of 1 to 7 percent were found, the mean error being 4
percent. This mean 4 percent correction was then used since it was felt that the variations
in the error for individual specimens may have been partially due to local changes in the
measured strain distribution. For the larger frame of Figure 9.75b, much smaller errors were
found and therefore no corrections for friction were applied.
It should be pointed out here that the apparently very rigid end plates and fixtures usually
employed in test rigs like the NASA one still have some flexibility. The resulting strains and
displacements have been found by many investigators to be far from negligible, and hence
they should be measured and considered in careful experimentation.
The hydraulic jack pushed its attachment to the loading frame up, resulting in compressive
stresses in the top of the cylindrical shell tested. Extensive strain gage data were taken from
the region subjected to the maximum compressive stresses. More strain gage data were
recorded in the later tests, which also investigated the general instability of stiffened shells.
The strains from the gages were recorded during the tests at the Langley central digital data
acquisition facility. Usually, two tests were made on each cylinder. For the second test the
cylinder was rotated 180° in its fixture after the first test. Thus, the generator subjected to
maximum tensile stress in the first test was the generator of maximum compressive stress
in the second one.
Many metal cylindrical shells, ring stiffened corrugated and integrally stiffened, were
tested in these test setups, with a wide range of Rlt = 120-1850. More recently, also,
graphite-epoxy cylindrical shells were tested in the larger test rig of Figure 9.75b.
A good example of a modem smaller laboratory-type bending test facility is that built
recently by Kyriakides and his co-workers at the Engineering Mechanics Research Labora-
tory of the University of Texas at Austin [2.71], [9.243] or [9.253]. Though developed
primarily for experimental study of inelastic buckling, it is also suitable for elastic instability
of cylindrical shells and represents the state of the art. The following description of the test
setup is based on [2.71] and [9.253].
The facility consists of a four-point bending machine (Figure 9.76), transducers for re-
cording geometry changes and a computer-based data acquisition system. The machine is
able to apply bending, as well as reverse bending, because it was designed for a test program
of cyclic bending (see [9.253] and [16.112]). The pure bending is achieved by applying
concentrated loads in the form of a couple to the ends of the test specimen assembly. The
concentrated loads are applied through rolling contact, which allows free axial movement of
the test specimen during bending and therefore always ensures essentially zero axial force.
Each test specimen is fitted with close-fitting solid extension rods, as shown in Figure
9.77. The rods extend more than three diameters into the free ends of the shell and are
carefully chamfered at the ends to reduce the effect of the discontinuity. The rollers engage
these extension rods. Four rollers are used at each end to allow bending in both directions,
if required. The rollers are located on two double-strand, freely rotating, sprockets (pitch
diameter 9.56 in. [243 mm]), which are mounted with pillow blocks on two heavy parallel
support beams, as shown in Figure 9.77. The beams are 70 in. (1.78 m) long and the distance
between the sprockets can be varied. Chains run around the sprockets and are connected to
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

the hydraulic cylinders and load cells, forming a closed loop.


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
706 Shell Buckling Experiments

(a)

(b)

Figure 9.76 Texas University pure bending test faci(ity fo r cylindrical shells: (a) side view, (b) sche-
matic diagram (from [2.71))

Bending is obtained by contracting either the top or the bottom cylinder, causing rotation
of the sprockets . The hydraulic system was designed to allow the idle cylinder to extend an
equal amount, which ensures that the periphery of the closed loop (consisting of the chain,
the hydraulic cylinders and the load cells) has constant length. Reverse loading can be
real ized by reversing the direction of the flow in the hydraulic circuit. For monotonic bend-
ing, the bending curvature capacity of the testing machine can be doubled by removing one
of the hydraulic cylinders and extend ing the second one to its full stroke of I 0 in. (0.26 m).
T he applied bending moment in the test speci men is directly proportional to the tension
in the chain, which is monitored by two calibrated load cells. The bendi ng defonnation of
the test specimen is uni fonn along the length, and the bending curvature is proportional to
the sum of the rotation of the two sprockets. The rotation of each sprocket is monitored by
a li near variable differential transformer (LVDT) as follows: A flex ible, inextensional, thin
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

cable nms over the circular hub of the sprocket. The cable is spring loaded and is connected

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Cylindrical Shells under External Pressure, Bending or Torsion 707

to the core of the LVDT. Thus, rotation causes linear movement of the core and a proportional
signal from the LVDT.
In the recent study on bifurcation and localization instabilities of long aluminum tubes
[2.71], the majority of the test specimens tested had an effective length of 30 in. (0.76 m).
With the test facility configuration used, such specimens could be bent to a curvature of
0.070 in. -t (2. 75 m 1). Larger curvatures can be achieved by reducing the length of the
shell. The capacity of the bending device shown in Figure 9.76 is 15,000 lb-in. (1,700 N-
m). It was designed to be stiff relative to the test specimens. For the experiments of [2.71],
the energy stored in the device was less than 4.5 percent of the energy stored in the strongest
of the test specimens (approximately half of this stored in the hydraulic system). In other
uses of the bending device the relative stiffness is even larger. For example, in the tests of
[9.253] the energy stored in the device was less than 1 percent. It should be noted that a
stiff machine is desirable for all buckling experiments and is essential for testing beyond
limit loads.
This computer-based data acquisition system, used to monitor the variables measured
during the experiment, consists of several signal conditioning units, an analog-to-digital
(A/D) converter, a desktop computer and peripherals such as digital plotters and an electronic
display unit. The system enables real-time display of some of the variables and provides
storage of the data for later analysis.
During a typical monotonic bending experiment on a shell (as in [2.71]), the variables
monitored were the moment (M), the curvature (K) and the ovalization at one or two discrete
points along the length of the shell (!:.DJ The data acquisition system stores a set of
{M, K, t:.D;}, whenever any one of the variables changes by a prescribed small value. The
number of sets of data collected from each experiment varied from approximately 125 for
the thinner shells to 450 for the thicker shells. The uncertainties in the recorded data were
as follows: moment ~0.8 percent; curvature ~0.6 percent; ovalization ~0.1 percent.
Since the ovalization of the cross-section of the tube is known to play an important role
in the stability of the structure, several lightweight transducers for monitoring small changes
in the geometry of the shell during bending (primarily the diameter) were developed. Their
main element is a rigid rectangular frame riding on the shell, which supports a sliding block.
The instruments are lightly spring loaded and contact the shell through knife edges in the
static ovalization transducers (Figure 6 of [2.71 ]), or rollers in the scanning ovalization
transducer (see Figure 9.78). The relative displacement between the two knife edges or
rolling contacts is monitored by a miniature LVDT. In some experiments two static ovali-
zation transducers were placed approximately I 0 shell diameters apart in order to establish
the onset of localized deformation in the specimen.
In at least one experiment for each shell R It ratio considered, a scanning ovalization
transducer (shown in Figure 9.78) was used to periodically scan the ovalization along the
length of the shell, contact with the shell being made through rollers. The transducer is
moved along the length of the shell manually. Its axial position is established by an encoder
system that works as follows: a thin flexible tape, with regularly spaced black-and-white
markings, is bonded on the tension side of the shell as shown in the figure. An emitter/
receiver photodiode mounted on the transducer produces an electrical pulse whenever it
encounters a white marking. The pulse, suitably processed, triggers the AID converter and
the value of ovalization is recorded in the computer. In the majority of the experiments, the
encoder had a resolution of five readings per inch.
In the case of thinner shells, short-wave-length ripples developed on the compression side
of the bent shell (see Figure 9.79). For the aluminum shells used in the experiments of
[2.71], with typical diameters between 1.0 and 1.5 in. (25-38 mm), the amplitude of these
ripples was of the order of 0.001-0.005 in. (0.025-0.125 mm). Thus, a more sensitive in-
strument with higher axial resolution than that shown in Figure 9.78 was developed for
measuring the geometry of these ripples. It consists of an axial guide with a fine lead screw,
supporting a vertical traversing system. Details are given in [2. 71].
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
708 Shell Buckling Experiments

(a)

VERTICAL PROXIMITY
TRAVERSER TRANSDUCER

PHDTOOIODE

ARrGGER

8
SIGNAL

DISPLAY PLOTTER

(b)

Figure 9.78 Transducer for scanning ovalization along shell leng1h: (a) lransducer mourued on lube,
(b) Lransduccr schcmalic and AID convener lriggering circuil (from [2.71))

Figure 9.79 Local buckling due LO bending: (a) axial ripple developed
on 1be compression side o f a shell wilh (Ri t) = 16. 1 in
pure bending, (b) local buckle in a shell wi1h (R/1) = 22
which occurred shor1ly afler Lhc ripples developed (from
(b ) 12.71))

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Cylindrical Shells under External Pressure, Bending or Torsion 709

A typical example of the experimental results obtained by Kyriakides and Ju [2.71] in this
test facility is presented here in Figures 9.80 and 9.81 to indicate their character, though
they relate primarily to inelastic buckling. The shell was a long 606 I-T6 aluminum tube of
1 in. diameter, with R It = 25 and L/ R = 60. In this test series, at least two experiments
were conducted for each Rlt ratio studied (altogether 11 Rlt ratios). The first experiment
involved monotonic bending at a strain of about 1 percent/min. to failure. In the second
experiment the bending was periodically interrupted for a complete ovalization scan till
failure. Hence, the example discussed includes two tests.
The results are presented in nondimensional form using the following normalizing pa-
rameters:
(9.14)
where D0 is the shell mean diameter and u, the yield stress. Figure 9.80 shows the bending
response of the shell: (a) the moment-curvature curve and (b) the ovalization-curvature re-
sponse. The predicted response was calculated with the analysis of [9.254]. In the figure,
the curvature at which ripples were first detected on the compression side is identified by
the symbol (l). This ripple curvature is usually obtained from the second bending experiment,
in which bending was interrupted while the shell was checked for ripples. The symbol (l)
denotes the curvature at which catastrophic collapse occurred, and the symbol (A) designates
the curvature at which the maximum (limit) moment was or would be achieved. (In Figure
9.80 this appears only on the theoretical curve since local collapse precedes the limit mo-
ment.) Figure 9.81 shows the ovalization along the shell at different curvatures, obtained by
the scanning ovalization transducer of Figure 9.78. Just prior to catastrophic collapse, a
pocket of short wavelength ripples, with A 0.28R, (of the type shown in Figure 9.79) was =
detected, and a detail scan with the special transducer of higher axial resolution is shown in
the inset of the figure. Such ripples usually lead to a local collapse of the type shown in
Figure 9.79b.

1.2
AI-6061-TS


.
1.0 25
------~-

::E
' .8
::E
A
I.
.6

.4 --EXPERIMENT
UNIFORM TUBE
----PREDICTION
.2 I USING Ref. 9.254)

lal

0 .2 .4 .6 .8 1.0 1.2 1.4 1.6


-KIKI

06
/
/
/
/
/
.04 /
~
/
0 v
<I
.02
I b)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

0 .2 .4 .6 .B 1.0 1.2 1.4 1.6


-KtKI

Figure 9.80 Bending response of aluminum shells with (Rit) = 25: (a) moment-curvature,
(b) ovalization-curvature (from [2. 71])

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
710 Shell Buckling Experiments

1.2 (in)

.08

o 0
---.32

c .04 ~L~-......,.~~-·-====--=----~ 22
<I

i 0 --+--\ X/l e-

Figure 9.81 Ovalization along shell length at different curvatures, for a tube with (R/t) = 25. The
inset shows detail of short wavelength ripple (from [2.71])

A second test setup of similar design was developed at the University of Texas in the

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
mid-eighties for combined bending and pressure experiments (see [9.255] and [16.112]).
This improved bending test facility can be operated remotely (with the aid of a closed-loop
servo-hydraulic control system, described in [ 16.112]) inside a special cylindrical pressure
vessel with a working pressure of 5000 psi (350 bar). This second test device, which also
included modification of transducers, signal cables and hydraulic lines for operation in a
wet, high-pressure environment, was developed for a series of combined loading tests related
to pipelines in deep offshore waters, which were canied out by Kyriakides and his co-workers
in the late eighties.
The main interest in bending instabilities in recent years has been related to inelastic: tubes.
Further discussion is therefore relegated to Chapter 16.

9. 7.3 Torsion
The earliest significant experiments on the buckling of thin cylindrical shells under torsion
were the 1932 tests by Lundquist at the NASA (then NACA) Langley Research Center
[9.256] and the 1933 tests by Donnell at the California Institute of Technology [2.35], as
well as the 1931 tests by Sezawa and Kubo on nine rubber models at Tokyo University
[9.257]. Lundquist carried out extensive tests on 15 in. and 30 in. diameter duralumin tubes
of (Rit) = 329-1424 and (L/R) = 0.2-5.0. Donnell, on the other hand, employed smaller-
sized specimens, of 5 I 16 in. to 6 in. diameter, made of steel and brass shim-stock 0.002 in.
to 0.006 in. thick, with (Rit) = 83-1370 and (L/R) = 0.1-335.
In his 1933 NACA Report [2.35], well known as the pioneering document of the Donnell
shell theory, which, as pointed out in Chapter 2, facilitated simplified shell analysis for half
a century, Donnell also showed his ingenuity as an experimentalist. First he made a case for
small-scale models "because of the great ease and cheapness of construction and testing"
and, comparing their results with those of the significantly larger shells used by Lundquist,
concluded that "there is no great disadvantage or danger in using such small specimens,"
since using the very thin sheet meant that "the proportions were such that the stresses were
always well below the elastic limit." He then described the fabrication of these specimens:
carefully rolling the sheets around rods of proper diameter, soldering the longitudinal seams
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Cylindrical Shells under External Pressure, Bending or Torsion 711

and then soldering the tubes to heavy end pieces, with jigs holding the sheets in a true
cylindrical form and preventing local waving during soldering operations. Except for some
initial waves in the specimens with the largest (RI t) ratios, "no departure from true cylin-
drical form could be detected by eye or fingers." Modern, more sophisticated methods of
measurements would have no doubt indicated some initial imperfections.
The influence of the longitudinal seams was discussed and shown to be mostly negligible
(viewed from the authors' experience, though not mentioned by Donnell, this was probably
because the stiflening effect of the seams cancelled the weakening effect of the minor wav-
iness or small residual stresses near the seams). The edge conditions and their careful fab-
rication were also precisely detailed for short- and medium-length shells.
Donnell's small torsion-bending-compression testing machine was the forerunner of the
combined loading test setups of the fifties and sixties. A diagrammatic top view of the

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
machine is shown in Figure 9.82a. The specimen is attached to two L-shaped members ABC
and DEF, which are balanced on practically frictionless universal joints B and E. The ends
of the specimen are therefore free to rotate in any desired direction. Axial loads are applied
through these universal joints, which ensures the desired line of action of the load. For
bending, downward forces are applied at D and C through wires that extend down to a
crossbar D 1C 1 under the specimen. As seen in the schematic side view, Figure 9.82c, a crank
is used to press down on a fulcrum mounted on D 1 C 1• The crank and fulcrum are movable
along D 1C 1, permitting variation of the ratio of the forces at D and C and therefore of the
bending moments at the two ends of the specimen. Torsion is applied to the shell by pulling
down on F through a wire by means of a crank. As a reaction, joint A is prevented from
vertical motion by vertical wires, which, however, permit some horizontal motion (see Figure
9.82b). Axial load is applied by moving point B towards E with a crank H. Joint E is mounted
on one of the cantilever springs J, which measures the axial load. The arms BC and AB are
themselves cantilever springs, whose deflection measures the bending and torsion moments
respectively. The dial gages that measure these deflections are mounted on unstressed arms
of the frame (not shown). The report also discusses details of the construction of the test rig
and of the measuring instruments.

A A

F F
(a) TOP VIEW (c) SIDE VIEW

~HC
D E
----...---1
1 ----1 CP

r
I • l
0' C'
(b) FRONT VIEW

Figure 9.82 Donnell's 1933 torsion-bending-compression testing machine-schematic


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
712 Shell Buckling Experiments

Donnell's report is worth reading even roday, after 60 years, for its logic and engineering
sense.
In Figure 9.83, the test data of Donnell [2.35], and Lundquist [9.256], as well as some
results from the 1934 paper of F.G. Bridget et a!. [4.31] and from Moore and Westcoat's
torsion tests on stiffened circular cylinders [9.258] are compared with linear theory predic-
tions by Batdorf et a!. for classical simple supports [9.259]. As can be seen, there is indeed
not much difference among the different test data. Some more recent results, from tests by
Yamaki on six polyester film (Diafoil) cylindrical shells (with Rlt = 405) in the early
seventies [9.111], have been added in Figure 9.83, and fit the earlier data quite well.
One should remember that for buckling of cylindrical shells under torsion, the rotational
boundary conditions are of prime importance only for short and thick shells, for which the
Batdorf shell geometry parameter Z = Vl - v 2 (F I Rt) < 50 (see for example Figure 2.3
of [2.47]). However, as in the case of external pressure, the influence of axial restraint (u =
0 instead of Nx = 0), though smaller, is noticeable for all geometries (the predicted difference
being about 9 percent).
In contrast to the cases of axial compression or external pressure, only very few experi-
mental investigations on the torsional instability of circular cylindrical shells have been
reported in the literature, after the extensive tests of the thirties discussed above, and some
parallel studies on thicker tubes (for example the 1937 NACA experiments, [9.261], which
showed that elastic instability can dominate already when Rlt > 25). It appeared as if by
the end of the fifties the reasonable correlation between linear theory and the available
experimental data shown in Figure 9.83 (similar fair agreement was obtained for wave num-
bers) was accepted as satisfactory, with an average reduction factor of about 16 percent for

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
the buckling stress (see for example [9.260]). Some large deflection analyses then motivated
additional experiments that focused on the postbuckling behavior. In the late fifties, Nash at
the University of Florida [9.262] tested 26 thin-walled stainless steel cylindrical shells of
(Rit) = 232-685 and (L/ R) = 0.7-9.5 in order to corroborate the validity of his nonlinear
large deflection analysis. The cylinders were tested in a special torsion machine designed to
accommodate the shells of the widely spread range of lengths. Initial imperfections (called

.. _ lif
-'I- THEORY (BATDORF, STEIN, AND SC:HILDCROUT ·REF 9.259 )~ -

MATERIAL

~O-DONNELL REF. 2.35 STEEL


t. MOORE AND::-:WE:=SCO=::AT:-:cRE:=f~9.2=5=Bt-2:C:0:-:'24:::'_'::T::-3i
REf 4.31 , STEE L

• YAMAKI REF.9.111 : MYLAR

1(~--~10"3----~-----::!105
Z =~I- l/ 2 ( i!' I R!)

Figure 9.83 Circular cylindrical shells under corsion-comparison of test data with approximate linear
theory for classical simple supports (from [9.260] and [9.111])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Cylindrical Shells under External Pressure, Bending or Torsion 713

by Nash "circularity prior to loading") were measured at 10-12 stations along the length
of the shell and at 24 points around its circumference, with a dial gage mounted on a radial
arm rotating with an axially aligned shaft. To overcome the errors that result from variations
in the centroid of the shell along its length, the initial radius of curvature was also measured
separately at each of the 24 points along the circumference by a special instrument, which
was essentially an adaptation of a spherometer. Strain gage rosettes were then mounted at
points of maximum initial departure from perfect circularity, as well as at a few points
exhibiting almost perfect circularity. Note the attention paid to initial imperfections at a time
when their importance had not yet been fully realized.
Buckling occurred rather suddenly into readily visible waves. but not as violently as is
common for axial compression. The long shells buckled into 1 or 2 waves, the intermediate
ones into 4-7 waves occurring simultaneously, and the short ones into 6-11 waves, again
occurring simultaneously. These helical waves covered only part of the circumference, but
when extrapolated their number correlated well with the wave number predicted by linear
theory. In 21 of the 26 specimens the first buckling wave was found to center around the
point of maximum measured initial imperfection, or if several waves appeared simultane-
ously, one of these waves embraced this point. In the remaining 5 shells the waves centered
around the point of second largest measured imperfection. As discussed in Chapter 10, such
behavior can now be explained by the relation of the shape and wavelength of the imper-
fection to the inherent wave pattern of the corresponding perfect shell. For the long and
intermediate-length shells the buckling waves were found to extend over 113 to 2/3 of the
length of the shell, and not over its entire length as assumed in analyses. A similar obser-
vation was earlier made in tests of the long but thicker tubes of [9.261], that failed by elasto-
plastic two-lobe buckling, with the waves extending only over the central portion of the tube.
The buckling stresses measured were 40-95 percent of the linear theory predictions, with
the large knock-downs corresponding to the larger imperfections. The agreement with non-
linear large deflection theory was better, but the correlation was only partial, by bracketing
of the experimental points with computations for certain imperfection amplitudes.
In the early seventies, as numerical nonlinear postbuckling analyses were flourishing, ad-
ditional careful postbuckling experiments were also carried out on cylindrical shells subjected
to torsion. As pointed out earlier in Section 9.4, the extensive theoretical and experimental
studies of Yamaki and his co-workers at Tohoku University, included an experimental in-
vestigation of the buckling and postbuckling behavior under torsion [9.111].
A schematic diagram of the test setup is shown in Figure 9.84. It is essentially the same
as that used in Yamaki's earlier postbuckling experiments, except for the different loading
system, designed to apply pure torsional loads. Again the system is fixed in the rigid frame
of a large tension test machine, with (1) the table of that machine and (9) its movable cross-
head. The test cylinder (3) rests on a base plate (2), and a stiflening steel plate (4) is attached
to the cover plate. Four equidistant cylindrical rods (5), which ride without axial constraint
in ball bearings (17), are fixed to the stifTening plate (4) and transmit the torque to the test
cylinder. The weight acting on the test shell is counterbalanced by two sets of weights (6),
hung on thin wires (7) through pulleys (8).
The loading system (9)-(23) starts at the cross-head (9) with a reduction gear (I 0), through
which the movement of the lever (II) is translated into rotation of the central shaft (12). A
universal joint (13) eliminates the effects of eccentricity of loading and a spline shaft (14)
prevents axial constraint. A torquemeter ( 15) is attached to the spline shaft. The loading
head (16) consists of a rigid shaft, whose lower end (16A) is a circular disc with 4 holes,
in which the four cylindrical rods (5) that transmit the torque to the test shell ride on ball
bearings (17). The loading head is supported axially, via a ball bearing (18), by a casing
(19) that is attached to the four-legged supporting stand (20), which anchors it to the cross-
head. A cylindrical cap (21) is screwed onto the top of the casing (19) and supports the
loading head through two thrust bearings (21A). Its rotation with the lever (22) adjusts the
vertical position of the loading head relative to the test cylinder. This rather complicated
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
714 Shell Buckling Experiments

(a)

Figure 9.84 Schematic diagram of Yamaki's test setup for torsion (from [9.110] and [9.263], courtesy
of Professor N. Yamaki): (a) test setup, (b) measurement of angle of twist, (c) measurement
of axial shortening

loading system ensures that almost pure torsion is applied to the shell without any significant
effects of axial forces or bending moments. Combined loading of torsion and compression
is also possible with this loading head, and then axial compression is applied to the stiffening
plate (4) via a load cell (23). External or internal pressure was also later applied in combi-
nation with torsion in this test setup (see 19.263]), after it was connected to the pressure and
vacuum system mentioned earlier (see Figure 9.64).
The deflection measuring apparatus (24)-(29) is the same as that discussed earlier in
Section 9.4 for axial compression [elements (24)-(29) here are (15)-(20) in Figure 9.31].
which permits precise measurements of both axial and circumferential distribution of the
deflection without affecting the buckled surface. The axial shortening is also measured in
the axial compression experiments [the strain gages (30) and the calibration dial gages (31)
in Figure 9.84c appear as (21) and (22) there in Figure 9.31]. Two sets are again used in
series to obtain the average value of edge shortening at diametrically opposite points of the
stiffening plate (4). The same sets are employed for measurement of the angle of twist by
arranging them horizontally as in Figure 9.84b.
As already mentioned, the six specimens were made of polyester film (Diafoil), with
(Rit) = 405 and (L/R) = 0.2-1.6. The relation between torque T and the angle of twist 'l'
for one of the cylindrical shells, with (L/R) = 1.14 and therefore Z = 500, is shown in
Figure 9.85. The solid curve 0-A-B-C-D-E-0 represents the actual results obtained with an
X-Y recorder. Up to the critical point A, the torque-twist relation is linear. Then the shell
snaps to the stable postbuckling state with 15 circumferential waves. The path B-C-D rep-
resents the stable postbuckling state, and at D the shell snaps back to the stable prebuckling
state at E. The dynamic snap-through occurs due to the insufficient torsional rigidity of the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

loading system, which from the slope of curves AB and ED was estimated to be about 1.9
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Combined Loading 715

kNmlrad. the hysteresis loop in the range BC indicates a small frictional force against axial
movement of the shell after buckling. From experiments on another shell of the series, with
stepwise changes of axial load, this frictional force was estimated at about 1 percent of the
critical axial compressive load. Hence the curve for the ideal case of zero friction and no
snap-through can be estimated. It is shown by a chain line in Figure 9.85. It can be seen
that the postbuckling equilibrium torque decreases monotonically with twist. The experi-
mental critical torque was 3.26 N (which is 92 percent of the theoretical one).
From the results for the other shells in the series (see the discussion in [9.111] and [2.47])
it is observed that the slope of postbuckling torque versus twist strongly depends on the shell
geometry. For small Z the postbuckling torque increases monotonically with twist, changing
direction as Z > 50, and decreases more distinctively with twist as Z increases (see Figure
9 of [9.110] or Figure 3.14 of [2.47]). Earlier experiments by Weingarten [9.264] yielded
lower postbuckling curves than those for Yamaki but showed an approximately similar trend
in the influence of shell geometry.
Figure 9.86 depicts contour lines for typical postbuckling configurations under torsion
measured on a shell with Z = 500 (from [9.111]) for two sets of torque T and angle of twist
'¥. One may note that the waveform is symmetric with respect to the center of each buckle
and that as the angle of twist increases, from (a) to (b) the inward buckles (the solid contour
lines in the figure) become dominant and cover more of the surface at the expense of the
outward buckles (the dotted contour lines in the figure). The comparison between theory
[9.265] and [2.47]) and experiment showed good correlation also for postbuckling behavior
under torsion.
The buckling behavior of cylindrical shells subjected to torsion in combination with other
loads was a theme that drew the attention of many investigators (see for example [9.186]
and [9.266]). In particular, as mentioned earlier in relation to axial compression, the stabi-
lizing influence of internal pressure was extensively studied in the late fifties and sixties by
Harris et al. [9.184] and Weingarten [9.264] and more recently by Kodama et al. [9.263]
and Foster [9.267]. Some of these studies are discussed in the next section, together with
those on other load combinations.

9.8 Combined Loading


In practice, shells are often subjected to a combination of loadings. Since the interactions of
the buckling modes due to different types of load are usually nonlinear and may depend on
the boundary conditions and on the imperfections, buckling under combined loading has
been the focus of many recent experimental studies. The combination of loadings may be
destabilizing, as for example in the case of axial compression and external pressure, or
stabilizing, as in the case of the addition of internal pressure to any of the fundamental loads.

9.8. 1 Buckling of Pressure Stabilized Shells


As mentioned earlier, the use of pressure stabilized cylindrical shells in missile structures in
the fifties initiated many studies of the buckling behavior of pressurized shells, as for example
the extensive tests of Harris and Suer et al. [9.168], [9.184] and [9.185], Dow and Peterson
[9.252], or Weingarten et al. [9.103], [9.104] and [9.264]. The main purpose of the pressure
stabilization of the shells was to stabilize the initial postbuckling behavior and thus reduce
their imperfection sensitivity, allowing them to carry loads of nearly perfect shells. For
example, Figure 9.87 (from [9.104]), which presents the variation of the critical axial stress
coefficient 0:"' = aj(Et/R) with the internal pressure parameters p = PcJE(R!tf for two
series of Mylar and 7075-T6 aluminum alloy cylinders with (Rit) = 685-800, clearly shows
the trend of approach to the perfect shell values with increase in internal pressure. This was
especially important for the cases of axial compression and bending, but also significant for
--`,`,`````,`,````,,``,

torsion.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
716 Shell Buckling Experiments

I A

30~
-~I
t:'
~ c
1-
20
--EXPERIMENT
---ESTIMATE FOR
ZERO FRICTION
AND NO SNAP- THROUGH
z" 500
10 N • 15

0 2 4 6 8
>jltrod.I0- 3 )

Figure 9.85 Relation between torque T and angle of twist '{r for a typical Diafoil cylindrical shell no.
500-4, with Z = 500 (from [9.110], courtesy of Professor N. Yamaki)

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

n~o·==1i2D::::'=1=3o:':·==1,=o·==1si:o=.~1:::;:6o=.:::: -1 o 1 Zmm
o--- w-

110' 120' 130' 11,0' 150' 160'


b o-- w-

Figure 9.86 Contour lines for typical postbuckling configurations in torsion (in mm) of a typical
cylindrical shell no. 500-4, with Z = 500 (from [9.110], courtesy of Professor N. Yamaki):
(a) T = 2.173 Nm, 'It = 3.25 X l0- 1 rad, N = 15, (b) T = 25.1 Nm, 'It= 5.43 X 10- 1
rad, N = 15
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Combined Loading 717

0.7r-------------------,

0.6

MYLAR - - - -

I
~ 0.5
0: R/t o 800
-'
L) ~ 0.4
L/Ro20
R/t LIR
0 685 D. 987
0 688 0.987
0 717 0987
02 o~--~OA~-~OB~--~1.2~-~16~--72D

p
p~ E(R/t)2

Figure 9.87 Variation of critical axial stress coefficient 0:" with internal pressure parameter p, for Mylar
and 7075-T6 aluminum alloy cylindrical shells with (Rit) = 685-800 and (L/R) = 0.99-
2.0, showing the trend of approaching the perfect shell values with increasing internal
pressure (from [9.104] and [9.252])

As pointed out earlier in Section 9.4, the etiect of internal pressure was also manifested
by a pronounced change in the buckling pattern (see Figure 9.28). As a matter of fact, in
general the changes in buckling modes or their interactions offer a physical explanation to
the buckling strength of shells under combined loading.
The destabilizing combinations of loads, however, are the more important ones in practice
and have therefore been studied more extensively (see for example [1.17], [6.13], [9.63] and
[9.268]). In civil and offshore engineering combinations of axial tension with external pres-
sure and/ or bending or torsion appear often, where the tension stabilizes in elastic buckling,
but destabilizes in plastic buckling as will later be discussed in Chapter 16.

9.8.2 Combined Loading Test Setups

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
The essence of a combined loading test setup is to ensure the independence of the different
loading elements, or in other words to ensure that the interaction between the effects of the
loads is entirely the result of the shell behavior and not partially due to some test rig coupling
of the applied loads. For example, a thrust bearing is usually inserted between the load plates
of the testing machine or test rig and the top end plate of the cylindrical shell to allow
rotation of the top end of the cylinder under torsional loading independently of the axial
load that may be applied concurrently.
The capability for combined loading was pointed out in the descriptions of the different
buckling test setups. For example, in Donnell's 1933 test rig (Figure 9.82), which was one
of the earliest combined loading setups, a number of practically frictionless universal joints
ensured the independence of the effects of torsional loading from those of axial compression
and bending, as pointed out earlier. The North American Aviation test rig of the late fifties
for combined axial compression, bending and torsion loads (Figure 9.26b) was typical of
careful industry experiments at the time. Bending moments and torques were applied to the
bottom head by two cables, which were activated by hydraulic jacks. Heavy-duty ball bearing
aircraft pulleys and 1 I 4 in. steel cables were used to minimize the friction in the pulley
system, which was indeed found to be negligible. A special bearing was probably fitted in
the axial strut to permit unrestricted torsional rotation.
A similar test setup built by Ekstrom in a university laboratory [9.266] for combined axial,
torsional and hydrostatic loading is shown schematically in Figure 9.88. Axial loading was
applied there by means of a worm gear jack at the base of the test rig, and torsional loading
by means of a lead screw pulling a cable wrapped around the movable lower base plate.
External pressure was applied by evacuating air from the cylinder. Note the thrust bearing,
above the axial load dynamometer, which allowed the cylindrical shell to twist while under
combined axial and torsional loading. The radial load bearing next to it ensured the axial
position of the specimen under the torsional load. The shells tested were made from foil
gage 301 stainless steel and 3003-H19 aluminum strip, by rolling around a thick tube and
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
718 Shell Buckling Experiments

MERCURY
TO VACUUM PUMP+--- iC MANOMETER

FIXED END FLANGE

CYLINDER

MOVABLE
END FLANGE
TORSION
DYNAMOMETER
THRUST AND RADIAL
LOAD BEARINGS

WORM GEAR JACK

Figure 9.88 Schematic diagram of Ekstrom's testing machine for combined torsional, axial and hy-
drostatic loadings (from [9.266])

joining the ends with a lap joint, similar to the specimens of Harris and Suer et a!. Their
geometries covered (Rit) = 267-534 and (L/R) = 2-8.1.
The practice of inserting a thrust bearing between one end plate and the load plates or
axial load jack to permit unrestricted rotation of that end of the cylinder under torsional load
has now become universally accepted (see for example the 1979 UTIAS experiments on
laminated glass/epoxy tape wound cylinders of [9.269] or the 1965 Technion experiments
on conical shells [9.271] to be discussed in Section 9.9).
In the civil and marine engineering industry the test setups are considerably larger, to
accommodate specimens fabricated in a similar manner to that of the full-scale shells. A
good example of such a test setup for combined loading is that built at the Chicago Bridge
and Iron Company (CBI) for their extensive test programs [9.63].
The test setup was built around a specific fabricated test shell, 6.35 m (250 in.) in diameter
and 3.41 m (134.4 in.) long, with (Rit) = 500 and (L/ R) = 1.075, which represented the
space simulation chambers, some as large as 30 m in diameter and 38 m high, which were
built by CBI for the USA space program. The test shell was erected by a field crew using
standard field procedures in order to simulate full-scale construction techniques, according
to the "realistic model" philosophy, discussed in Section 9.3. Tests were performed in shells,
with and without ring stiffeners (see Figure 9.90b), under different combinations of axial
compression, external pressure and concentrated loads.
The test setup is shown in Figure 9.89 and a cross-section of the test shell, with typical
stiffener arrangements, in Figure 9.90. Axial compression was applied to the test shell by
tightening the turnbuckles (A) on the vertical rods (B) shown in Figure 9.90. External pressure
was applied to the specimen by pressurizing the annular space (D) between test shell (C)
and the outer cylindrical shell (£). The test shells were instrumented with approximately
500 transducers that measured applied loads, strains and the radial displacement of the shell
due to the applied loads.
Nine shells were tested under combined axial compression and external pressure. All tests
were in the elastic range, and for all specimens except one, shell no. 9, local buckling failures
occurred (between the rings). In shell no. 9 the small mid-height stiffener together with the
adjacent bays failed by general instability. The shells were first loaded by external pressure
only (actually also with a very small ax1ial load), and then the axial compression was in-
creased in steps with a corresponding decrease in the measured buckling pressure. Typical
results for shell no. 6 are shown in Figure 9.91 compared with interaction formulae proposed
in [9.63]. The same specimen was used for the six combined loading tests, but a degradation
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Combined Loading 719

(a) QNlW. l!ml


--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(b) TOP VI!>/ (c) INSIDE: VI!>/

figure 9.89 CBI test setup for large stiffened and unstiffenecl steel circular cylindrical shells under
combined axial compression and extemal pressure (from [9.63]. courtesy of Dr. C. D.
Miller)

in buckling strength wi th repeated buckling was nOled. Mi lle r pointed o ut that for steel shells,
even with (Ri r) = 500, previous buc kling, primarily unde r ax ial compression. affects the
resu lts. in contrast to the Mylar shells employed in many o f the tests related to aerospace
vehicles, which yield re peatable buckling loads also unde r ma ny load applicatio ns. He nce.
he recomme nded that for interaction data o n thicker steel shells with (RI r) < < 500. separa te
test specimens should be used for each point.

9.8.3 Repeated Buckling Approach


In general. there are two approaches to combined loading tests providing d ata for interaction
curves: (I) use of separate. no minally ide ntical shells for each point on the inte ractio n c urve,
a nd 12) use of re peated buckling of the same specime n for d ifferent load combinations. S inger
e mphasized alread y in the sixties [9.270) that the scatter produced by damage due to repeated
buckl ing in careful tests is m uc h Jess than tha t resulting from the initial di ffe re nces in shape
and propen ies of separate specimens, even if they are manu factured fairly accurately. Th is
is demonstrated by Figure 9.92. whic h presents a comparison of the two a pproaches for the
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
720 Shell Buckling Experiments

~~~RFACE OF TEST
~- CYLINDER

55.8em0.6cm 154.5 em. 77.3em


e
u
111
u
-
I ~-
~
t::l-- N

~
STIFFENER5
I @
R= 317.5em. r
L
6.35~ 76.2 em. 7.94mm

t f2.2clii .
TEST
CYLINDER

~:.!B-
E
..,.
E
r
r
--...Cl
LJ
<ta:
.... o..::>
(!)

...u!:::!z- e
..,.u
I
@_ "'
Ill ~
IJ..
jg:
en en
a:w
en __® ~
"'
0 .... ::>
I ~~ enJ:
tJu
f<l: r
za:
zo
<t"-
en"'
~!!!- r
~
I f2.2clii.
- !::::=
..._ §
R 20.32cm. r-.: ~

I 111
N

I
r· ·• ~ .t· .
..
. ·.b . .
.-- .......
... .,
. 4
•,'<I
<I.
' .. . .. . I . . ...•
..
~ - '.
'
. 4 . ~j

(a)

28.45 em.
t= 6.35mm (typocal) 56.90cm.f
4101.60 x 75.20x.63mm
~ R=317.5 cm(typocal)
(typical unle• •>fherwl..
0 noted)
"': NO STIFFENERS
~ Bar 76.20x .63mm
"'"
..J
L=228.50cm.

L_ ~.6.90 r 28.40 em.


--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

em. 28.45 em.

TEST SETUP NO. I TEST SETUP NO.2 TEST SETUP NO.8

(b)

Figure 9.90 Cross-section of CBI ring-stiffened cylindrical test shell (from [9.63], courtesy of Dr. C.D.
Miller): (a) cross-section of shell in test setup, (b) typical ring stiffener arrangements

conical shells under combined torsion and external pressure, reproduced from [9.270]. The
shells were truncated conical shells with a semi vertex angle of 40°, a small radius R 1 = 1.97
in., a large radius R 2 = 12.50 in. and a thickness t = 0.0157 in., and therefore a mean radius
of curvature to thickness ratio (gj t) ~ 600. The theoretical critical stress for external pres-
sure was 21 percent ofuvicid of the Alclad 2024-T3 aluminum alloy from which the shells
were made, and for torsion the relatively high Tmax was less than half of the yield stress.
Figure 9.92a shows the experimental data obtained from separate nominally identical shells,
correlated to a theoretical interaction curve, whereas Figure 9.92b shows data obtained from
two repeatedly buckled similar shells. The scatter relative to the theoretical interaction is
indeed reduced considerably by the use of repeated buckling. The scatter can be further
reduced by using a more appropriate reference critical pressure, as discussed in [9.270]. It
is evident from the figure that for elastic buckling the second approach of repeated buckling
appears preferable.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Combined Loading 721

u,
uy .4 .3 .2

0
0 EXPERIMENT .I

0
.2
c;;· o.5 u8 ue
uy
.3

TEST NO. 6 .4

Figure 9.91 Test data for a CBI fabricated steel ring-stiffened cylindtical shell under axial compression
and hoop compression, compared with proposed interaction formulae (from [9.63])

This approach of repeated buckling of the same specimen was successfully employed in
an extensive program on the buckling of unstiiiened conical shells under combined loading,
axial compression, torsion and external pressure, carried out at the Technion Aircraft Struc-
tures Laboratory in the sixties [9.270]-[9.272], which will be discussed in more detail in
Section 9.9. It was also widely used with Mylar shells (see for example [9.264]). This
approach was also employed in the eighties by Foster at the University of Tasmania in similar
combined loading tests on thin-walled cylindrical shells [9.267], as well as at the Technion
in combined loading tests on stringer-stiffened cylindrical shells [9.273], [9.274], to be dis-
cussed in Chapter 13.
Already in the early studies on repeated buckling of thin conical shells under external
pressure or under torsion (see for example [9.275]), it was noted that for very thin shells,
for which the maximum theoretical buckling stress was only a few percent of the yield stress
of the material, the pressure or torque could be arrested early enough to ensure practically
entirely elastic behavior. The tests on those specimens could therefore be repeated consis-
tently a number of times on the same cone. As pointed out in [9.272], where repeated
buckling was studied specifically, there was a small decrease in buckling load (of about 1
percent or less per repeated test) due to slight plastic deformation (too small to be detected),
but this was small enough to make repeated buckling a feasible approach to combined
loading tests. One particular 2024-T3 aluminum alloy conical shell, of (Ra) t) ~ 260, was
even repeatedly buckled there 162 times, with a cumulative degradation in buckling load of
only a few percent! These series of tests on conical shells provided definite support for the

---- ---
1.0..-;:::c-------------------J
-- --- -,
!.0·.---;:----------------,

.
~
0.8

.=.as
----;--,,
THEORETICAL 1NTERACTON CURVE
(FROM A•f 9.289 )FOR CONE OF LARGE
e \
~
::o.s
0.8

THECIRETlCAL
-,!--.. . . .
INTERACTION CURVE
'I
\
(FROM R•( 9. 289) FOR CONE Of' LARGE \

.
0 \
Q 6.~
e ~ TAPER RATto (SPECIMEN 421)

i
TAPER RATIO (SPECIMEN 421)
~
\ lr I
"'04 "'0.4 I
\ ~
~
• ALCLAO 40° (SPECIMEN 421 ) 0 421/20 \
I lr b. 421/21 I .O.V\
\ ~ V 421/21 I
0.2 0.2

Q L __ _L_~--~--~-~ Q L __ _ L_ _L __ _L_~--~

0 0.2 0.4 06 08 1.0 0 0.2 0.4 0.6 o.s 1.0


PRESSURE RATIO (A I A~ PRESSURE RATIO (A/ Ao)

(a) (b)

Figure 9.92 Comparison of the separate specimens approach with the repeated buckling of single
specimen approach for aluminum alloy conical shells under combined torsion and external
pressure (from [9.270]): (a) data for separate shells (reproduced from [9.275]), (b) data
Copyright Wiley
--`,`,`````,`,```

for two repeatedly buckled shells


Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
722 Shell Buckling Experiments

recommendation that interaction curves for buckling of shells are best obtained from repeated
tests on single specimens. This was also :reconfirmed later for other combined loading tests
on cylindrical shells.
It may be noted here that in practically all of the combined loading experiments the order
of the application of the loads was found to be immaterial unless significant plastic defor-
mations occurred. This was specifically emphasized by the authors discussing combined
loading tests (see for example [9.184], [9.186], [9.187], [9.266], [9.267], [9.271] and [9.273]).

9.9 Conical Shells


9.9. 1 Conicity Effects and Definitions
Buckling of conical shells is essentially similar to that of cylindrical shells, except when
their small end is much smaller than their large one. The conicity of conical shells is deter-
mined by the cone angle a and the taper ratio 'It defined as:
(9.15)
where R 1 and R 2 are the radii of the small and large ends of a truncated cone (see Figure
9.93). Conical shells with small cone angles are called steep cones, and their buckling be-
havior resembles that of cylindrical shells. Conical shells with large cone angles are called
fiat cones, or shallow cones, and their buckling behavior differs somewhat from that of
cylindrical shells and has some of the characteristics of circular plates.
Except for very shallow cones, the buckling strength of conical shells is usually estimated
by consideration of an equivalent cylindrical shell, with the equivalent geometry given by:
Rcq = Pa, = (R 1 + R 2 )12 cos
L"'~:: e, the slant length
a} (9.16)
feq - f

Since the correlation between conical shells and their equivalent cylindrical ones depends
on the loading, this "equivalent geometry" is merely a convenient convention. The corre-
lation with equivalent cylindrical shells has therefore been studied extensively both in theory

Figure 9.93 Notation for conical shells


--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Conical Shells 723

and experiment (see for example [9.275]-[9.281]). Other definitions of equivalent cylindrical
shells have been suggested, but only that of Eq. (9.16) is widely used.
One should remember that in conical shells, in particular those of large taper ratio, high
stresses occur near the small end under axial compression or torsion loading. Under external
pressure, on the other hand, the large circumferential stresses occur near the large end.
Furthermore, the effective radius-over-thickness ratio, which also determines the imperfection
sensitivity of the shell, is largest near the large end. The buckling behavior depends therefore
on both the type of loading and the taper ratio; and the location of the initiation of buckling
and the prominent buckling deformation is determined by a balance between the location of
high membrane stresses and that of highest effective radius-over-thickness ratio. Under ex-
ternal pressure the buckling waves usually cover the whole shell, with the maximum deflec-
tion nearer to the large end (sec for example Figure 9.94). Under torsion, the buckling

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
deformation generally converges to the small end of the conical shell, where the high shear
stresses appear, especially when the taper ratio is large (see for example Figure 9.95). Under
axial compression, however, though buckling initiates at the small end, where the membrane
stresses arc large, the common two-tier diamond postbuckling pattern soon develops and
covers most of the conical shell (see for example Figure 9.96), or at least its upper half.
High-speed photography of the buckling and postbuckling process of steep and fiat Mylar
conical shells [9.248] confirmed this behavior. It may be pointed out that for conical shells
under combined internal pressure and axial compression, the stabilizing effect of the internal
pressure is greatest in the wider part of the cone (since the meridional and circumferential
tensile stresses due to the pressurization grow with increasing radius). This also shifts the
buckle pattern towards the small end of the conical shell (see [9.282]).
As in the case of cylindrical shells, the boundary conditions, including the in-plane ones,
affect the buckling load and behavior of conical shells. For conical shells, classical simple
supports are usually defined as:
Zero radial displacement w = 0 }
Zero circumferential displacement v = 0
No restraint in the direction of the generators N,. = 0
Free rotation of generators M,. = 0 (9.17)
At the ends s = s 1, s2
These boundary conditions represent bulkheads that are very rigid in the radial direction and
very flexible in the direction of the generators. Another definition of simple supports for
conical shells, by Seide [9.276], replaces w = 0 by the zero displacement normal to the
cone axis:
w cos a - u sin a = 0 (9.18)
while the other boundary conditions remain the same. This represents bulkheads that are
very rigid in a plane perpendicular to the axis and very flexible in the axial direction. The
boundary conditions of Eq. (9.18) offer slightly less resistance to buckling than those of Eq.
(9.17), but since u is of a smaller order of magnitude than w, the difference is small.

9.9.2 Early Buckling Experiments of Conical Shells


Apparently the earliest experiments on buckling of conical shells were those carried out by
Tokugawa in 1932 [9.283]. They and most of the other scarce early tests (for example
[9.284]-[9.288]) focused mainly on external pressure loading, and some of them dealt with
complete cones (which are stiffened by the restraint at the apex). Only a few of the early
experiments considered axial compression or torsion (for example [9.289]). In the late fifties
and early sixties, in parallel with many theoretical studies, extensive experimental programs
on the instability of conical shells under external pressure, axial compression, torsion and

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
724 Shell Buckling Experiments

Figure 9.94 Fully developed elastic buckling under hydrostatic prc.-surc of a thin, electrofonned nickel
truncated conical shell with cone angle a - JO•. taper ratio 'I' = 0.75. (p, Jr)"' 12000
and R, = 5.00 in. (from [9.92))

combined loading were carried out at Space Technology Laboratories (now T RW), Los An-
geles [9. 103), [9.104) and [9.237 1 and at the Techn ion Aircraft Stmctures Laboratory, Haifa
[9.2011, [9.270]- [9.272], [9.275)), as well as some smaller programs at other laboratories
(for example [9.290)- [9.292)).
The experimental investigations on conical shells at Space Technology Laboratories were
pan of the test program, also including cyl indrical shells, that was mentioned in Section 9.4.
The conical test specimens were also mostly Mylar shells, with some steel ones for com-
parison. The si mple test fixture used for conical shells is shown in Figure 9.97 for axial
compression loading. Later versions were modified for the addition of torsion and external
or internal pressure loading. Wi th these relatively simple experimen tal tools, Seide et al.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 9.95 Typical buckle paucrn for a conical shell of large taper ratio subjected to torsion. Only
the plastic deformation remaining after removal from the test rig is shown. The specimen
is a 2024-T3 Alead shell with cone angle & = 30•. taper ratio 'I' = 0.843, (p,J r) = 530
and R, = 317.5 mm (from [9.275])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Conical Shells 725

Fij,'llre 9.96 Typical buckle pauerns for conical shells under axial compression (on the left) and under
torsion (on the 1i ght). Only the plastic deformation remaining after removal from the test
rig is shown. The specimens are 2024-T3 Alcad ;hells with cone angle a = 40", taper
ratio 'V = 0.678 for the A series (upper specimens) and 'i' - 0.500 for the B series (lower

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
specimens). (p.J t) "' 340 and 300 respectively. and R, = 140 nun (from (9.271])

A, TEST FIXTURE
8. LOADING SCREW
C. BALL ..ONT
O. STATHAM UNSOHOEO
STRA IN GAGE LOAD CELL
E. lOADING BALL
f . UPPER CLAMPING FIXTURE
G. LOW'ER CLAMPING f iXTURE
H 8Ast: PLATE
J , CENTER SCREW
K. LOCATING PINS
L . POST
M. SPECIMEN

figure 9.97 Diagram of the simple test fixture for conical shells under axial compression employed
in the Space Tcchnolllgy Labll~<llories experiments in the late fifties (from [9.293] and
(9.294]). In later version; the perspex clamping fixtures were replaced by metal end plates
with grooves into which the shells were fixed with a low melting alllly
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
726 Shell Buckling Experiments

tested many Mylar conical shells, covering a broad range of geometries with (p,) t) "= 350-
2300, and carefully correlated the data with concurrent theoretical studies. Then, after con-
firmation of similar buckling behavior and results by some relevant steel cones, the data
became the basis for early design criteria for conical shells (see [9.103], [9.104], [9.237],
[9.294] and [9.295]).

9.9.3 Technion Experimental Program


In the Technion experimental program, which was planned to complement and extend the
data needed for design criteria, the specimens were made of aluminum alloys and stainless
steel, their scale was mostly about twice that of the Space Technology Laboratories ones
and they were fabricated by standard aircraft manufacturing techniques. The earlier Technion
test series [9.201] and [9.275] focused on loading under external pressure.
In the design of the test rig an effort was made to approach simple supports as far as
possible, in order to facilitate correlation with analyses, which at the time were all for
classical simple supports given by Eqs. (9.17) and (9.18), in contrast to previous experiments
in which the specimens were actually clamped (it being assumed that on account of the low
bending rigidity of the thin wall of the shell, the generators were nearly free to rotate
immediately outside the clamping fixture). A diagram of the test setup is presented in Figure
9.98. Detail A there shows that the end fittings have a circular profile and that the outer
rings have 0-seals at the contact line, so that the rotation of the generators is much less
restrained. The top and bottom fixtures are similar. The displacement along the generators
is restrained by the friction of the seal and inner ring, but the ends are free to move axially
as a whole. At the time, the effect of the restraint on u, along the generator, on the buckling
pressure was regarded to be very small, whereas the restraint on edge rotation was considered
to have a significant effect on it. The test boundary condition appeared therefore to be a fair
approximation to the theoretical simple support conditions of Eq. (9.17).
The influence of partial clamping on the buckling pressure was studied experimentally in
[9.275] by assessing the effect of the tightening of the bolts connecting the end rings. Seven

@ i FIXTURE
ELECTRICAL
LAMP S CIRCUIT

MECHANICAL
PRESSURE GAGE

STATHAM
POTENTIOMETER
PRESSURE
GAGE

'=====9 ~i,~~~~IC
SYSTEM

(b) EIOTTOM ENO FIXTURE FOR


TORSION WITH SMALL RADII

Figure 9.98 Schematic diagram of Technion test setup for conical shells under external pressure (from
[9.201]) or torsion (the torque loading system, which was added later, is not shown).
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under licenseDetails
with WILEYA and B show the end fixtures for externalIncpressure
Licensee=McDermott and torsion
- India/8215328006, respectively
User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Conical Shells 727

thin Alclad specimens with (pa) t) = 530-600, for which repeated tests were easily carried
out, were first attached with simple supports at the top (large) end by minimum tightening
of the top end fixture bolts (fingertight) and tested and retested to the fully buckled state a
number of times. Then the top end fixture bolts were tightened appreciably and the test
repeated with the top end partially clamped. This partial clamping of the top end, in the last
portion of each test, resulted in increases of 10-29 percent in the buckling pressure, which
clearly demonstrated the importance of the boundary conditions for buckling under external
pressure, even for very thin shells. As already mentioned, the rotational restraint imposed
by the partial clamping was considered at the time to be the major cause of the significant
increase in buckling pressure. More recent theoretical studies indicate, however, that the
increased frictional restraint on u, along the generators, showed that for conical shells under
external pressure, in-plane restraint along the generators can increase the buckling pressure
by 20-45 percent, especially in the long and thin shells, as was noted for cylindrical shells
in Subsection 9.7.1. But in cones the influence of the rotational restraint becomes important
for all shells with large cone angles (shallow cones) and not only for short shells, as in
cylindrical shells. It may be pointed out that the stiffening effect of u = 0 compared to
N, = 0 was found to be slightly less for clamped shells than for shells with no rotational
restraints at the ends.
The test rig (Figure 9.98) consisted of the following elements:
1. The heavy pressure vessel and the end fixtures, which form a closed unit with the
specimen. A rotating lever is attached to it to facilitate loading of specimens and to
carry the weights that counterbalance the lower end fixture.
2. The hydraulic system, which supplies the pressure loads on the specimen, either directly
or via air for the thinner shells.
3. The pressure measuring system, which includes a sensitive Statham gage and a parallel
mechanical one, permitting continuous calibration of the Statham gage.
4. A visual electrical contact indication system (discussed in detail in Section 9.5).
5. A strain measuring system, consisting of six strain gages attached at equal distances
along a quarter circumference, was also employed in the first tests but was discarded in
later ones.
This simple, reliable test setup was used for many years (with additions and improvements)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

in the succeeding phases of the cone buckling program.


Since the Technion experimenters meticulously reported their considerations, problems
and attempted solutions, which may be of value in future experiments, parts of their work
are presented here in detail.
Much effort was devoted to arriving at an optimum type of test specimen. In order to
obtain this with standard available sheet thicknesses, shells having large ( Pa, It) ratios, ap-
proaching those used in aeronautical practice, large-sized specimens were required. The
maximum dimensions of the specimens were determined by the size of the available sheet,
and hence the base radius R 2 was fixed at 317.5 mm (12.5 in.), yielding specimens with
( Pa) t) = 126-620. Because the specimens were fairly large, they could be fabricated by
standard aerospace methods, which enhanced the design value of the test results.
Since the aim was to approach an ideal uniform conical shell as closely as possible, lap
joints were dismissed on account of their local stiffening effect, and a welded butt joint was
decided upon. Thus, a weldable alloy was required and the first series of test cones was
made of a French aluminum alloy, AG5-X516, with mechanical properties that differed only
slightly from the American 5052 alloy. The cones were fabricated by Israel Aircraft Indus-
tries, Lod, to utilize their experienced welders. The developed cones were photo-lofted onto
the sheets, cut and rolled into shape and then joined by a butt weld. The manufacturing
problem posed was therefore the production of a butt weld of thin sheet, down to 0.5 mm
(0.020 in.) without weakening the adjacent material and with a minimum of waviness in the
Copyright Wiley
Providedneighborhood ofwith
by IHS Markit under license theWILEY
weld. Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
728 Shell Buckling Experiments

A series of tensile, bending and buckling tests was carried out to assess the quality and
characteristics of the butt weld in thin aluminum alloy sheets. In these preliminary tests,
butt-welded specimens, and butt-welded specimens filed fiat on one side, were compared
with uniform specimens of the same sheet. The tensile tests were the standard ones; in the
bending tests, cantilever specimens with the weld near the root were compared; and the
buckling tests were for simply supported plates (supported at all four edges) with the weld
parallel to the loaded edge. In the latter tests the Euler load was obtained from Southwell
plots. Forty specimens of 0.6 mm (0.024 in.) and 0.8 mm (0.048 in.) thickness were tested.
The results indicated that the butt weld (not filed fiat) caused no reduction in tensile strength,
a very slight reduction in bending stiffness (about 5 percent) and a slight increase in buckling
load (about 5 percent). This average increase in buckling strength was probably due to a
combination of increased friction in the supports and the local lateral stiffening of the weld.
Hence the butt weld seemed suitable. Filing the butt weld fiat resulted in appreciable reduc-
tions in strength and stiffness and was therefore ruled out.
As a further check on the effect of the butt joint on the buckling pressure, 10 specimens
were joined by adhesive bonding instead of welding. A double strap joint, with aluminum
alloy strips of half the shell thickness and bonded by Shell Epon VI adhesive, reinforced by
a very thin fiberglass mat, replaced the butt weld. But though the joint presented a local
reinforcement of the shell, there was some waviness near the joint, which more than canceled
the strengthening effect of the joint. It was therefore decided to use butt-welded joints for
all the other specimens in this series.
In fabrication, argon-arc welding was employed for the thicker shells and oxyacetylene
welding for the thinner ones. Extreme care was taken during welding to reduce waviness,
but some slight waviness always remained near the weld.
The same approach was employed in the second test series, in which the specimens were
made of an annealed French stainless steel, Z I 0 CNT18, a non-heat-treatable stainless steel
similar to the American 18-8-321. This type of steel was chosen because of its excellent
weldability, and extremely careful argon welding resulted in very uniform welds. However,
some slight waviness still remained near the weld as in the earlier aluminum shells for the
first test series. Though the measured initial out-of-roundness (geometric imperfection) of
the steel shells was not worse, and usually even better, than that of the earlier tests, the
external pressure test results for the steel cones were on the average much below those of
the aluminum alloy ones. Residual welding stresses were suspected to be the cause of this
consistent reduction.
Tests were therefore carried out on two typical steel shells, of taper ratio qr = 0.843,
a = 30° and 40° and (p,".lt) = 530 and 600 respectively (see [9.298]). In these special tests
22 and 20 strain gages, respectively, were attached to the rolled shell prior to welding (at
each location, one on each side of the sheet, and both connected in series so as to measure
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

only membrane stresses) and balanced. The shell was then welded in exactly the same
manner as the regular specimens, and finally, after completion, the change in strain was
recorded. Circumferential residual strains of 20-30 X 10- 6 were recorded, which correspond
to residual circumferential stresses of 500-800 psi. Compressive and tensile stresses ap-
peared, which varied and changed signs rapidly along a generator (due to the method of
welding, in which a considerable number of tack welds preceded the continuous seam) but
exhibited only slow changes along the circumference. More extensive testing and strain
gaging would be required to give good quantitative estimates of the welding stresses, but
the tests demonstrated very clearly the presence of residual compressive circumferential
stresses of up to 25 percent and even 45 percent of the maximum critical buckling stress
under external pressure, whose mean effect may reduce the critical pressures by 15-25
percent, as observed. A similar test on an AG5 aluminum alloy specimen with a = 40° and
of double thickness ( Pa, It = 300) revealed similar magnitudes of residual strains, which,
however, correspond to stresses that are less than 15 percent of the critical stresses. Smaller
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Conical Shells 729

reductions in critical pressures of 5-l 0 percent should therefore result. These residual weld-
ing stresses appeared, however, also to be an important factor in the interpretation of the test
results of the first series with butt-welded aluminum alloy cones.
As a further check on the cause of the low results for the steel cones, strain gages were
attached to one steel specimen (no. 311114 of [9.298]), and the theoretical buckling pressure
that would appear for a perfect conical shell was determined from the strain gage readings
by the extension of Southwell's method, discussed in Section 9.6. The slope method men-
tioned there was used, and it yielded a perfect cone buckling pressure of about 1.29 psi, 31
percent below the theoretical critical pressure! Since even after the elimination of the effect
of initial imperfections a very low result was obtained, the residual welding stresses appeared
indeed to be the prime cause of the reduced buckling pressures. Hence it was decided not
to include the results obtained from the welded steel specimens in the evaluation of the
methods of analysis.
On the other hand, theoretical results obtained at the time on the marked inferiority of
stringers as stiffeners for cylindrical shells against general instability under external pressure
[9.299], initiated a review of the earlier dismissal of lap joints on account of their local
stitlening effect. Since fairly strong stringers, distributed evenly around the circumference
of a shell, were found to have raised the critical pressure only by a few percent, it was
concluded that for buckling under external pressure the stiffening of a single lap joint is
entirely negligible. The third series of specimens was therefore made with an adhesive
bonded lap joint. In the absence of weldability restrictions, a material with a relatively high
yield stress, as appropriate for elastic buckling tests, could then be chosen. Ale lad 2024-T3
was used, and the shells were joined by an Epon-Versamid bond, with a very thin fiberglass
inter-surface mat to improve adhesion.
In these bonded 2024-T3 Ale lad conical shells of the third external pressure test series,
the buckling stresses were a much smaller portion of the yield stress. The maximum theo-
retical critical stress for perfect 2024-T3 shells was only 1.6-6.3 percent of the yield stress,
compared to 5-20 percent of the yield stress for the earlier welded aluminum alloy and
stainless steel ones. Indeed, only when the buckling stress was a very small fraction of the
yield stress, as in the case of the thinnest 2024-T3 Alclad shells (for which maximum
theoretical ucnt is about 2 percent of uvicict), could the pressure be arrested in time to ensure
entirely clastic behavior. The tests for these specimens could be repeated consistently a
number of times on the same cone, while even for the thinnest steel cone consistent repetition
was difficult (though occasionally very nearly achieved). The noticeable difference in the
buckling behavior of the Alclad shells was therefore the absence of the nearly instantaneous
transition of the elastic buckling waves to plastic deformation that was always observed in
the earlier tests.
The buckling behavior of the shells was carefully studied during the entire experimental
program. For external pressure loading three values of buckling pressure were recorded in
each test: (l) the onset of buckling, (2) fully developed elastic buckling and (3) plastic
collapse. The onset of buckling is the formation of the first true buckling wave, as distinct
from the initial waviness, which increases gradually with increase in pressure. This gradual
increase in initial waviness is an equilibrium phenomenon, which only near the critical load
transforms into an instability phenomenon (usually with a clear audial indication, "puck").
The onset of buckling is indicated by a loud "puck," accompanied by a small pressure drop
for air loading (when the pressure is applied via air), or by a sudden large pressure drop for
oil loading. This sudden large pressure drop is caused by the slight volume change occurring
at the onset of buckling, which is too small to affect the pressure appreciably in the case of
air loading.
The transition, with slow increase in pressure, from the first wave to the fully buckled
condition is either gradual, the waves appearing one after the other (or in pairs) along the
circumference, each with a "puck," or sudden, the waves appearing along the entire circum-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
730 Shell Buckling Experiments

ference simultaneously with a large "puck." In most of the good specimens, with small out-
of-roundness, the onset of buckling brought out all the lobes and represented therefore also
complete buckling.
After the shells were fully buckled (and repeatedly buckled a number of times), the tests
were continued into the plastic regime till plastic collapse, characterized by folding of buck-
les and a large pressure drop.
Since the standard tolerances for the thin sheet of which the cones were made permit
thickness variations up to ± 10 percent in a batch of the same nominal size, the thickness
of the cones was accurately measured at 25 points for each specimen (and for 10 specimens
at 70 points each) and averaged. The results indicated very small variations (on the order of
0.01-0.02 mm) in the thickness of each ,.pecimen, but slightly larger variations of thickness
between specimens (up to ± 0.03 mm), i.e. up to 3-8 percent). Hence for some specimens
the nominal thickness had to be corrected accordingly.
For a test, the lower end fixture (of smaller radius) was attached first to the specimen. It
was then lowered into position from the counterweight lever and centered prior to final
tightening of the top end fixture bolts. Centering was carried out with a dial gage mounted
on a rotary arm of adjustable length and vertical position. The same device was used to
measure the out-of-roundness of the cones once they had been fixed in position.
For most specimens, the out-of-roundness was measured at a radius slightly larger than
the mean, where the maximum deflection upon buckling was expected. For some cones these
measurements were taken at three radii and compared, and for some the initial out-of-
roundness (initial geometric imperfection) was mapped (see for example Figure 9.99). The
maximum out-of-roundness A0 , an imperfection amplitude parameter which was used exten-
sively in the fifties and early sixties, was then computed from the measurements by Holt's
method [9.300], which defines A0 as the maximum plus or minus radial deviation from a
perfect circle measured over an arc length corresponding to the one-half lobe length of the
out-of-roundness pattern. Typical initial circularity contours for two similar butt·welded alu-
minum alloy shells, and the correlation with onset of buckling, are shown in Figure 9.99.
While the initial out-of-roundness A0 for the two shells, specimens 108/ I and 108/2 of
[9.201], did not differ very much, the shape of their initial circularity contours did. The first
shell (Figure 9.99a) had noticeable initial local waves near the weld, which came out at the
onset of buckling. The second specimen (Figure 9.99b) had a much smoother contour, and
on buckling all waves appeared simultaneously. In this case, specimen 108/2, the initial
contour also biased the circumferential distribution of waves.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

35

2~ 22 21
(a)

Figure 9.99 Typical initial circularity contours (the out-of-roundness is magnified 20 times), and po-
sition of buckles at onset of buckling, for welded aluminum conical shells tested under
external pressure at Technion (from [9.201]): (a) specimen 108/l with noticeable local
Copyright Wiley
Provided by IHS Markit under licensewaves near the weld; (b) specimen 108/2
with WILEY with a smoother
Licensee=McDermott contourUser=G, Boopathi
Inc - India/8215328006,
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Conical Shells 731

It may be noted that considerable effort was expended in the evaluation of the test results
in order to elucidate the causes of scatter. For example, in [9.201] the buckling behavior of
the six shells, yielding the lowest and highest Pcxr/p,h ratios for onset of buckling in each
of the three groups of specimens, was studied in detail and correlated with the measured A0
and with the respective fully developed elastic buckling pressure.
From Figure 9.99 and similar correlation attempts on other specimens from this and sub-
sequent test series, it was concluded that initial geometrical imperfections affected the pre-
buckling behavior and the onset of buckling, but influenced the pressure for complete buck-
ling only in the case of the thinner and "better" specimens of the third series. In general,
the imperfection amplitude, the maximum out-of-roundness A 0 , was found in the Technion
studies not to be a reliable criterion for estimation of the buckling behavior of shells and to
be of less significance than the shape of the initial circularity plot-a finding confirmed later
by the more sophisticated imperfection measurements of the seventies and eighties, which
are discussed in Chapter 10.
The pressure vessel of Figure 9.98 was adapted for torsion loading by adding a stiff torque
arm riding in its main bearing that was attached to the top flange of the vessel. The conical
shells were mounted as before, but in order to ensure that the applied load was pure torsion,
the cone and torque arm floated in a central bearing anchored to the bottom of the vessel.
The torque was applied via load cells by a pair of jacks attached to side frames and fed by
a separate hydraulic system. As shown in Figure 9.98 in detail B (the torque loading system
is not shown in the figure), the end fittings were similar to those used in the earlier tests,
except that the inner rings were serrated to prevent slippage of the specimens. Also, for most
of the cones with small lower end radius, the outer lower ring was replaced by an aluminum
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

ring that had a straight wedge profile instead of the circular one with the o-seal used oth-
erwise. This end fixture was introduced to prevent slippage (note that a small radius results
in very high shear loads), though it caused slightly more restraint. Comparative sample tests
of a typical shell with this fixture and the original one, however, indicated only very small
differences, well within the experimental scatter of identical specimens. In general, the end
fixture bolts had to be tightened much more in the torsion and torsion plus external pressure
tests than previously. This removed the test conditions considerably from the assumed simple
supports, but they were probably still further from the theoretical clamped boundary con-
ditions than some other commonly used clamped ends, such as ends cast in grooves with
Cerrolow.
For torsion, the fitting and centering of specimens and the out-of-roundness measurements
were similar to those for external pressure tests, except that an additional circularity contour
was obtained near the small end of the cone, where the maximum shear stresses occur and
buckling commences. The angle of the twist was measured with a light ray from a scale
reflected with a mirror (attached to the axis of the cone) to a reading telescope. The torque
was increased in small increments till buckling occurred. Buckling appeared very clearly on
the telescope as a sudden running of the scale, or even a strong vibration of the image when
the wave formation was very sudden and violent. Simultaneously, the load cells showed a
sudden drop of torque, which was a very sensitive indication of buckling. Audial indications
("pucks") could not be relied on for torsional buckling, which often occurred quietly (when
it concentrated near the small end of the cone).
Most of the specimens for the torsion and combined torsion and external pressure tests
were made of Alclad 2024-T3 with a bonded lap joint, as for the third external pressure test
series, since it was concluded, from considerations of equivalent stringer-stiffened cylindrical
shells under torsion, that the stiffening effect of the lap joint is only on the order of 1 percent.
As mentioned earlier, the torsional buckling behavior of conical shells of small and me-
dium taper ratios resembled that of cylindrical shells, while for shells of large taper ratio the
waves concentrated near the small end (see Figure 9.95). The experimental results, however,
correlated equally well for all the taper ratios. The measured buckling torques for the Alclad
specimens were 80-97 percent of those predicted by linear theories, but one should recall
Copyright Wiley
that these were for classical simple supports, whereas
Provided by IHS Markit under license with WILEY
in the experiments the ends were partly
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
732 Shell Buckling Experiments

clamped. Also. the interactio n curves for combined torsion and external pressu re for conical
shells of large taper ratio differed from those for cones of smal l and medium taper ratios.
The Technion experimental program on the larger conical shells included 153 specimens
with (p,.Jt)"' 125-725. It was extended by two additional test series, one o n buckl ing under
combined torsion and axial compression [9.271] and o ne under combined axial compression,
torsion. and external or internal pressure (9.272 ]. The 69 specimens in these two series were
somewhat smaller joggle-lap-jointed Alclad 2024-T3 alumi num alloy shells. with cr = 40°,
taper ratios 'i' = 0.500 or 0.678. R, = 140 mm and t = 0.4 mm, and therefore (p,,)t) "'
300-340.
The test setup is shown in Figure 9.100. Axial compress ion. torsion and ex ternal or internal
pressure can be applied simu ltaneously in the load frame. T he vertical cylinder in Figure
9. I 00. control led from an adjacent universal testing machine. applies axial compressive load
via a movable load transfer shaft . a low friction thrust bearing (to el iminate any load cou -
pl ing) and a load cell. Torque is applied by means of two horizomal hydrau lic cy linders that
are control led in parallel from a hand-operated oil pump. The cyl inders act on load cells
attached at either end of the tOrque arm. External pressure is applied by partial evacuation
of the conical vessel. formed by the spec imen and the clamping fixture. wi th a vacuum
pump. For internal pressure the vacuum pump is replaced by a high-pressure system. In
either case the pressure is measured with a mercury manometer, connected to the conical
vessel through a second openi ng in the large clamping fixt ure.
Each of the contoured clamping fix tures used to hold the conical shells consists of a
circular plate whose edge is mach ined at an angle corresponding to the cone angle. Each
end of the specimen is clamped over such a plate by usc of a ring wi th a mating inner edge.
For some specimens it was necessary to kn url the conical edge of the fix ture to prevent
slippi ng of the specimen during torsional loadi ng. The boundary conditions were therefore
close to fully clamped ones. In early tests the narrow end of the specimen was clamped to
the load frame and the wide end to the torque ann. In later tests the specimen was installed
with the narrow end clown. which resulted in better alignment and improved installation
procedure. In either configuration the torq ue arm is suspended from the spec imen. and its
weight must be taken into account d uri ng the tests. For better alignment of the test speci mens
in the load frame, the clamping of the shel l to the smaller end fixture was done in a special
alignment jig. which ensured that the edges of the shell were parallel to each o ther and to
the surface of the clamping fix ture. Initial o ut-of-roundness measurements were taken on
each specimen, with a rotating device. holding a dial gage. that is attached above the upper
clamping fixture (see Figure 9.100).

SPE0r.£N

TO PRESSURE TOROLE I..OAO


SC)JRCE OR .. ~ - .... . ~ YLINOER
Vt(:WM PUMP --/,

........._l'OROUE A.RM

· ---- LOAD FRAME

TORQUE LOAD / AXIAL LOAO


CELL - - CYLINDER
AXIAL LOAD
CEU.

Figure 9.100 Technion test setup for tests on smaller Alclad 2024-T3 conical shells under combined
lo"ds: axial compression and torsion. or axial compression. lorsion. and extemal or in-
temal pressure (from [9.272])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Conical Shells 733

During tests, motion is preserved along and about the vertical axis by means of a guide
pin attached to the small end fixture and a mating sleeve attached to the large one. Rotation
is measured, as in earlier Technion torsion tests, by recording the change in the reading of
a scale, which is reflected by a mirror attached to the lower clamping fixture.
Onset of buckling was determined in the first of these test series by noting the sudden
drop in load or the sudden running of the angle of twist. In the second series more accurate
determination was obtained with strain gages. Up to six strain gages were attached around
the midsection circumference of the shells. This permitted early detection of buckling, and,
as a result, inelastic effects at buckling were minimized. The buckling load could also be
more accurately ascertained by use of gages when external pressure was one of the loads
applied during the tests.
As mentioned in Section 9.8, the approach of repeated buckling of the same specimen
was employed in these two conical shell test series. The average number of tests carried out
on each specimen increased as more experience was acquired and reached 15 for the second
test series. The average rates of decrease of buckling torques and buckling pressure were 0.8
percent and 1.2 percent per test, respectively (with down to an order of magnitude less in
one shell). One should remember that the specimens were fabricated by standard aeronautical
techniques (though with care). The data for the interaction curves indeed showed relatively
low scatter and therefore reiterated the preferability of the repeated buckling approach.

9.9.4 Stanford University Experiments


Many of the lessons learned in the Technion test program on buckling of conical shells under
external pressure were applied in a similar experimental study at Stanford University in the
mid-sixties [9.921. and [9.95]. The aim of that investigation was to settle by careful exper-
iments the then prevalent controversy over the validity of linear theory also for conical shells
of large taper ratios. The design of the test rig (shown in Figure 9.101) strived to achieve
greater accuracy while retaining the simplicity of the Technion setup. The test specimens

SUPPORT
HOOK

Figure 9.101 Stanford University test setup for tests on electroformed nickel conical shells under
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

external pressure-specimen and support structure (from [9.92])


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
734 Shell Buckling Experiments

were very thin electroformed nickel shells, with t = 0.0028-0.0051 in. ( = 0.007·-0.13 mm),
discussed in Section 9.3. Large (paJ t) = 570-1540 could therefore be obtained with fairly
small dimensions, R2 = 3.78 in. ( = 95.9 mm) and 5.00 in. ( = 127 mm).
The rig (Figure 9.101) consists of the following major components: a support structure
and out-of-roundness measuring device, a test specimen, a vacuum system and a pressure
measuring system. The support structure consists basically of the bottom support disk, upon
which rest concentrically an 0-seal, the test specimen and the bottom support ring, in se-
quential order. The top support structure is similar to the bottom one, except that its entire
weight, including that of the out-of-roundness measuring device, is balanced by a gallows-
type counterweight system (not shown in the figure).
The vacuum and pressure control system consists of a small vacuum pump that leads
through a valve to a vacuum storage tank from which the flow may be regulated and mi-
croregulated by two additional valves. One branch of the flow leads to an inclined water
manometer, designed to indicate the decrease in pressure within the specimen to an accuracy
of ±0.0015 psi. The second branch proceeds through a manual and electrical solenoid pres-
sure release valve to the test specimen. The pressure measuring system, though simple, was
found to be very accurate.
In a typical test the specimen was first set up on the support structure, with great care
being taken to ensure proper sealing, centering and support. The vacuum storage tank was
partly evacuated, and finally the actual evacuation of the shell was begun by slowly opening
the microvalve, which completely controls the rate of evacuation from the shell. Throughout
the test this rate was kept down to approximately 0.005 psi/ sec so that the precise buckling
pressure reading could be visually located on the manometer at the instant of buckling. At
buckling, an immediate pressure rise and a "pop" sound occurred. The solenoid valve al-
lowed an immediate release of the vacuum and snap-back of the shell to its original shape
before any noticeable plastic deformation had occurred. This provision was made so that a
given specimen could be buckled a number of times and the successive reductions in buckling
load measured.
Because of the high quality of the shells, buckles appeared around the whole circumference
simultaneously in all of the specimens. Hence onset of buckling and complete buckling
occurred concurrently, as in the better specimens of the earlier Technion tests. All the shells
were buckled a number of times, and the repeated buckling showed a small reduction in
buckling load, with a trend, after 7-10 repetitions, to an asymptotic value of buckling pres-
sure about 7 percent below the initial one, a similar reduction to that observed in the second
test series, with the smaller conical shells . at Technion [9.272].
For the second test series [9.95] some improvements were incorporated in the test rig.
The original gallows-type counterweight system was replaced by a standing type drill press,
which also assisted in the axial alignment during assembly. Great care was taken to obtain
uniform and equal tensioning of the end rirrg units. The shells were all clamped down tighter
than in the corresponding earlier tests, which resulted in clearly recognizable higher buckling
pressures.
The test results of both series conclusively settled the controversy and verified the validity
of the theory also for cones of large taper ratio. The scatter was fairly small, well within
± 7 percent of the average for all the experimental results and much less for the second test
series. These experiments were a good example of a successful use of electroformed spec-
imens, discussed in Section 9.3.
Many of the techniques employed in all these experiments in the sixties may today seem
rather naive, in view of the sophisticated modern measurement and recording systems, but
the concerns, considerations, emphases and most of the results still fully apply today and
will probably also apply to future tests. Therefore it was felt that their rather lengthy dis-
cussion was warranted. --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Conical Shells 735

9.9.5 Ghent University Tests on Liquid-Filled Cones


To conclude this section on conical shells, a recent experimental program that emphasized
the effects of conicity of the shells is discussed. For over a decade Vandepitte and his co-
workers at Ghent University studied the buckling of liquid-filled conical shells [9.301]-
[9.304]. As pointed out by the researchers, the motivation for this study originated in the
collapse in 1972 of a large steel water tower with a conical bottom while it was tested by
being filled for the first time. The investigation into the cause of the collapse, which included
tests on Mylar models, "showed that the accident was caused by buckling of the thin conical
wall of the tank in the region of its smallest radius."
When a conical shell with a vertical axis and supported on its lower edge, which has the
smaller radius (see Figure 9.102), is partly filled with a liquid, compressive stresses arise in
the meridional direction and tensile stresses in the circumferential direction. The compressive
meridional stresses acting along a parallel circle AA' (in Figure 9.1 02a) equilibrate the weight
of the liquid contained in the toroidal space of triangular cross-section, bounded by the wall
of the shell and by the vertical cylinder ABB'A', as well as the weight of the shell down to
circle AA' (which is usually small compared to the weight of the fluid). The weight of this
toroidal mass of liquid increases more than linearly as we lower the parallel circle considered.
Moreover, the perimeter of the parallel circle then decreases. The compressive meridional
membrane stresses therefore increase rapidly, for both these reasons, as the base of the shell
is approached. At the lower end, the compressive meridional membrane stress a is given by
yh 2 [R 1 + (h/3) tan a] tan a
a= (9.19)
2R 1t cos a

when the weight of the shell itself is neglected. Note the highly nonlinear dependence on
the liquid level h. As this h gradually rises, the compressive meridional stresses eventually
cause the bottom part of the shell to buckle, in spite of the stabilizing effect of the concurrent
circumferential tensile stresses (due to the normal pressure exerted by the liquid). Since the
load is a gravity load, the shell fails suddenly and catastrophically.
Vandepitte and his colleagues carried out an extensive experimental program involving
tests on 694 specimens made of Mylar, brass, aluminum and steel, as well as many numerical

Iih

4u;
!
L - --c 1c'
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(a)

Figure 9.102 Liquid-filled conical shells: (a) geometry, (b)


small end boundary conditions for Ghent Uni-
(b) versity specimens (from [9.302])

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
736 Shell Buckling Experiments

calculations with two programs (Bushnell's BOSOR 4 and 5, [1.27] and [2.62], and Esslin-
ger's F04B08, [9.305]). It may be of interest here to quote from their 1988 paper [9.304]:
The writers would not venture to propose as a basis for design against buckling of shells information
resulting exclusively from calculations. The reasons are that:

1) experimental results in this field are more conclusive than theoretical results, although numer-
ical shell buckling calculations deserve increasing confidence;
2) the BOSOR programs and F04B08 program can make allowance only for axisymmetric im-
perfections, whereas real shells and also our shell models have random imperfections.

The writers regard results of numerical calculations as valuable background and corroborative
material.

As pointed out in Chapter 10, numerical calculations can indeed be a reliable basis for
design only if the initial imperfections are known, either from actual measurements or from
an imperfection databank, also to be discussed in Chapter 10.
Since buckling deformations occur mainly close to the lower and smaller edge, the bound-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

ary conditions there are of prime importance. The Ghent researchers tried different types of
edge supports: clamping of Mylar cones by casting into resin, clamping of metal cones
between two mating outer female and inner male steel cones, simulating real conical shell
structures for Mylar specimens with a special set of steel rings that approach simple supports
with weak elastic restraints (see [9.302]). Finally they arrived at the support shown in Figure
9.102b, where the lower rim of the test cones was placed on a thick conical steel bearing
plate whose generatrices were perpendicular to those of the test specimen. This hardly im-
pedes rotation of the lower edge of the specimen but could result in sliding of the bottom
edge, which was therefore prevented by placing 16 studs around and against its perimeter
and by fastening them to the bearing plate. It was not easy to achieve contact between the
bearing plate and the test cone all along their circumference. Whenever a slit was noticed
between the cone and its support, calibrated shims, not just one shim but a series of different
thicknesses, were inserted into and over the length of any gap wider than 0.05 mm. Finally
a soft plastic material was smeared into the groove between the two conical surfaces in order
to prevent or minimize leakage of fluid during the test. In all types of support the upper
edge is completely free, which approximates real structures.
The fabrication of the metal conical shells presented some problems. The sheets of steel
needed to make a test cone were too thin to be welded. The joints between the sheets were
therefore soldered. The soldered joints had to be protected with a special coating. In the
absence of such protection, the mercury (which was the liquid employed in some tests, as
will shortly be explained) rapidly formed an amalgam with the solder during the test and
caused the tensile strength of the joint to vanish almost completely.
Furthermore, since no machine was found that was capable of rolling sheet steel suffi-
ciently accurately into the desired conical form, a simple device was designed and made in
the Ghent laboratory (see Figure 9.103).
As in any modem shell experiment, the initial geometric imperfections were measured
and recorded, but measurement could be limited here to the lower part of the cones, to which
the buckling deformation was confined.
Most of the 674 Mylar and metal conical shells were tested with water as the loading
medium. The loading was increased in 19 steel models by filling them with a mixture of
molasses and water (about 1.35 times heavier than water) or with a suspension of barium
sulfate (about twice as heavy as water). The results obtained with the heavier liquids verified
the functional influence of the specific weight obtained earlier by dimensional analysis (see
[9.302]). To explore the elasto-plastic buckling range, higher compressive stresses were
needed. This could be generated either by very large steel models filled with water or by
conical shells of moderate size (say, with an upper rim of 1.75 m diameter) and loading
them with a fluid many times heavier than water, which in practice meant mercury.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Conical Shells 737

Figure 9.103 Special device for rolling steel sheet into a conical shape
designed at Ghent University (from [9.304]. courtesy of
Professor D. Vandepille and Professor G. Lagae): (a) the
device. (b) making of a conical shell with the device

Earlier, two steel cones of large size (diameter of the upper rim = 7.50 m) and with a

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
rather thin wall (1 = I .5 mm) had been tested in order to investigate the infl uence of welds
on the buckling load (9.302]. But making and testing a series of such large models- and
even larger ones, some too high to fit inside the laboratory, that wou ld have been required-
would have been prohibitively expensive. On the other hand, mercury too is expensive and
requires great care in handl ing to prevent health hazards from contact with it or its vapor.
After the pros and cons were considered, mercury was chosen and 24 steel cones were tested
with it as the loading flu id.
T he test setup for the mercury loaded shells is shown in Figure 9.104. Basically, it is
similar to that employed in the earlier tests. However, in order to decrease drastically the
quanti ty of mercury required, a second steel cone was placed inside the test cone. The inner
cone was tilled with concrete to prevent it from being cntshed by the pressure exerted by
the mercury. Notwithstanding the concrete ballast, the inner cone had to be anchored to the

Figure 9.104 Test setup for mercury loaded conical shells at Ghent University (from [9.304] , courtesy
of Professor D. Vandepille and Professor G. Lagae)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
738 Shell Buckling Experiments

bearing plate by means of a strong anchoring bolt (see Figure 9.104) to keep it from floating
on the mercury. The space between the tc;;t model and the inner cone, later to be filled with
mercury, was about 1 em wide.
In order to prevent complete collapse of the test cone after its failure and minimize contact
between the mercury and the air in the laboratory, a two-piece steel structure into which a
third cone was incorporated was placed around the test cone.
The mercury needed for loading the test cones was pressed from a steel container of
100 L capacity through tubing and through a hole in the bearing plate into the space between
the test cone and the inner cone by pressurizing the air in the container. After the model
had failed, the mercury flowed back into the container. Great pains were taken to avoid
spillage of mercury and protect the laboratory personnel from it and its vapor.
Vandepitte and his co-workers at Ghent University attempted early in the program, without
success, to interpret the experimental results with reference to existing design formulae for
cylinders subjected to simultaneous axial load and internal pressure. They concluded that
because of the considerable physical differences, the most effective approach was to deal
with the stability of liquid-filled conical shells as a separate problem. As was seen, the
conicity effects dominated these conical shell experiments, which include many unique fea-
tures. In addition to the development of special experimental techniques for conical shells,
the results of this extensive test program and its concurrent numerical studies also produced
a basis for design.

9.1 0 Spherical Shells


As in the case of cylindrical shells under axial compression, the buckling and postbuckling
behavior of spherical shells under unifom1 external pressure has challenged and sometimes
perplexed investigators for decades. Extensive theoretical and experimental investigations
have been carried out, and interaction berween analytic and experimental studies has been
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

essential for progress (see for example Kaplan's review, [9.224], or that of von Karman and
Kerr [9.306], or Kollar's review, [9.307]).

9. 10. 1 Spherical Caps


The first experimental study on the buckling of spherical caps under external pressure by
Bach in 1899-1902 ([9.16], mentioned in the historical introduction) preceded the earliest
classical analyses by nearly two decades. The experiments on wrought copper and wrought
iron specimens showed clearly the violent nature of the buckling, as well as that the buckling
deformations were asymmetric and occurred near the edges of the spherical cap. The buck-
ling pressures measured were about 15-30 percent of what the later classical theory would
have predicted, similar to those measured in the tests half a century later. However, Bach's
experiments were apparently forgotten even in Germany.
The classical theory for buckling of thin spherical shells under uniform external pressure,
developed by Zoelly [9.308], Schwerin [9.309] and van der Neut [9.310] in 1915--32, arrived
at a critical membrane stress for a complete spherical shell identical to that of a cylindrical
shell under axial compression:
uc\ = [3(1 - v2 )] 112
E(t!R) = 0.605E(t/R) for v = 0.3, (9.20a)
where R is the radius of the sphere. Note that in Chapter 2 a similar result was obtained for
a shallow spherical shell under external pressure (except that t was denoted h there). The
corresponding critical pressure, Eq. (2.272) there, for sphere and spherical cap alike, is
P<:~ = 2uc 1(t/R) = 2£[3(1 - v 2 )]- 11 '(t!R) 2 = l.2lE(t!R)" for v = 0.3. (9.20b)
In the forties, fifties, and sixties most of the analytical and experimental investigations
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Spherical Shells 739

dealt with spherical caps rather than complete spheres. The geometry of a spherical cap is
determined by a parameter
A = 1'12(1 - v2 )(S/VRt) = V12(1 - v 2 ) VRfi a (9.2la)
where S = aR is the arc length and a the semi-apex (or semi-included) angle of the spherical
cap. For shallow spherical shell caps a < < 1 and the shell segment geometrical parameter
becomes
(9.2lb)
where a is the base radius and His the rise (or height) of the cap. A increases as the spherical
caps become deeper and thinner.
Since the classical buckling theory for spherical shells indicated that the complete spheres
develop small local buckles, the shell in the vicinity of the buckle could be considered
shallow, and therefore the behavior of a clamped shallow spherical cap appeared to be the
key to the understanding of the behavior of a complete spherical shell. Hence most of the
earlier efforts focused on spherical caps, which also experimentally seemed more tractable.
Already the exploratory tests of E.E. Sechler and W. Bollay in the thirties (for the design
of the Mount Palomar Observatory Dome, see [1.22] and [9.32]), on thin-walled spun copper
hemispherical shells (with Rlt as high as 1800) immersed in a mercury bath showed that
the measured buckling pressures were about one-fourth of those predicted by the classical
theory of Eq. (2.272). As mentioned, the much earlier (and forgotten) Bach experiments had
yielded similar low buckling pressures. Another set of exploratory experiments was those
carried out on clamped spherical shell segments under approximately constant volume load-
ing at the Caltech in the early forties by L.G. Dunn and H.B. Crockett (and reported by
Tsien in [7.15]). They resulted in somewhat higher buckling pressures, but still considerably
below the classical theory predictions. These larger discrepancies motivated the use of non-
linear analyses, initiated by van Karman and Tsien in 1939 [9.412] and developed further
by many prominent investigators since, which are briefly discussed in Chapter 3.
The surprising scarcity of experimental work on the buckling of spherical shells in the
first half of the century can perhaps be explained by the difficulties in the fabrication of
satisfactory spherical specimens (for example, Bach's 1901 copper spherical caps were fab-
ricated by a manufacturer of pans for breweries and cooking utensils). When special forming
techniques for spherical caps evolved in the fifties and sixties (discussed in Section 9.4),
there was a sudden gush of experimental investigations employing similar types of specimens
(for example [9.122]-[9.127], [9.143]-[9.145], [9.150]-[9.154] [9.313], and [9.314]). It was
also noted there that it was the search for good spherical shell models that first motivated
the development of electroforming [9.85].
The first systematic experiments on spherical caps were carried out in the early fifties by
KlOppel and Jungbluth at Technical University Darmstadt, Germany [9.311] and Kaplan and
Fung at Caltech [9.223].
The Darmstadt specimens were made from ductile steel sheet, drawn (or pressed) onto
special molds. They were medium and deep spherical caps of R = 520 mm and R = 250
mm respectively, with a base radius a = 200 mm, and hence with a shell segment geometrical
parameter range A = 18-31. Over 40 specimens with (Rit) ~ 300-1900 were tested. The
shell thickness measurements showed deviations from the mean of up to 5 percent in the
shallower shells, of R = 520 mm, and up to I 0 percent in the deep ones, of R = 250 mm.
The imperfections of the shells (mainly variations in radius of curvature and thickness and
non-uniformity of the boundary condition) were revealed by the irregular stress field recorded
by the strain gages bonded to the specimens. This motivated stress relieving to verify that
the residual stresses were not a primary cause of these measured stress irregularities, and
indeed stress relieving did not significantly change the stress pattern.
In the test rig, air pressure was applied to the shell either by vacuum inside the cap, or
by external pressures. The small pressure drop during snap-through depends on the ratio of
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
740 Shell Buckling Experiments

the small volume enclosed by the cap to the much larger volume of the storage tank. Buckling
therefore represents approximately a constant pressure process. The measured buckling pres-
sures were 0.11-0.27 of the classical theoretical values.
Buckling failures were sudden and noisy and did not allow a clear observation of the
snap-through. Kli:ippel and Jungbluth therefore employed high-speed photography (1000
frames/sec) to study the buckling process. Thus they observed that the snap-through started
with a localized asymmetric buckle (usually at the boundary), even when the final pattern
was axisymmetric.
The specimens of Kaplan and Fung at Caltech were made by spinning from fiat sheet.
After unsuccessful attempts to heat-treat aluminum spinnings, magnesium alloy QQ-M-44
was selected because of its relatively high (0') E) ratio (of yield stress to Young's modulus,
which is the prime suitability parameter for elastic buckling and postbuckling models, as
discussed in Section 5.6 of Chapter 5). Si.nce the magnesium was spun while hot, most of
the fabrication residual stresses (one of the sources of imperfection) were eliminated. The
specimens were shallow spherical caps, and their spinning proved to be difficult, but after a
combination of spinning on concave and convex molds good-quality shells were obtained.
Kaplan and Fung tested 23 spherical caps, having a base radius a = 4 in., nominal radii
of curvature of 20 in. and 30 in., and thicknesses varying from 0.029 to 0.101 in .. and hence
with (R/ t) = 198-1034 and a shell segment geometrical parameter range A = 4.04-10.1.
The edges of the specimens were clamped between two rings, which were bolted to a heavy
circular plate, thus providing clamped edges and a closed pressure chamber. For each of the
two radii of curvature a separate set of clamping rings was used. The shells were subjected
to either oil pressure (in 17 of the specimens) or air pressure (in 6 of them). The oil pressure
tests provided an approximation to a com:tant volume process, while the air pressure tests,
in which an accumulator tank was connected close to the test rig, provided an approximation
to a constant pressure buckling process. Similar buckling loads were obtained with both
systems of loading, the rigid system (oil) and the soft system (air), contrary to the predictions
of von Karman and Tsien [9.412]. This insensitivity to the rigidity of the test system was

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
reconfirmed later for complete spherical shells [9.84], to be discussed below) and for cylin-
drical shells (see Subsection 9.2.2).
The results of the Kaplan and Fung tests and those of Kloppel and Jungbluth are presented
in Figure 9.105, in which the results of some of the significant test series for spherical caps
under external pressure carried out in the fifties and sixties have been compiled. Additional
test series are presented in Kaplan's 1974 review [9.224], where a list of experiments with
some details is also given, and in Kollar's 1982 review [9.307]. The results of the earlier

~--------------------~~------------------,-----------
AISYMMETRIC NONLINEAI~ THEORY

1.1

I \ COMPLET~ SPH~R,E
0 KLbPPEL &

. ..
JUNGBLUTH
CLASSICAL THEOR'T' FOR _ _ _ _ _ _ _ __ -- (1953) Ref 9 311
1.0
e KAPLAN a FUNG
0.9 .&. .&. .t. x x \ AS'1MMETAIC NONLINEAR THEORY (1954) R"f 9 223

0.8 "" ............. ' .. [). t\.' "'' 8 0 HOMEWOOD, et ol,


(1961) Ref 9 145
1J t '
ll(l(ooP'K xClxfi~xD ·~
A EVAN-IWANOWSKI
07 .&. 8; LOO
(1962:) Ref 9 313
v xX .p
PeL 0.6
V X PARMETER
( 1963) Rtf 9 96
0.5
+ LITLE
(1964) Ref 9 314
0.4
C. ADAM& KING
(196~) Ref 9122
0.3
0 a
KRENZitE KIERNAN

o.2 ~ o 0 @B
(19641 Ret~ 9 151
9.312
a.

01~ VWANGelcl

al~----~----~~----~·----~----~~----~--J__"~_•_'_"'_'._'_'·-~
0 10 15 >.. 20 25 30

Figure 9.105 Experimental results for clamped spherical caps tested in the fifties and sixties (from
Copyright Wiley [9.122], [9.145], [9.151], [9.224], [9.311]-[9.315) and [9.318))
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Spherical Shells 741

experiments in Figure 9.105 (Kaplan and Fung, KlOppel and Jungbluth and Homewood et
al.), which employed drawn and spun metal specimens, exhibit a continuous trend. There is
a dip in the buckling pressure between A = 4 and 6, and then as A increases the buckling
pressure decreases monotonically. In the later experiments shown in the figure (Wang, Adam
and King, Evan-Iwanowski and Loo and Litle), better specimens were made of PVC or
similar plastics by thermal forming, a process discussed in Section 9.2. These better or more
near-perfect specimens significantly reduced the discrepancies between experiment and the-
ory, especially the nonlinear axisymmetric [9.315] and asymmetric theories [9.316] and
[9 .317] that had been developed. The buckling pressures were also higher since the edge
supports used in these experiments minimized the residual stresses that arise due to clamping,
though the edges were still weaker than with fully rigid clamping, as will be discussed later
in this section.
Two of the test series shown in Figure 9.105 employed other techniques for fabrication
of the specimens. Krenzke and Kiernan [9.312] machined their shells very carefully from
solid 7075-T6 aluminum alloy. Their shells were relatively thick, with (Rit) = 76-213 and
A = 10.7-30.0, and the end rings were an integral part of the specimens, resulting in high-
quality shells with practically stress-free, nearly complete clamping. Parmeter's shells, on
the other hand, were very thin electrofonned copper specimens, with (R/t) = 740-4350 and
A = 5.7-20.3, which were nearly perfect and therefore buckled at pressures fairly close to
the predictions of classical linear theory, (Pexr/ Pc~) = 0.62-0.90. In order to mount the thin
electroformed shells into the clamping rings without inducing stresses or moments and with-
out distorting the spherical shape, Parmeter set the edge into an epoxy glue, confined to the
two matching end rings (see [9.317]). One may note that the results of both Krenzke and
Kiernan and Parmeter for those spherical caps for which A > 5.7 do not significantly depend
on A, as is the case for asymmetric nonlinear theory. The results of Adam and King as well
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

as those of Wang and Evan-Iwanowski & Loo for A > 9 and those of Litle show a similar
trend. (There are discrepancies between some of the experimental points in Figures 9.105
and 9.106 and those in the corresponding figures in Kaplan's 1974 review, [9.224]. It should
therefore be noted that here the experimental points were plotted directly from the original
references.)

9. 10.2 More Recent Experiments on Spherical Caps-Effect of Boundary


Conditions
The later experiments on clamped spherical caps, the results of which are shown in Figure
9.106 (Kriiger and Glockner and Sunakawa and Ichida), reconfirm this trend. Tillman's re-
sults, also shown in the figure, are for A < 6.4 and mostly for relatively shallow shells,
which buckle axisymmetrically and confirm Budiansky's axisymmetric nonlinear theory
[9.315]. They therefore do not relate to the large A trend. The specimens were again made
from PVC or similar plastics by thermal forming, a technique that by then had been per-
fected, accumulated considerable experience and produced very accurate spherical caps.
As pointed out earlier, the boundary conditions exert a very significant influence on the
buckling behavior of spherical shells. Wang et a!. [9.124], [9.319] and [9.320] and Litle
[9.314] extensively studied these effects. In his test setup Wang employed different edge
supports for his spherical caps (see Figure 9.107) and compared their influence on the buck-
ling pressures. The specimens were polyvinyl chloride (PVC) spherical caps with R = 18
in., (Rit) = 314-643 and were tested in two series. In series A, which included 52 tests on
18 shells, each specimen (except for two) was tested with fixed, hinged and roller-supported
edges, and therefore the changes in buckling pressure could be attributed to the difference
in support conditions alone. In series B, which included 35 tests, five additional boundary
restraints were studied. All models were loaded by evacuating air from beneath the shell.
The roller-supported edge is shown in Figure 9.107a. The edges of the specimens were
cut in a lathe to ensure that they were in one plane, and therefore the fiat ftexiglas base met
the shell everywhere providing vertical support. The horizontal displacement restraint was
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
742 Shell Buckling Experiments

p
Fh AXISYMMETI~IC NONLINEAR THEORY ------~
1.1 o TILLMAN 119701
't-~~~~ 1~~~~ FOR Ref. 9125
1.0
---------------r·r~m~~~rt-r4~
0.9 Rol. 9.127
ASYMMETRIC (MEAN Of DATA
0.8 lf,.A~__,._--t----t---~~N~~~~ ~~!'!_Y__ • hN>. ~E~MENJS
VARIANC~
0.7 .O.KRUGER"ii

0.6
t GLOCKNER
(1971)
Ref. 9.126

0.5
f
0.4

0.3

0.2

0.1

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
0 ___.L -~ ~~~~~~--J.-~~_L---~
00 5.0 10 15 >. 20 25

Figure 9.106 Experimental results for clamped spherical caps tested in the seventies (from [9.224],
[9.126] and [9.127])

limited to the friction between the shell edge and the base, which was minimized by the
pool of oiL Buckling tests with oil having three different viscosities yielded the same critical
pressures.
The hinged edge, shown in Figure 9.107b, was intended to provide complete restraint to
vertical and horizontal edge displacements while permitting free edge rotation. The thin Saran
membrane prevented contact between the shell and the epoxy cement, and therefore the free
edge rotation was only slightly hindered by friction between the shell edge and the
membrane.
The fixed-edge condition, shown in Figure 9.107c, is obtained by the epoxy cement, which
bonds firmly to the specimen and to the heavy aluminum ring. This eliminates the bending
stresses that usually accompany mechanical clamping of the edge.
The influence of the boundary conditions on the buckling pressure is shown in Figure
9.108, where the results of the shells of series A are plotted versus the shell segment ge-
ometry parameter A. This series included combinations of three basic radii, a = 6, 9 and
12.875 in., and two groups of thicknesses, Group I with t = 0.0278~0.0290 in. and Group
II with t = 0.0533-0.0574 in., yielding a range of A = 10.7-32.3. For each shell, the buckling
pressure for the fixed edge condition is larger than that for the hinged-edge condition, and
the buckling pressure for the roller-supported edge is significantly lower in each case than
that for the other edge supports. Hence it is apparent that restraint of horizontal edge dis-
placement is more influential than restraint of edge rotation. Note also that the buckling
pressure ratios (Pex/Pc;) were practically independent of A and of the base radius but were
slightly higher for the thicker shells.
In series B additional boundary conditions were studies for one nominal shell geometry,
R = I 8 in., a = 9 in. and t = 1132 in. In one type of edge support the shells were rigidly
attached with epoxy cement to simply supported edge rings (Figure 9. I 07d). Both PVC rings
(denoted SP) and aluminum rings (denoted SA) were used. Since the restraint imparted by
the edge rings depends directly on the modulus of elasticity of their material, the edge
restraint of the SA models was 22.5 times that of the SP models. Then three types of back-
to-back edge supports were investigated (Figure 9.107e and f). In the back-to-back config-
uration, symmetry imposes a condition of zero edge rotation (and therefore complete clamp-
ing) up to the point of buckling. However, when buckling occurs near the shell edge, one
cannot assume complete edge fixity, and edge rotation depends on the ring flexibility. The
three types of edge support were: (I) essentially
Copyright Wiley
Provided by IHS Markit under license with WILEY
no edge ring (DN, as in Figure 9.I07e); (2)
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Spherical Shells 743

(b) (e) ~NOMETER

SHELL SPECIMEN
113"
(c) (f)
Tt
!18"F~--========~
TO E0GE
MANOMETER RING
SHELL SPECIMEN

Figure 9.107 Experimental edge supports for PVC spherical caps tested by Wang eta!. (from [9.124]):
(a) roller supported edge, (b) hinged edge, (c) fixed edge, (d) shells rigidly attached to
PVC or aluminum edge rings, (e) back-to-back arrangement without edge rings, (f) back-
to-back arrangement with heavier PVC or aluminum edge rings

a PVC edge ring (DP) having twice the cross-sectional area of that employed in the SP
models of Figure 9.107d, and (3) an aluminum edge ring (DA) with twice the cross-sectional
area of that used in the SA models (see Figure 9.107f). With these stiffer rings, buckling
occurred in the DP and DA models sufficiently far from the edge to justify the assumption
of complete edge fixity.
Wang et al. also studied the effect of applied edge moments and other boundary distur-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

bances on the buckling pressure (see [9.320]). Similar studies on the influence of different
edge supports and of applied edge moments on the buckling pressure of spherical caps were
carried out earlier by Wang's teacher Litle at the M.I.T. Laboratory for Structural Models
(see [9.314]). Litle also compared buckling under air pressure loading with that under discrete
weight loading and concluded that provided the grid spacing is small enough, discrete load-
ing systems can be satisfactorily substituted for continuous ones. He noted, however, that
the buckle position was sensitive to the load direction and that with hanging weights the
edge conditions were even more critical.
Not only the thermally formed plastic test specimens but also the test rigs of most of the
spherical cap experiments in the sixties and seventies were similar. Their essential elements

r- ..
Pupli
pel
08

0.6-"
~
----




-------~---

.,
••+_,_'•'

••
~::OLLER .-G-ROUF' 1

I
,
t:. ROLLER , GROUP U
I + HINGED • GROUP I

~~. HINGED • GRouP n ,

"'l
+
+
t I· FlXEO • GROUP I I

I· FIXED
I
,G_ROUP llj
I
~&
0: ~ cfl&. 1!1> I
i
_l_ ___ __________l_____ J
10 14 18 22 X 26 30 34

Figure 9.108 Influence of edge supports on the buckling pressure for PVC spherical caps tested by
Copyright Wiley Wang et a!. (from [9.124] with additional data)
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
744 Shell Buckling Experiments

appear in the example of Figure 9.47, presenting a later test setup of Yamada and Yamada
[9.211], shown earlier in Section 9.5 in connection with the discussion on recording of
buckling. It may, however, be useful to examine one typical test setup in more detail. The
1974 one of Sunakawa and Ichida [9.127] has been chosen as an example.
This experimental setup at the Institute of Space and Aeronautical Science, University of
Tokyo, is shown schematically in Figure 9.1 09a. The test specimen (1) was bonded with an
epoxy resin adhesive (3) to the relatively rigid aluminum alloy supporting ring (2) to provide
clamped boundary conditions. Details of a supporting ring, in particular the steps to ensure
exact, rigid and repeatable bonding, are shown in Figure 9.109b. There were two such rings,
one for specimens with base radius a = 100 mm and one for those with a = 150 mm. The
efforts to ensure exact and complete clamping (which included extensive and repeated mea-
surements of the mechanical, physical and chemical characteristics of the epoxy adhesive)
are reported in more detail in [9.321]. The external pressure was applied to the specimens
by evacuating the air beneath them. The test system was designed to ensure carefully con-
trolled quasi-static loading of the specimens. A vacuum pump (4) evacuated a vacuum tank
(5) and via regulating valves (7), a loading tank (6), which again via a fine regulating valve
(8) evacuated the space under the specimen. The pressures were measured in the tanks and
underneath the specimen by a manometer (I 0) and pressure transducers ( 11 ). The output of
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

these transducers was transmitted via an amplifier (14) to a recorder (15). Strain gages of 2
mm gage length ( 13) were attached to both sides of the specimens in the meridional and
circumferential directions, and their output was decomposed into in-plane and bending com-
ponents and recorded via an amplifier. Five differential-transformer-type displacement trans-

CD SPECIMEN @ LOADING T~NK @ PRESSURE TRANSDUCER


(2) SUPPORT RING (/) REGULATING VALVE @ DISPLACEMENT TRANSDUCER
@ ADHESIVE @ FINE REGUL.~TING VALVE @STRAIN TRANSDUCER
@ VACUUM PUMP @ EXHAUST v.~LVE I@) AMPLIFIER
@ VACUUM TANK @MANOMETER @I RECORDER

(a)

LPRESSURE
TRANSDUCER
~400

Figure 9.109 Sunakawa and lchida's test setup for spherical caps under external pressure (from
[9.127]): (a) schematic diagram of setup, (b) details of an edge supporting ring
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Spherical Shells 745

ducers (with semi-spherical ends of 2.3 mm radius and a measuring force of 8 gr.) were
installed on a special displacement scanning device, permitting measurement of the displace-
ment over the entire surface of the test specimen.
A series of preliminary experiments were performed with the displacement scanner and
strain gages to confirm that the prebuckling deflection pattern and strain distribution were
axisymmetrical. Indeed, in all the more than 100 cases measured the prebuckling deforma-
tions were axisymmetrical. After these preliminary experiments the locations of the displace-
ment transducers were fixed, and then changes in displacement pattern and strain distribution
with external pressure were recorded in each test.
All the instruments used in the experiments were repeatedly calibrated on-line. The ex-
ternal pressure was applied gradually and the magnitudes of load, deflection and strain were
recorded continuously. When necessary, the load was held constant to permit detailed mea-
surements of deflection and strain distributions. These measurements were canied out until
buckling, and for some cases snapping back was also checked. In the tests measurements
were repeated and excellent repeatability was obtained.
As mentioned in Section 9.3, the thermal forming process of the PMMA (perspex) was
carefully controlled (see Figure 9.22) and the clamped shells were nearly perfect. The 174
specimens tested covered a wide range of geometries, with radii of curvature R = 300-1500
mm, A = 2.64-22.5 and semi-apex angles a = 0.067-0.52 radians (or 3.8°-29.8°).
The experimental results are presented in Figure 9.110 versus the shell segment geomet-
rical parameter A. The mean of the experimental data in segments of ~A = 2 appears in
Figure 9.1 06, together with other test data obtained in the seventies. One notes that the
experimental buckling pressures for these carefully fabricated PMMA (perspex) caps are
very close to the classical buckling pressure of a complete spherical shell. The buckling load
is seen to be nearly constant up to A = 13, and thereafter it decreases with A. There appears
to be no minimal point near A = 4 for these shells, whereas established nonlinear theory for
symmetric buckling of perfect spherical caps clearly predicts there a large trough, or valley.
This significant discrepancy between experiment and theory has generated considerable
discussion. For example, in 1976 Galletly [9.161] not only commented on the problem, but
carried out some experiments on machined metal spherical capes (with integral edge rings)
having A = 4, to clarify the apparent absence of the trough. His preliminary experiments
also yielded buckling pressures higher than the theoretical predictions, though not as high
as those of Sunakawa and Ichida. One may note in Figure 9.105 that the tests of Krenzke
and Kiernan on similar metal machined spherical caps [9.151] yielded results that clearly
confirm the predicted trough, and also that in Figure 9.106 Tillman's results [9.125] agree

NONLINEAR SYMMETRIC
BUCKLING THEORY
&r,f:l -------..._- ~<JN_ UN--E-AR_ ASY-~MET-RI~
+UCJT x ~CKUNG THEORY
"t +++,..+ xxl-X.~-: ~ ,._ Jo:. ---·~ -----

+ + + ..fX +x n )I.

""
X
X

Figure 9.110 Experimental buckling pressure for Sunakawa and Ichida's clamped spherical caps under
external pressure versus geometrical parameter 'A. (from [9.127])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
746 Shell Buckling Experiments

quite well with the theoretical symmetric buckling curve. Since Sunakawa and Ichida's ex-
periments as well as those of Krenzke and Kiernan and of Tillman were apparently carefully
executed, there was a dilemma. More recently, in 1984 Yamamoto and Kokubo [9.322]
showed that nonlinear calculations, which assume a small inward boundary translation, can
eliminate the trough and bring the predictions very close to the experimental results in the
region near A = 4. A possible source for these small inward boundary translations could be
a contraction during aging of the epoxy adhesive. According to Yamamoto and Kokubo, the
assumption of certain small axisymmetric imperfections could also narrow the gap between
experiment and theory. This may be the solution to the dilemma, but it also stresses again
the importance of precise measurement of imperfections and secondary effects such as
boundary translation and rotation in buckling experiments.
Since their tests included many specimens, covering a wide range of geometries (as can
be seen in Figure 9.110), Sunakawa and Ichida could study the influence of various para-
meters and arrive at other general conclusions about the buckling behavior of spherical caps.
They found that the buckling pressure was practically independent of (R It) but was affected
by the semi-apex angle a to some extent, decreasing with increase in a, this decrease being
the main cause of the lower buckling pressure at large A, observed in Figure 9.110. The
shell deformation patterns before, during and after buckling were found to be axisymmetric,
which was also reconfirmed by high-speed photography. The number of deflection waves in
the meridional direction was observed to grow with loading, and it increased with A.
Note that the more recent test setup of Yamada and Yamada shown in Figure 9.47 is
essentially similar to that of Figure 9.109 and to earlier ones (like those of [9.122] and
[9 .126]), though the emphasis in the 1982 experiments was on the postbuckling behavior of
geometrically imperfect spherical caps. As pointed out in Section 9.5, the initial imperfec-
tions and the large deflection behavior were recorded accurately and the buckling and post-
buckling behavior agreed very well with theoretical predictions (see Figure 9.48).

9.10.3 Complete Spherical Shells


For complete spherical shells, however, both the fabrication of specimens and the test rigs
differ significantly. Relatively few buckling experiments have been carried out on complete
spheres, and not very many on hemispheres. After the explanatory tests of Sechler and Bollay
on copper hemispherical shells, mentioned above in this section, the earliest ones were the
experiments on machined aluminum alloy and steel hemispheres, integrally connected to
ring-stiffened circular cylinders, performed at the David Taylor Model Basin of the U.S.
Navy in the early sixties [9.150]-[9.153], [9.323] and [9.324].
As in the shallow spherical caps of [9.151], which appear in Figure 9.105, Krenzke's
hemispherical specimens [9.150] were machined very carefully, probably representing the
upper limit of commercial fabricability. Most of the 26 small shells of this series, made of
6061-T6 and 7075-T6 aluminum alloy, with R = 0.8-0.85 in. and (R/t) = 11-105, were,
however, too thick-walled to buckle elastically. Only 5 of the 7075-T6 ones, with (Rit) =
50-107 were thin enough to buckle at stress levels within the elastic range of the material
and are plotted in Figure 9 .111. The other shells buckled plastically and will be discussed
in Chapter 16. Probably, however, the thicker ones of the 5 nominally elastic specimens
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

were also in the plastic range, on accounll of the increased local bending stresses induced
by the imperfections.
In an effort to achieve higher buckling coefficients, Krenzke attempted even more careful
machining of two larger hemispheres of R =
2 in. with Rlt = 165 to ensure elastic buckling
[9.323]. They were again made of 7075-T6 aluminum alloy and were rough-machined inside
and out before final machining. Then the model was held in place by a pot-type fixture, and
the final inside contour was obtained with a special accurate tool. The final outside spherical
contour was achieved by supporting the inside contour with a mating mandrel and then
generating the outside surface, using a lathe with a ball-turning attachment. Finally the
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Spherical Shells 747

0
0.6 0
0

0.4 ,~

0
~-~-~-

KRENZKE {1962)
--l
Ref. 9.150

0.2 • KRENZKE 1196 3 l Ref. 9. 323 J


I

~- DA~LEY ~6~f9.324 I
0
L __ _ _ __ L_ _ _ _ _ _ _ _ _ _ __L,_ _ _ _ _ _ _ _ _ _ -L~--~A ____ _L_

20 50 100 150 v 300


tR/f)

Figure 9.111 Experimental elastic buckling pressures for David Taylor Model Basin machined hemi-
spheres (data from [9.150], [9.323] and [9.324])

exterior surface of the integral supporting ring-stiffened cylinder was machined. Extremely
fine feeds together with light cuts were used in the later stages of machining of each surface
to minimize residual stresses.
Every effort was made to minimize departure from sphericity, variations in thickness and
adverse boundary conditions. Accurate measurements of the wall thickness indeed showed
a maximum variation of 0. 75 percent from the mean, while the maximum measured departure
from sphericity was less than 0.01 percent of the mean radius. With the stiffened cylindrical
portion (with the procedure evolved in [9.325] for cone-cylinder junctures) designed for
conditions of membrane deflection and zero edge rotation, the boundary conditions allowed
the hemisphere to behave as a portion of a complete sphere. Foil strain gages to measure
both meridional and circumferential strains inside and outside were bonded along one vertical
periphery of each specimen (36 gages on one and 44 on the other).
One specimen buckled at 73 percent of the classic buckling pressure and the second at
90 percent (see Figure 9.111 ). The first failed with a five-sided dimple in the spherical portion
away from the boundary, whereas the second, with the higher buckling pressure, had a dimple
in the spherical part immediately adjacent to the boundary. Collapse was sudden and no
significant nonlinear strain behavior was observed in either test prior to buckling, though in
the first shell the local inward dimple was not too far from the peripheral line of strain gages.
One may note here that since the dimple is local, one must either be lucky to have located
strain gages at the position of the buckle, or bond an excessive number of gages to increase
the probability of on-the-spot gages. Alternatively, one may use initial imperfection scans to
improve one's chances. However, in the case of near-perfect shells there will usually be only
very minute strain nonlinearities prior to the sudden buckling failure.
Two years later Dadley [9.324] tested six slightly larger high-yield steel hemispherical
shells, five with R '= 3~4 in. and (R/t) = 27~145 and one with R '= 2 in. and (R/t) = 304.
They were machined with the usual DTMB care, except the small one, which was again
machined by the special method of Krenzke [9.323] just described, except that it was thinner.
Only one of the larger models, with R/t = 145, and the smaller, specially accurate one were
thin enough to buckle elastically. Their results are therefore also plotted in Figure 9.111, the
larger one yielded an experimental collapse pressure of 54 percent of the classical one and
the specially accurate shell buckled at 74 percent of the classical buckling pressure, a similar
result to that of the two specially accurate aluminum alloy models of Krenzke.
Though the DTMB specially accurate models represented nearly perfect hemispherical
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`

shells, their (R/ t) values were rather small because they were mainly aimed at elasto-plastic
buckling behavior. Furthermore, the boundary of the hemisphere, even in these carefully
designed specimens, remained a source of uncertainty, especially if the buckles initiated near
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
748 Shell Buckling Experiments

the edge ring. Therefore, the search for high-quality specimens, and in particular complete
spherical shells, continued.
In the mid-sixties Sabir at University College Cardiff [9.326] tested copper and aluminum-
alloy hemispherical shells with R = 3 in. and (R/t) = 150-200, but because of geometric
imperfections and inelastic effects obtained rather low buckling pressures. He then proceeded
to prepare more accurate specimens by etching selected circular areas of the copper hemi-
spheres to a thickness of less than a third the original one. The specimens obtained were in
effect elastically supported spherical caps of (R It) ~ 500 and had thickness variations of
less than 2 percent. Thirty of these specimens were tested with external pressure applied to
the complete hemisphere, and all buckled in the thinned segments at 50-87 percent of the
classical pressure (most shells buckled at about 60 percent Pel and three just below 0.9 Pe 1).
Since, however, the specimens were essentially shallow caps with some uncertainties at the
boundary, they could not be considered a reliable simulation of complete spherical shells.
As mentioned in Section 9.3, in 1960 Thompson [9.85] introduced electroforming (or
electrodeposition) as a method for making high-quality spherical copper shells, and the tech-
nique was developed in the sixties to nickel models and extensively applied to cylindrical,
conical and spherical shells by many investigators. As a result of the work at Stanford
University, electroforming became one of the main methods for producing high-quality thin
specimens of complete spherical shells.
The Stanford spherical shells [9.84], [9.89] and [9.90] were fabricated by electroforming
nickel on a wax mandrel. The process consisted of three steps: (1) the manufacture of the
spherical wax mandrel, (2) the plating and (3) the removal of the mandrel from the shell
specimen. Since the electroplating requires an electrically conducting mandrel for a cathode,
a special conducting wax was used. A shell casting procedure was developed to reduce the
amount of wax required, which then also decreased some of the solidification problems and
simplified the subsequent removal of the wax from the electroformed shell specimen. The
wax was placed inside two hemispherical molds, which were bolted together along the
flanges. This assembly was then mounted on a rotating suspension device, providing contin-
uous rotation about two axes, which was positioned inside an oven. The wax was heated in
the mold to the liquid state. Then the oven was turned off and the rotating mechanism turned
on. As the rotating mold cooled, the wax solidified as a thick shell casting of uniform
thickness. To give the surface of the mandrel the desired degree of smoothness, a shaving
operation was performed with a hand tool, followed by rubbing of a solvent over the surface
and polishing it.
A propriety nickel sulfamate solution containing nickel bromide was used for the plating.
The bath included also boric acid, anti-pitting agent and naphthalene 1-3-6 trisulfonic acid
sodium salt to reduce the internal stresses. Temperature, current density and the pH of the
bath were monitored and controlled throughout the plating operation, and impurities were
minimized by additions of distilled water and continuous filtration, since all these influence
the mechanical properties of the shell. The thickness distribution of the specimen also de-
pends on the suspension method of the mandrel in the bath. Using the same twin-axis rotating
suspension device as in the casting process of the mandrel yielded the most uniform shell
thickness with variations of about ± 5 percent. It was found, however, in preliminary tests
that the abrupt changes in thickness at the electrode holes were sources of weakness, which
demanded changes in the production process. Two different approaches were adopted. First,
the shell thickness was gradually increased as the diametrically opposite electrode holes were
approached. Secondly, regions of slightly reduced thickness were produced by chemical
polishing in the neighborhood of the apex (and within a cone angle of about 100°) of a
hemisphere whose base circle passes through the two electrode holes. (For testing, one of
these holes was later sealed and a pressure fitting was added to the other.) Thus, if the overall
variation in thickness was small, an elastically supported thinner deep cap was formed whose
edge conditions approximated those of a cap of a uniform complete sphere. As pointed out
in Section 9.3 and shown in Figure 9.20, the electroforming process for spherical shells
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

required
Copyright Wiley
a considerable development effort, but
Provided by IHS Markit under license with WILEY
the skill and patience of the Stanford team
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Spherical Shells 749

eventually produced complete spherical shells of (Rit) = 1570-2120, which buckled at 72-
86 percent of the classical buckling pressure (note that half of the last 10 specimens of
Figure 9.20 carried 82-86 percent of the buckling pressure of perfect shells).
When the plated mandrel was removed from the slightly heated plating bath (electroplating
occurs at a slightly elevated temperature), the wax mandrel contracted more upon cooling
than the nickel shell and a small gap was formed (see Figure 9.112). The resulting built-in
deflection limitation permitted the study of the development of the buckle wave pattern under
forced ideal conditions and allowed the specimen to be air tested (by partial evacuation of
the air inside it) without being destroyed. With the mandrel inside, it was found that for
shells that buckled at low pressures (less than 50 percent the classical one) a single dimple
appeared at buckling. It was approximately circular, with a radius of about one-tenth that of
the sphere, or less, and occurred at a flaw. If the test was continued, additional dimples
popped in at higher pressures, usually singly. At higher initial buckling pressures, the shells
buckled with the simultaneous formation of many dimples (see Figure 9.113 ). The transition
to this fully buckled state was very fast, and visual detection of a dimpling sequence was
therefore impossible.
One aim of the investigation of the behavior of complete spherical shells was an experi-
mental evaluation of the influence of the rigidity of the testing system on the buckling
pressures, similar to the studies discussed in Subsection 9.2.2. Here, in the rigid or "hard"
system the volume change of the shell is controlled, whereas in the "soft" system the pres-
sure difference acting on the inner and outer walls of the shell is regulated. The rigid system
(Figure 9.114) consisted of a cylindrical pressure vessel filled with water, into which the
spherical shell was placed (also filled with water), as well as the necessary valves and a
pressure transducer. When the water level at the valve (7) and the reservoir (9) were even,
valves (7) and (8) were closed and external pressure was applied by a fine threaded pressure
piston (3) till buckling. The volume of the water remained unchanged during buckling.
In a soft system, after initiation of buckling, the external pressure should remain constant.
The soft system was approximated here by the use of air as a pressure medium (see Figure
9.115). With valves (2) and (6) closed and (3) open, the vacuum pump (4) was turned on
and a large tank reservoir was evacuated. Valve (3) was then closed and the needle valve
(2) slowly opened, which evacuated the test shell gradually.
After the six preliminary tests, all the 26 shells were first tested with the mandrel inside.
They were first air tested (soft system), and sometimes repeatedly air tested, and then water
tested (hard system), still with the mandrel inside. Finally, the wax mandrel was removed

Figure 9.112 Stanford University tests on electroformed nickel complete spherical shells-view of
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley section through specimen with wax mandrel after cooling (from [9.84])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
750 Shell Buckling Experiments

Figure 9.113 Stanford University tests on electrofonned nickel complete spherical shells-air system
test with wax mandrel inside the specimen (from [9.84]. courtesy of Professor R.L.
Carlson)

by melting the wax and draining it through the pressure filling. Any residue left was removed
by submerging the shell in a container wit h solvent and heat ing for a prolo nged period. T he
specimen could then be tested under more real istic conditions, usually in the soft system.
which resu lted in complete collapse of the shel l. Similar tests without the mandrel in the
rigid system sometimes permiued retesting of the specimen. since for the large (R/t) nickel
specimens, with relatively high u,, practically no plastic deformatio n <>Ccurred.
Comparison of the buckling pressures obtained in the hard and soft systems showed that
they were nearly the same, withi n a few percent, except when the specimen was damaged
during preparation for testing in the rigid test system. Each specimen was tested in both
systems. eliminating possible uncertainties due to differences in speci mens. if on ly simi lar
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

but different shells were compared. Hence the earl ier conclusion. pointed out in Subsection
9.2.2, that the rigidity of the test system has no influence on the buckling load is reconfirmed.
In this connection, it may be of interest to mention an experimenta l study on the buckling
of shallow spherical sandwich shells carried o ut by Lin and Popov at the University of
California, Berkeley, in the late sixties 19.3271. Their specimens were eight shal low spherical
sandwich caps, with 0.006 or 0.008 in. thick 5052 aluminum alloy shell facings and an 0.125
or 0.250 in. 5052 aluminum al loy honeycomb core, fabricated by a vacuum and pressure

Figure 9.114 Stanford University tests on complete spherical shells-


•·igid test system (from [9.84]): (I) shell. (2) specimen
connection. (3) threaded pressure piston. (4 pressure
transducer. (5) lead to pressure transducer. (6) top cover.
(7) valve. (8) valve. (9) reservoir

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Spherical Shells 751

~
~r/~~
c
J(-
(i) TO tANK
Figure 9.115 Stanford University tests on complete spherical
RESERVOIR shells-soft test system (from [9.84]): (1) shell, (2)
needle valve, (3) valve, (4) vacuum pump, (5) in-
_/ clined mercury manometer, (6) valve

cold-forming process similar to that developed in the mid-sixties for PVC spherical caps and
described earlier in this section. The test setup consisted of a 26 in. diameter autoclave, a
pressurizing system and a 96-channel digital recorder. During an experiment the sandwich-
shell specimen acts as a bulkhead separating the autoclave into two chambers. The setup
was designed for two loading systems: a hard system, in which bleeding of confined liquid
in the lower chambers (after equal pressure had been applied to upper and lower chamber)
controlled the displacement of the shell, and a soft system, in which the lower chamber was
left open to atmospheric pressure and the specimen was failed by pressurizing the upper
chamber with nitrogen. Of the eight specimens (with shell segment parameters A = 16-18),
seven were tested by the controlled volume (hard) system and one by the constant pressure
(soft) system. The experimental buckling pressure ratio (normalized in turn by the predictions
of four theories) of the shell tested by the soft system was within 6-11 percent (depending
on the theory chosen) of the mean of the three most similar shells tested by the hard system.
This again confirms that the stiffness of the loading system practically does not influence
the buckling pressure. This study is also worth noting for other experimental techniques.
Returning to the Stanford experiments, Berke and Carlson [9.90] extended them to studies
of postbuckling behavior and high-speed photography of the buckling process. They im-
proved the rigid loading system by replacing the pressure vessel of Figure 9.114 with a
plexiglas tank to facilitate the high-speed photography, using distilled water as a working
fluid and applying the volume changes as reductions in the internal volume of the sphere.
This was obtained by withdrawing water from the sphere with a high-precision micrometer
syringe. The loading was therefore applied to the shell by reducing its internal pressure.
The same plexiglas tank was used for the soft loading system, where the tank was closed
and pressurized with air while the specimen was vented to the atmosphere, the pressure
difference being measured with a manometer.
The specimens were again very thin electroformed nickel spherical shells, with R = 4.25
in. and (Rit) = 2125. The variation of the wall thickness was small, less than 5 percent in
the neighborhood of the buckle or inward dimple.
Two types of experiments were conducted in the rigid system. In the first, the wax mandrel
was left in the nickel sphere, as in Figure 9.112. Spherical shells were buckled in this
configuration in order to obtain high-speed motion pictures of the initiation of the buckle.
With the mandrel, the buckling tests were repeatable because the deflection was limited by
the narrow gap, and many small dimples developed, as shown in Figure 9.113. The location
of the first of the many small buckles was determined by analyzing the high-speed film taken
of a large suspected area before a closeup of the first dimple was attempted. However, even
with this preparatory wider high-speed photography, several tests were required to obtain
successful closcups, also because the first buckle did not always occur at exactly the same
location in repeated tests, though the buckling pressure was the same. The high-speed motion
picture frames of the closeup of the first dimple, discussed in [9.90] (see Figure 9.116),
confirmed the fact that buckling of a complete spherical shell initiates as a small circular
inward dimple. The mandrel did not have any effect on the initiation of the first dimple, or
on its growth, until the bottom of the dimple touched the mandrel. Then further inward
motion of the first dimple was restricted and instead a large number of small dimples de-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
752 Shell Buckling Experiments

Figure 9.116 Stanford University tests on complete spherical shells- high-speed movies of the devcl·
opmem of the first dimple. Buckling load at approximately 0.9 P.,. speed of 7000 frames/
sec and a grid of 20 lines/in. (from [9.90])

veloped at other locations (as can also be seen in the last frames of Figure 9.116). Finally.
the patterns observed in the earlier studi es, of many dimples covering the whole shell (as in
Figure 9.113), were arrived at.
T he second type of experiment in the rigid system was performed with the wax mandrel
removed from the shell. For the very thi n speci mens tested, the comrol achieved by the
prescribed volume change practically eliminated any noticeable plastic de fotmation, and
therefore reproducible buckli ng behavior could usually be obtained without the mandrel. The
primary purpose of these experiments was to obtain pressure-volume change data in the
postbuckl ing range. One should note that whereas for axially compressed cylindrical shel ls
or longitudinally compressed plates the end shorteni ng is used as the deformation parameter.
for spherical shells under pressure the volume change or increment is usually employed.
Pressure-volume change curves were obtai ned for loading and unloading, wh ich exh ibited
similar characteristics. As the inward displacement increased. the mode shape of the initial
circu lar dimple changed to an ellipse, then to a triangular shape, then to a four· and five-
sided shape. These transitions from one mode shape to another occurred both for loading
and un loading, but it was most clearly observed du ring unloading in the rigid (hard) load ing
system. The mode changes were apparentl y caused by the strain energy becoming less for
modes with straight sides than for the circular dimple at larger inward displacemems.
Simi lar experiments were also performed in the soft loading system in order to study the
postbuckl ing process under constant pressure. Since no difference in buckling pressure was
expected between the hard and soft system, the spheres in the soft system were loaded to
within I 0-15 percent of the hard system buckl ing pressure. The pressure was then raised a
few percent during the first half of the running time of the high speed camera (at 2000
frames/sec the running time was 2 seconds), and this procedure was repeated till buckling
occurred duri ng one of the load steps. Again the first displacement modes detected were
small ax isymmetric dimples, and then wi th increasing inward displacements progressive tran·
si tions to other modes were observed. even under the dynamic conditions of the soft loading
system. Fi nally, al l the mode shapes and mode transitions were observed on spherical shells
witll large Rlt (;;;2000) and cou ld be modified in different Rlt ranges.
Similar very thin complete spherical shells (wi th Rlr = 450- 1700), made by electro-
deposition of copper on a wax mandrel, were later tested at Oxford University [9.98] and
yielded similar 2-3-4·5 lobed postbuckling patterns. --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Spherical Shells 753

9.10.4 Large Fabricated Spherical Shells


In parallel with the near-perfect small specimens like the Stanford electroformed nickel
shells, larger fabricated spherical specimens were tested to simulate industrial uses of large
complete spherical shells.
One example was the Moss Rosenberg liquid natural gas (LNG) ships with spherical
independent tanks, which were developed in Norway in the seventies [9.128] and [9.239].
The tanks were large unstiffened complete spherical shells supported at their equator by
stiffened cylindrical shells, four to six such tanks being installed in each ship. The spheres
were thin-walled, with (R/ t) ratios from 350 (for aluminium shells) to 900 (for 9 percent
nickel steel shells), and their thickness was mostly determined by buckling criteria. Since
the available design requirements and test data were not adequate for all the loading con-
ditions, an extensive investigation on spherical shell buckling was initiated. Theoretical stud-
ies and tests on small-scale PVC shell segments were first carried out by Pedersen and Jensen
at the Technical University of Denmark, Lyngby [9.328] and [9.128], and then in 1976-77
a series of 10 larger aluminium alloy spherical shell segments were tested at Det norske
Veritas near Oslo [9.182], [9.183], [9.329] and [9.330].
The main purpose of the Lyngby small-scale experiments was a correlation of experimental
buckling loads with those predicted by an initial postbuckling analysis. The two most im-
portant axisymmetric loading conditions for the lower hemisphere are shown in Figure 9.117.
In empty tanks a biaxial stress field with equally large compressive membrane stresses can
be introduced by a resulting external pressure p "' due to a difference between the internal
pressure in the tank and the external one to which the tank is simultaneously subjected
(Figure 9.117a). Compressive membrane stresses that can cause buckling can also arise as
a result of internal fluid pressure exerted by the liquified gas when the tanks are only partially
filled. The regions in the lower hemisphere close to and above the surface of the liquid gas
are then in biaxial stress fields with a compressive circumferential membrane stress a 0 and
a tensile meridional membrane stress a,1, (see Figure 9.117b). These two stress fields and
their combinations were obtained in the test setup for truncated hemispheres loaded with an
external pressure Pe and ring load Fat the lower boundary (see Figure 9.117c).
The test hemispheres were made from polyvinylchloride (PVC) plate by an air vacuum
framing process at elevated temperature, similar to the ones described in Section 9.3. The
specimens had a radius of 150 mm and (Rit) ~ 500. The 25 test specimens had a rotationally
symmetric thickness variation introduced by the thermoforming operation, which could be
represented by a fourth-order polynomial. Any shell whose thickness variation deviated more
than 6 percent from this polynomial representation was discarded. Extensometer measure-

+
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(OJ

(C)

Figure 9.117 The two independent loading systems for the LNG spherical tanks (from [9.128]): (a)
external vapor pressure, (b) partial filling, (c) model tested
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
754 Shell Buckling Experiments

ments of the elastic modulus showed that it varied with the variation in thickness, from
2.90 X 109 N/m2 near the equator to 3.5 X 109 N/m 2 near the pole, but there was only
negligible orthotropy. It could therefore also be represented as a simple function of the
thickness variation. Poisson's ratio was obtained from strain gage measurements as v = 0.37.
The experimental setup is shown in Figure 9.118. The test specimens were clamped in a
jig that held the upper support ring. At the lower end of the test specimens three different
boundary conditions were employed. First the hemispheres were exposed to an external
pressure and a nearly concentrated force at the pole, as shown in Figure 9.118a. After
completion of tests with the point load, the hemispheres were truncated to spherical segments
clamped at </> = 77.5°, with the boundary support of Figure 9 .118b. After completion of tests
with this setup, the shell segments were further truncated and the boundary supp011 of Figure
9.118c was applied, clamping the segments at </> = 45°.
As shown in Figure 9.118a, the uniform external pressure Pe was applied as a static water
pressure exerted by the water columns in two tubes of small cross-sectional areas to the left
of the test jig. Using a needle valve, a suitable slow flow through one of the tubes from a
water reservoir was obtained and thereby a slow pressure increase. Two magnetic valves
made it possible to shut off this water flow the moment the shell buckled. Manometer
readings from the second tube continuously recorded the water pressure. The inside of the

J:
WATER SUPPLY

EEDLE VALVE
AGNETIC VALVE F
~(DEAD WEIGHT)
CLAMPED SUPPORT

...
..."'
...~:z

MAGNETIC
VALVE
(a)

(b)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`

(c)
Figure 9.118 Experimental setup for the Lyngby tests of hemispheres and spherical segments (from
[9.128]): (a) schematic representation with point load, (b) lower 12S boundary support,
(c) lower 45° boundary support
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Spherical Shells 755

test shell was also filled with water in order to obtain the same external pressure at all levels
of the surface of the specimen. The load at the lower boundary of the specimen consisted
of a contribution from the water pressure and from an external force F that counteracted the
water pressure. The force F was kept constant during each test and was applied with weights
via an almost frictionless lever system.
Since there was only water above the shell, buckling could be easily observed and pho-
tographed. The buckles always appeared quite suddenly, and because of the small cross-
section of the two tubes, they were followed by a sudden drop in the manometer reading.
This immediate pressure drop after buckling prevented damage to the specimens. Thus some
of the tests were repeated with the same results.
The correlation between the experimentally obtained results and the theoretically deter-
mined imperfection sensitivities for spherical shell segments was not based on direct mea-
surements of initial geometric imperfections. Instead, imperfection amplitudes were chosen
for three cases such that the experimentally obtained reduction factors were predicted for
one particular loading condition in each case. Then it was shown that the variation of the
experimentally observed buckling loads for various combinations of two external loadings
agreed well with the variation of the theoretically predicted loads for these load combina-
tions, based on the chosen imperfection amplitudes. Fairly good general correlation was
found in this manner between experimental buckling loads and theoretical predictions based
on an initial postbuckling analysis, and this gave more confidence to the analysis and design
methods proposed in [9.328].
The buckling behavior of similar partly filled spherical storage tanks under vertical and
lateral acceleration was later also studied experimentally with thin cast epoxy models (of
Rlt ~ 900) at Oxford University, already discussed in Section 9.3 (see [9.137]).
However, it was concluded that experiments on models with measured imperfections were
needed for more reliable design methods, which motivated the Det norske Veritas tests on
spherical shell segments on a scale nearly one order of magnitude larger than the Lyngby
hemispheres, though still one order of magnitude below the actual LNG tanks.
As mentioned in Section 9.3, the models were welded aluminum alloy truncated spherical
shells with R = l m and (Rit) = 400 and 500 (see Figure 9.25). Each model contained
three meridional welds, whereas the full-scale LNG spherical was assembled from a large
number of segments by circumferential and meridional welds. The full-scale welding pro-
cedures and also, therefore, the weld-induced residual stresses and displacements were dif-
ferent. Hence the results obtained on the 10 models tested were not directly applicable to
the full-scale spherical tanks but served primarily for general correlation between measured
imperfections and the buckling predictions for imperfection sensitive shells.
The experimental setup employed is shown in Figure 9 .ll9a. The stiff steel rings (1) and
(2) were designed to transform the concentrated loads applied by four hydraulic jacks (3)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

into evenly distributed edge loads acting on the shell segment (4). The machined bottom and
top rings (6) and (7) provided clamped boundary conditions for the specimen. They also
served as the reference surface for the measurements of initial imperfections and radial
deflections. The cylinder cover (5) and the rubber profile (8) served as the pressure chamber,
which was filled with pressurized water to apply external pressure on the model. The pressure
was measured by a manometer at the top (9).
The rigidity of the loading system was analyzed by a finite element model. However,
before the first model was tested, the vertical deflections were measured, without the cylin-
drical pressure chamber in place, at many positions inside the shell, in order to check the
rigidity of the test rig. No significant bending or twisting of the steel rings was observed.
In addition to the ultrasonic thickness measurement mentioned in Section 9.3, the initial
geometrical imperfections of the shell segments were mapped by a special device shown in
Figure 9.119b. The machined end rings, (6) and (7), served as rails for a carriage (10), which
was equipped with an angle counter (11) and seven inductive displacement transducers (12).
The transducers could be moved vertically and were calibrated prior to the test of each
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
756 Shell Buckling Experiments

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Figure 9.119
(a) = (b)

Experimental setup for the Det norske Veritas tests on large spherical shell segments
(from [9.182] and [9.329], courtesy of Dr. J. Odland): (a) the test rig with pressure
chamber, (b) the initial geometry mapping system

model. Before loading, measurements were taken at 21 meridional posttwns, but during
testing the transducers were fixed in certain vertical positions and the carriage was moved
manually circumferentially. The circumferential position, recorded by the angle counter ( 11 ),
was stored on tape together with the data from the displacement transducers, and therefore
constant circumferential speed was not required. If the transducer carriage was moved too
fast, however, vibrations occurred that disturbed the measurements. The radial deviation (the
imperfection) was measured within ± 0.1 mm accuracy ( ± 5 percent of the equator shell
thickness t) and related to the mean radius, assumed to be the radius of a corresponding
ideal sphere. A computer program was developed for the analysis of the data, and the de-
viation from the mean radius was plotted on line on a screen, from which hard copies could
be obtained. Both initial imperfections and their growth were scanned. As will be discussed
in Chapter 10, this system was one of the first industrial systems, but it is typical of modern
imperfection measurement systems.
With the hydraulic jacks and the pressure chamber, the models could be loaded by any
combination of axial force P (compression or tension) and external pressure p. At the equator
(where the meridional angle ¢ = 90°) the membrane stress resultants are:
Nq, = -(pR/2) cos 2 ¢ 0 - P/2nR (9.22)
and
(9.23)
where ¢ 0 is the meridional angle of the spherical shell segment. It was always possible to
choose p and P in such a way that any desired ratio (Nq,l N,) could be obtained. Three
loading conditions were defined for the testing:

LCl: (N,b!N 11 ) : (u1,/u0 ) :


LC2. (Nm/ N 0) - ( u,j u 0 ) -
-1}0 (9.24)
LC3: (Nm!N 11 ) = (umlu0 ) = l
Computing the prebuckling stresses with the BOSOR 4 program [2.92] showed that the
stresses away from the edges could be calculated with sufficient accuracy from the membrane

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Spherical Shells 757

solution, Eqs. (9.22) and (9.23). The three loading conditions defined by Eq. (9.24) could
therefore be obtained as follows:
LCl: p = 0, P =I= 0 only axial compression or tension }
LC2: P = -pR 2 1T cos 2 cp0 (9.25)
LC3: P = pR 2 1r(1 - cos 2 cp) uniform external pressure
For a chosen loading condition the load was increased in steps. At each step the following
were measured: the water pressure, the jack force, continuous radial deflections along the
circumference, vertical deflections at four circumferential positions, and strain gage mea-
surements at four positions on the equator. Here the strains were measured for additional
control of the loading condition and the load level, and not for nonlinear buckling stresses.
Usually the model was loaded until buckling occurred in one loading condition and then
was reloaded in a different loading condition until buckling.
When the critical load was reached, the specimen always buckled locally at the location
where the geometrical imperfection was most severe. As the model was then unloaded, the
permanent deflections were generally found to be small. They were faired and the model
was backed up with carefully shaped wooden blocks, leaving only a small gap between them
and the shell (see Figure 9.120). Then the specimen could be reloaded under the same or
another loading condition. The stress state at the equator was always checked, and it was
usually found that the model could be loaded until buckling had occurred at three different
locations before the stress state in the remainder of the shell was disturbed. This was the
case because the specimens always buckled first close to the three meridional weld seams
where the geometric imperfections were most severe.
One of the models was also loaded by a point load (applied with a small hydraulic jack)
acting on the equator towards the center of the sphere.
The correlation between theory and experiments consisted of comparisons with relations
between predicted relative buckling loads and imperfection magnitude. The predictions were
obtained with initial postbuckling theories (like those of [2.54] and [2.57]) and were based
on the assumption that the shape of the initial imperfection was that of the natural buckling
mode, while its magnitude was equal to the amplitude of the measured regular imperfection
pattern. Since, however, the measured imperfections were very irregular (as can be seen in
the Appendix of [9.329]), an equivalent imperfection amplitude had to be defined. Such an
equivalent amplitude should be fairly easy to acquire not only for test specimens, which
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

WOODEN BLOCKS

-A

Figure 9.120 Del norske Veritas tests on spherical shell segments-


arrangement for backup of buckled regions (from
[9.329] and [9.182], courtesy of Dr. J. Odland)

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
758 Shell Buckling Experiments

could be carefully measured in the laboratory, but also in situ for full-scale tanks installed
in a ship. After several alternatives were evaluated (see [9.329]), it was finally decided to
define the imperfection amplitude as the maxi~um local out-of-roundness measured radially
from a circular template of a specified length L and a radius equal to the nominal radius R
of the sphere. One of the reasons for the choice of this alternative was the fact that buckling
in spherical shells always occurred locally and at the location of the maximum local out-of-
roundness, apparently entirely independently of the imperfections in other remote parts of
the shell.
The appropriate length of the template had still to be determined. Usually the dominating
measured imperfections were more long-waved than the natural buckling mode of the shell.
Therefore, if the template were longer than the natural half-wavelength, the measured equiv-
alent imperfection amplitude would be larger than for the case that the template length was
equal to the natural half-wavelength. One would expect imperfections having the same shape
as the natural buckling pattern to be the most severe; but more long-waved imperfections
with larger amplitude could have the same severity. Hence the appropriate length of the
template was taken to be somewhat larger than the natural half-wavelength and defined (after
a close examination of the predicted eigenvalues) for the different loading conditions as:
L= 2Vift for LCl and LC2
(9.26)
and I= 4Vift for LC3
With these specified lengths the equivalent imperfection amplitudes could be derived from
the recorded imperfection plots.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

The correlation between the experimental relative buckling loads (sometimes called the
imperfection factor or knock-down factor) versus the equivalent imperfection amplitudes and
the corresponding theoretical predictions for the Det norske Veritas tests are shown in Figure
9.121. For LC3, (crq,lu0 ) = I, corresponding to a sphere under uniform external pressure,
the relative buckling loads p = (Pma) Pc~) were in the range of 0.25-0.45, values that agree
well with similar fabricated cylindrical shells under axial compression.

1.0 0
0
0
0.9 0
0
0

0.8
0~ 0
~
..... 0.7
~ o oo~
"E
b 0
" 0.6
.,_
0::
e o.5
0
~
z 0.4
0
;::
0
...w
:!'i 0.3
a.
!
0.2
LCI OJ
LC2 0 EXPERIMENTAL VALUES
0.1
LC3'7
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
EQUIVALENT IMPERFECTION A'*'LITUDE Sl'
Figure 9.121 Det norske Veritas test results for spherical shell segments subjected to three different
loading conditions, plotted versus the equivalent imperfection amplitude (from [9.329])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Spherical Shells 759

For LC 1, ( crd,! cr0) = - 1, corresponding to a spherical segment under axial tension, the
relative buckling loads found p = ( PmaJ Pc~) = 0.80- LOO. These values were generally
higher than those reported earlier by Yao in tests on truncated hemispheres under axial
tension [9.331]. Since Yao's models were either shorter (smaller meridional angle), and
therefore more imperfection sensitive (according to [2.57]), or with a much higher (R! t) ratio
and therefore also with larger relative imperfection amplitudes, the notable difference in p
seems reasonable. Furthermore, it is likely that if the initial imperfections had been measured
in these 1967 tests, they would have been found to exceed those of the Det norske Veritas
specimens.
For LC2, (cr, 1,! cr0) = 0, the relative buckling loads were in the range of 0.55-0.80. Ap-
parently no other comparable experiments have been carried out for loading condition LC2.
The correlation with the theoretical predictions in Figure 9.121 shows that the traditional
way of design with an imperfection factor is rather crude and conservative. As will be
discussed in detail in Chapter 10, more efficient methods are available today, but prior to
their general application, extensive data collection and evaluation of measured imperfection
data, like that initiated at the Delft-Technion Data Bank (see [10.13]. [10.20]) is necessary.
The experiments here showed clearly that the elastic buckling of imperfect spherical shells
is governed not only by the maximum compressive principal stress but also by the principal
stress ratio.
Another example of industrial uses of large thin-walled spherical shells is the spherical
steel containment shells considered recently for nuclear reactor designs at the Kemforschung-
szentrum Karlsruhe, Germany (see for example [9.332]). The design was supported by ex-
tensive theoretical and experimental studies of the dynamic behavior of spherical shells under
seismic loading, which excites beam-type oscillations, causing significant membrane stresses
that may lead to buckling phenomena.
The experiments consisted of a static buckling test on a horizontally loaded spherical shell
and dynamic buckling tests of a larger high-precision spherical shell model [9.333] and
[9.336]. The test rig for the static buckling test is shown in Figure 9.122. The specimen was
a steel shell, of (R It) '= 350. It was fabricated from 14 segments and a pole cap, cut from
a sheet of mild steel preformed by deep drawing and precisely welded together. For good
sphericity the shell was pressurized against a precise spherical mold so that imperfections
were suppressed by plastic deformations. The maximum shape imperfection was therefore
only 0.1 percent of the diameter. The shell was truncated at the bottom and bonded there to
a stiff base flange. The shell was loaded horizontally by a rod pulling a stiff pole cap, which
was also bonded to the shell. Thus, static stress distributions were obtained in the lower shell
region that were similar to the dynamic stress distributions caused by the seismically excited
beam type oscillations. Buckling phenomena similar to the dynamic case were therefore
expected.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

horizontal load

pole displacement

Figure 9.122 Karlsruhe test rig for static horizontal loading of a model of a spherical containment
shell (from [9.333], courtesy of Dr. S. RafT)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
761J Shell Buckling Experiments

As the load was increased, a growing meridional buckle near the clamped bouom end of
the shell was observed, which developed in the postbuckling region into an "elephant foot"-
type buckle (see Figures 9. 123 and 9.124). Since the experiment was displacement controlled.
the unstable postbuckl ing equilibrium path with decreasing load cou ld be followed, till the
test ended suddenly by a failure of the bond between the shell and its base flange. The final
meridional buckle had an amplitude of about 5 1, but opposi te to it shallow circumferential
buckles, with amplitudes 0.2 t, were revealed (as shown in Figure 9. 123).
For the dynamic buckling tests a larger spherical steel shell, about 1.3 m in diameter and
with (R!t) "" 650, was manufact ured wi th high precision by cutting it from a much thicker
shell on a lathe (see [9.334] and [9.335)). After different fabricat ion processes were evalu-
ated, machi ning from a thicker shell was chosen in order to obtain the desired precision.
The blank was made from two 38-mm thick hot-pressed semispherical 15 MnNi63 steel
shells welded at the equator. The model was truncated at <1> = 130° and the base llange
welded to it. The machi ning process had four stages. shown schematically in Figure 9.1 26.
To perm it accurate machining of the thin shell. a two-part cast support housing (or female
mandrel) was employed. After it was accurately machi ned, the specimen (Figure 9. 125) had
deviations from the ideal spherical shape of less than 0.05 percent and thickness variations
of less than 3 percent. It was mounted on a stiff base flange. which was elastically supported
and coupled by a rod to the vibrati ng mass of a horizomal electrodynamic shaker. In the
tests, the excitation intensity was increased stepwise. horizontal and vertical accelerations
were measured by 12 transducers. strains were measured by II strain gages in the lower
portion of the shell and the horizontal displacement was recorded at one location on the
equawr. Four tests were performed on the model. Since the measured time histories for tests
3 and 4 showed qual itative changes in comparison to tests I and 2. it was concluded that
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

~~
-· . "./ 1 -~.~

lfflL!'\ ~\\\
--\t\·t· · t--i~-J~J-· ·
'W ~_: uJI
m~ddionol ! circumlenmtiol
buckle buckles

Figure 9.123 Karlsruhe spherical containment shell lCSlS- lwo different buckling modes in !he static
lest (from [9.333). counesy of Dr. S. Rafl)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Spherical Shells 761

figure 9.124 Karlsruhe sphcoi cal containment shell tests- permanent meridional buckle of .. elephant-
foot .. type (fro m ]9.333 ]. courtesy of Dr. S . Rafl)

buckling must have occurred when the excitation was increased from test 2 to test 3. T he
visual inspection of the shell after tests 3 and 4 indeed revealed a permanent meridional
buckle close to the clamped bottom edge and o pposite the excitation rod, si milar to that
observed in the static test (see Figu re 9. I27). No buckles were found at the shell side near
the excitation. The colo ring of the buckle area indi cated that th is region of the shell must
have reached temperatures of about 300°C in the test with the highest excitation intensity.
This led to the conclusion that the shell must have huck led periodically, causi ng significant
plastic strains.

9.10.5 Spherical Shells Subjected to Concentrated Loads


Bel(lre leaving the topic of spherical shells, it should be mentioned that extensive theoretical
and experimental st udies were also carried o ut for the case of spherical shell s and caps
subjected to a central concentrated load (sec for example ]9.224), ]9.143 ), )9. 144], [9.313).
[9.326] and [9.335]-[9.337]). Ashwell 's experiments in the late fifties [9.335] were appar-
ently the earliest reported and agreed well wi th his analysis. Good agreement between theory
and experiment also characterized the tests of the sixties. S ince the experimental layouts fo r
central concentrated loading do no t di ffer significant ly from those for external pressure. they

Figure 9.125 Karlsnohe spherical contai nment shell 1es1s-high-precision model for dynamic leSIS
(from [9.333). counesy of Dr. S. Rafl)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
762 Shell Buckling Experiments

(a) (b) (c) (d)

Figure 9.126 Karlsruhe spherical containment shell tests~fabrication process of high-precision model
for dynamic tests (from [9.334] and [9.335], courtesy of Dr. S. Raff): (a) outside of
spherical model is machined (turned), (b) a two-part support light alloy housing is cast
and its internal spherical contour is machined (turned), (c) the two parts of the housing
are first heated, then the model is inserted and fixed in the shrinking housing; then the
remaining space between model and housing is evacuated to assure complete support,
(d) the model can then be machined internally to the final thickness

will not be discussed here. However, there are similarities in the buckling behavior and its
dependence on the shell segment geometrical parameter A.

9.11 Toroidal Shells, Torispherical Shells, Buckling under


Internal Pressure
9. 11. 1 Toroidal Shells
Complete toroidal shells (shells of revolution with a crown, also called sometimes torus-type
shells, see Figure 9.128a) are employed, primarily as fluid tanks, in various branches of
engineering. In particular, they were often used in the aerospace industry in the late fifties
and sixties and considered for underwater structures in the seventies. The most common
loading that may cause buckling of closed toroidal shells is external pressure, which has
indeed been the subject of many theoretical and some experimental studies.
The earliest analyses of bifurcation buckling of a toroidal shell subjected to uniform ex-
ternal pressure were carried out by Machnig in the late fifties [9.340] and [9.341] and pre-
dicted buckling in an axisymmetric mode. A more complete analysis was presented by Sobel
and Fliigge in 1967 [9.342], and many others followed (see for example [9.343]-[9.350]),
some of these employing nonlinear equations and considering prebuckling deflections.

I
e~ 1so·
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 9.127 Karlsruhe spherical containment shell tests~position and size of buckle observed after
Copyright Wiley dynamic tests of high-precision model (from [9.333], courtesy of Dr. S. Raff)
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Toroidal Shells, Torispherical Shells, Buckling under Internal Pressure 763

(a ) TOROIDAL SHELL

SPHERICAL CAP

(b) TORISPHERICAL SHELL

Figure 9.128 Geometry and notation of toroidal and torispherical shells: (a) toroidal shell, usually C
and D are called equators and E and F poles or crowns, (b) torispherical shells as dome
of a cylindrical pressure vessel

Two types of buckling modes were shown to be possible, one symmetric about the equa-
torial line (CD in Figure 9.128a), called mode A (Figure 9.129a), and one antisymmetric
about it, mode B (see Figures 9.129b and c). Both types of mode can be either axisymmetric
(n = 0) or asymmetric (n > 0). The critical pressures corresponding to modes A and B are
always close to each other. For the axisymmetric buckling mode (n = 0), mode B yields
lower buckling pressures than mode A, whereas for asymmetric buckling modes (n > 0)
mode A results usually in slightly lower pressures. Calculations show that nonlinear prebuck-
ling effects are insignificant and that toroidal shells under external pressure have a stable
postbuckling behavior and are therefore imperfection insensitive.
Only a few experimental investigations on the stability of closed toroidal shells have been
performed. The earliest tests were apparently those carried out at the Lockheed Missiles and
Space Company in the mid-sixties and reported by Sobel and Fliigge [9.342], who compared
their predictions with the experimental buckling pressures and found the agreement to be
within 10 percent. The tests included three almost complete steel toroidal shells, with a torus
radius r = 2.825 in., (rl t) ~ 70-80 and a slenderness ratio (blr) = 8.04, as well as one
titanium 180° toroidal shell with r = 3.5 in., (rl t) = 70 and (b/ r) = 6.32. Under the uncertain
assumption that the titanium half-torus was simply supported at its edges, its experimental
result was compared in [9.342] with the theoretical predictions for a complete torus buckling
into two circumferential waves.
As was pointed out by Sobel and Fliigge, these toroidal shells were rather slender, and it
was not certain whether the good correlation obtained between test and theory could also be
maintained for shells with small (bl r). Another series of experiments was therefore carried
out at Lockheed Palo Alto on "stouter" toroidal shells with (blr) = 2 [9.351]. Except for
the slenderness ratio, the dimensions of the shells were similar, with a torus radius r = 2.5
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

in.
Copyright Wiley
and (r/ t) = 50. To obtain a uniform thickness
Provided by IHS Markit under license with WILEY
distribution, the specimens were manu-
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
764 Shell Buckling Experiments

(a) MODE A, n• 2

(b) MODE B, n•2

Figure 9.129 Buckling mode shapes predicted for a toroidal shell with
(r/ t) = 100 and (hi r) = 4 (from [9.342]): (a) asymmetric mode
A with n = 2, symmetric about equatorial line, (b) asymmetric
mode B with n = 2, antisymmetric about equatorial line, (c)
axisymmetric mode B with n = 0, antisymmetric about equa-
(c) MODEB, n•O torial line

factured by casting an epoxy resin. The toroidal shells were cast in two halves, which were
then glued together by a room-temperature adhesive selected to match the material properties
of the shell, and good shape accuracy was obtained. To obtain accurate critical pressures, an
epoxy resin with minimum room-temperature creep was chosen. This selection yielded a
specimen material with a high modulus (E = 500,000 psi) but also brittle. Hence at buckling
the shells shattered and one could not observe their buckling and initial postbuckling be-
havior. One additional shell was therefore made from a very ductile epoxy material with a
low modulus for study of the development of the buckling pattern.
Three shells of the brittle material were tested, the external pressure being applied by
partial evacuation of the air in the shell. The experimental buckling pressures were within a
few percent of the theoretical predictions. In the test of the ductile shell that followed (see
Figure 9.130) an axisymmetric buckling pattern was first observed, as predicted theoretically,
and as the load was further increased a nonsymmetrical buckling pattern appeared.
In the late sixties and early seventies three series of experiments on toroidal shells sub-
jected to external pressure were carried out by Fedosov in the U.S.S.R. In the first series
[9.352], nine aluminum alloy toroidal shells were fabricated by spinning symmetric halves
of the torus sheet material in stages, with annealing between the stages, machining the ends
on special mandrels and welding the halves together. The specimens had a torus radius r =
80 mm, (rlt) = 100 and slenderness ratios 2.5, 3.325 and 4.0. The second series [9.345]
included eight stainless steel toroidal shells of the same dimensions as those of the first
series. The specimens were also made by a similar process, the axisymmetric torus halves
were stamped in stages from sheet with annealing between the steps, and machining of the
edges prior to being welded together. In the third series [9.348] there were also eight shells
of the same torus radius but with double wall thickness such that (rlt) = 50 and with only
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

two slenderness ratios (bl r) = 2.5 and 4. Four of the specimens were made of stainless steel
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Toroidal Shells, Torispherical Shells, Buckling under Internal Pressure 765

(o) UNLOADED SHEL L

( b) AXISYMMETRI C BUCKLING

lcl NONSYM METRIC BUCKL I NG

Figure 9.130 Lockheed Palo Alto test or a toroidal shell under extemal pressure. The specimen is a
ductile epoxy resin and shows the nonsymmetric postbuckling pattern (rrom [9.35 I].
courtesy or Dr. L.H. Sobel)

and four of alumjnum alloy. All models were agai n fabricated by stamping axi symmet ric
half-torus segments, with standard anneali ng between the stamping stages, machin ing the
edges and jo ining the halves with argon arc butt weld ing. Fairly good specimens were o b-
tained for the three test series, whose measured dimensions deviated by less than 6 percent
from the no minal ones and that did not have significant resid ual welding stresses.
T he shells were tested in a special rei nforced pressu re chamber, supported elastically at
the outer equatorial circumference of the torus and in the later tests a1 boch inner and o ucer
equators. Most specimens were 1es1ed under hydrostacic loadi ng and some under ai r loadi ng.
In all cases buc kling occu1Ted with an abrupl snap-through. Some of the shells buckled in
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
766 Shell Buckling Experiments

an axisymmetric mode and some in an asymmetric one. In both cases the mode shapes were
mostly of type B (antisymmetric about the equatorial line-see Figure 9.129), but some of
type A (symmetric about the equatorial line) were also observed. In all the tests the wave
formation was most pronounced in the neighborhood of the poles. The experimental buckling
pressures were fairly close to the theoretical predictions of [9.352] and [9.348].
Fedosov pointed out [9.348] the influence of an axial force that could arise from the weight
of the shell and its manner of support and distort the deflected shape. The effect of axial
loads on toroidal shells was also analytically and experimentally investigated in the mid-
sixties by Jordan at the Martin Company in the United States [9.353]. Shells subjected to
both axial load and internal pressure were considered, but the tests dealt only with stress
distributions and deflections and not with failure. Hence one model (see Figure 9.131), which
was retested many times, sufficed. It was a medium-sized stout aluminum shell with r =
11.5 in., (rlt) = 183 and slenderness ratio (blr) = 1.37. The torus was made from two
identical half-shells, each fabricated by explosive forming from a preform consisting of a
horizontal flat annular plate and two conical parts welded from flat sheet material. The
manufacturing operation that was most critical for the shape accuracy of the final model was
not the explosive forming (which was done in several steps), but the joining together of the
two half-shells. Therefore, careful machining of the edges in a special fixture preceded the
welding (as was also the case for Fedosov's specimens. discussed above). As can be seen
in Figure 9.131, the model was supported at its rim, or outer equator, by means of 100
T-shaped support lugs fixed to a heavy base cylinder. The lugs were bonded to the specimen
via 1/8 in. thick rubber buffers. The test specimen was instrumented with many strain gages
and mechanical dial gages (which today would obviously be replaced with recording trans-
ducers) as well as with a simple crown shift meter, consisting of a pointer (fixed to a small
block bonded to the shell at its crown) whose angular motion could be measured. The initial
geometric imperfections were measured by the dial gages. From these measurements it was
noted that at the location of the crown shift meter there was a local imperfection whose
straightening out by the internal pressure obscured the rotation of the shell that the shift

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
meter was planned to measure.
In the early seventies additional experiments on toroidal shells subjected to hydrostatic
pressure were carried out related to their use as underwater structures. At the U.S. Naval
Ship Research and Development Center (NSRDC), Washington, D.C. (the former David
Taylor Model Basin), Fishlowitz [9.354] tested eight plastic toroidal shells covering a wide
range of slenderness ratios, (bl r) = 1.20--8.00, and torus radius-to-thickness ratios, (r/ t) =

LEGEND:
\> DEFLECTION GAGE
0 STRAIN GAGE
CROWN SHIFT METER

1-aon--j
f-----------54in

Figure 9.131 Martin Co. model toroidal shell subjected to axial load and internal pressure and its test
arrangement (from [9.353], courtesy of Dr. P.F. Jordan)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Toroidal Shells, Torispherical Shells, Buckling under Internal Pressure 767

25-100, which complemented the earlier Lockheed tests [9.342) and [9.35 1) by focusing o n
the middle and lower range of hi r (see Fig ure 9. 132).
Since the objective was the study of elastic buckling behavior, Fishlow itz looked fo r a
model material with a hi gh ratio of limit of proportionality to Young's modulus. T he material
selected was a plastic formed by the mix ture of equa l parts by weight of Versamid 140
polyamide resin and Epo n 828 resin, with which the NSRDC sho p personnel had consid-
erable experience. For this plastic (IT,/ E) E:!! 0.012 (i .e. nearly tw ice that for aerospace g rade
aluminium al loys) where IT, represents the proportional limit, a ratio hig h enough to ensure
elastic buckling.
The shells were made by casting half-sections to the fi nal o utside dimensions in female
molds and then machining the final inside dimensions as well as the joini ng edges. Wooden
cores were used to reduce the machining necessaty and avoid cracks caused by uneven
temperature distributions in thick castings. The halves were joi ned together with the same
plastic mixture as that employed in casting the shells.
The model fabrication process is presented in detai l in [9.354], where specimen design
considerations as well as other test data and results are fully discussed. It is a meticulous
test report. worth close study by any experimentalist.
The speci mens were carefull y measured before joini ng the halves, indicating less than 7 .5
percent difference between average measured thickness and the design values but significant
nonuni formity in thickness (in some shells larger than 20 percent). T here were, however,
only very small variations in measured r and IJ compared to the design values.
1\vo loading schemes were employed in the tests: load application by pressurizing a tank
containing the model and by partially remov ing the air inside the model wi th a vacuum
pump. The latter type of test allowed visua l and photographic observation during the test,
which facilitated determinatio n of the buckli ng mode shapes . These, however. were actually
better defined by the strain gage records (t he speci mens were instrumented by dozens of
strai n gages, each on their o uter surface). The measured strains at collapse were elastic in
all the models, indicating that they failed by elastic instability. The Southwell method was
used to calcu late the coll apse pressure (for the corresponding perfect shell) from the exper-
imental strai n data. in the manner discussed in Section 9.6.
T he experimental collapse pressures were compared to prediction by Bushnell 's analysis
[9.344) and that of Sobel and FHigge [9.342) and wi th the exception of two shells (discussed
in [9.354) in detai l) the agreement was within ± 10 percent. T he buckling modes were similar

figure 9.132 Deformation of a U.S. Navy NS RDC toroidal plastic shell. no. 6. (from [9.354], courtesy
of Dr. E.G. Fishlowiiz): (a) no load. (b) first visible deformation at 85 percent of collapse
load. (c) at 96 percent of collapse load. (d) at 100 percent of collapse load. just before
collapse, (e) after fai lure. Note the side mirrors that permit beller observation of the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

deformation paucm
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
768 Shell Buckling Experiments

in nearly all the models. Typical experimental buckling and collapse modes are shown in
Figure 9.132 for one specimen, Shell no. 6.

9. 11.2 Torispherica/ Shells under External Pressure


The most widespread use of segments of toroidal shells is as the knuckle of torispherical
shells (see Figure 9.128b), which are the most prevalent type of dished ends for cylindrical
containers. Under external pressure loading, the buckling behavior of torispherical shells or
domes resembles that of spherical caps, and therefore they have often been studied together,
both theoretically and experimentally (sometimes elliptical domes were also included-see
for example [9.122]). In these studies the relevant hemispherical shells have then frequently
been taken as reference points (see for example [9.356] and [9.364]).
In the sixties some experimental studies were carried out on torispherical shells under
external pressure [9.357]-[9.361] in order to provide experimental data for designers on the
behavior of torispherical heads of pressure vessels. In particular, answers were needed on
whether the knuckle region acted as a stiffener and whether the torispherical head buckled
as a part of a sphere. The answers were incorporated in the design formulas, but more
information was required for better understanding of the influence of the geometric para-
meters and for verification of the computer programs developed for design and optimization.
Therefore, research on torispherical heads has continued, primarily that performed by Gal-

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
letly and his group at the Department of Mechanical Engineering of the University of Liv-
erpool. Their program included many experimental investigations (see for example [9.356],
[9.363]-[9.366]) and dealt both with external and internal pressure loading, which will be
discussed later in this section.
The experiments on shallow torispherical domes under external pressure carried out by
Galletly and his co-workers in the mid-eighties [9.364], represent a typical example of a
carefully executed and well-reported modern experimental study in the field and are therefore
discussed in detail. The study included 24 steel shells, and since its focus was on shallow
torispherical shells with sharp knuckle radii, 16 of the specimens had sharp knuckle radii,
yielding knuckle-to-diameter ratios (rl D) = 0.06-0.095, whereas the remaining eight control
shells had a rather generous knuckle ratio (riD)= 0.18.
In order to provide practical design data, the test specimens were chosen to be standard,
as-manufactured pressure vessel heads, complying with the shape limitations imposed by the
British BS 5500 Code, i.e. (rl D) > 0.06 and (RJ D) < 1.0. (The code actually specifies,
instead of r, the best-fit inside knuckle radius for one meridian r,, and instead of R,, the
best-fit inside crown radius for one meridian R,,, but for design of the specimens the nominal
values of r and R, suffice). The dimensions of the test heads were also limited by the
requirement that they should be suitable for testing in the Liverpool hyperbaric chamber
(Figure 9.133), a special test facility of 0.9 m diameter, 3 m length and maximum operating
pressure of 13.8 Nlmm 2 (2000 psi).
This test chamber (Figure 9.133), which has been used in a number of the recent exper-
imental programs at Liverpool University, consists of a thick-walled horizontal cylindrical
vessel having a torispherical head with a 12 in. (305 mm) bore nozzle at one end and a full-
diameter quick-release flanged door at the other end. In the cylindrical section of the chamber
are a number of instrumentation ports. The pressurizing fluid is insulating oil, which is stored
above the chamber when not in use. The test compartment is pressurized by an electrically
driven pump for coarse pressure control arrd by a hand pump for fine control. A bleed valve
is provided for fine control of reduction of pressure (see [9.367] for more details of the test
facility).
The test program was carried out in two phases. In phase I two shapes that were signifi-
cantly different from each other were investigated: one shape having fairly generous knuckle
radii, with (riD)= 0.17 and (RJD) = 0.9, and one having a sharp knuckle with (riD)=
0.065 and (RJ D) =
1.0. Each shape included both hot-pressed and cold-spun heads, denoted
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Toroidal Shells, Torispherical Shells, Buckling under Internal Pressure 769

(a)

(b)

Fi~u re 9.133 The Li verpool hypcrb:u·ic chamber: (a) gcncrnl view (counesy of Professor G.D. Gal-
letly). (b) details (the hydraulic circuiting is not shown)

by <,pecimens P and S. respectively. The nominal cylindrical body diameters of the specimens
were D 676-769 mm and the nominal crown radii R, = 495-800 mm. The nominal
thickness of the pressed heads was dictated by manufacturing requirements to be not less
than 6 mm . whereas the spun heads were thinner, with a nomi nal plate thickness of 2.5 mm.
Four specimens of each type were made in this series. eight pressed domes with (R,It) =
121 - 150, and eight spun ones with (R.J r) = 259-334. T he pressed heads were made or a
carbon-manganese pressure vessel steel and the spun ones or a low-carbon steel. In phase II
two ha~i c head shapes with sharp knuckles were chosen. one with (r/ D) "' 0.10 and
(R,ID) "' 1.0 and one wi th (riD) ""' 0.06 and (R,/0 ) a 0.7. All the eight specimen~ of this
series were hot-pressed. made of similar material to that of the pressed heads of the first
series. two of each shape with nominal thickness of 6 mm and two of 8 mm. referred to as
A and B. respectively.
In phase I very detailed shape and thickness surveys were performed. The external surface
dome shapes were measured by the profile traverse method: the specimens were mounted
on a rotati ng table located on a horizontal boring machi ne and were measured wi th a dial
gage, which was attached to a spi ndle that cou ld move horizontally and a boring head that
could move vertically, thus yielding complete profi les at a number of meridians (actually
12). The corresponding thickness variations were also recorded wi th an ultrasonic probe. A
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
770 Shell Buckling Experiments

special computer program, TORISP, was written to process the shape and thickness data and
utilize it to determine best-fit shape parameters and imperfection amplitudes and generate
data suitable for input to the BOSOR 5 elastic-plastic shell buckling program [2.63], which
was employed to calculate the theoretical buckling pressures. These measurements also
showed the considerable thinning that appeared in the thin-spun heads toward the knuckle
and flange regions, whereas the thicker hot-pressed domes had fairly constant thickness.
In phase II less detailed shape and thickness measurements were carried out. There the
radii in the crown region were measured with a chord gage (as proposed by Kendrick in
[9.368]), whereas the meridional radii of the knuckle regions were measured with a shape
gage in the form of a series of needles mounted in a frame. The material properties were
measured on tensile specimens taken from plates that had been subjected to the same heat
treatment as the domes themselves. The material behavior was found to be essentially elastic-
perfectly plastic up to a strain of about 10 times the yield strain and could therefore be
modeled in the computations as elastic-perfectly plastic. It may be pointed out here that the
failure of most of the specimens was indeed by elastic-plastic buckling rather than purely
elastic buckling. One may also note that rhe buckling model parameter (a) E), discussed in
Section 5.6 of Chapter 5 and high values of which suggest the likelihood of clastic buckling
(if instability is the mode of failure), is fairly low in these domes. Here (a) E) = 0.0020-
0.0023, compared to that of a typical high-strength steel (a) E) = 0.0057 or of an aerospace-
grade aluminum alloy (a) E) = 0.0072, indicating that collapse of the heads may involve
both buckling and yielding.
For testing, the phase I domes were each welded to a similar, but thicker, backup head to
form a complete test vessel, as shown schematically in Figure 9.134. To check for possible
weld distortion, chord gage measurements of the crown region were compared with those
prior to welding, and the distortions were found to be insignificant. Prior to the phase II
tests, calculations with BOSOR 5 showed that for the test heads of the program the boundary
conditions at the flange end were not important. Therefore, all the phase II test heads were
welded to a reusable 70 mm thick circular fiat plate as backup head.
Volumetric displacement and some strain readings were recorded for all the 24 shells. In
the tests the pressure was increased in increments until collapse, which was marked by a
bang accompanied by an immediate pressure drop. The ratios of experimental and computed
buckling or collapse pressures for all the domes of the program, as well as the respective
predicted mode shapes, are presented in Table 9.5 (condensed from [9.364], where more

HYPERBARIC
CHAMB,ER
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

!INSTRUMENTATION
1 PORT

DISPLACED
OIL

VENTING VALVE
!SEALED DURING TEST)

Figure 9.134 Liverpool University tori spherical pressure vessel heads~ method of testing in hyperbaric
chamber (from [9.364])

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Toroidal Shells, Torispherical Shells, Buckling under Internal Pressure 771

Table 9.5 Buckling of the Liverpool shallow torispherical domes-experiment and theory (from
[9.364])
Dome r!R, R/t Mode shape p"P/ Phi PL''-P/Pimp
No. for perfect
shell

Pl/1 0.173 121 N=O 0.93 0.91


Pl/2 0.175 122 0.92 0.91
Pl/3 0.176 126 0.91 0.93
Hot- Pl/4 0.177 124 0.90 0.91
pressed P2/l 0.057 149 N=O 0.86 0.93
P2/2 0.055 147 0.90 0.92
P2/3 0.056 146 0.88 0.94
Phase I P2/4 0.057 150 0.88 0.90

Sl/1 0.201 259 N=8 0.86 0.78


Sl /2 0.202 25R 0.73 0.84
S1 /3 0.203 260 0.70 0.80
Spun S1/4* 0.201 261 0.95 1.08
Sl/1 0.067 332 N=O 0.99 1.05
Sl/2 0.071 330 0.99 1.04
Sl/3 0.075 335 0.94 1.04
Sl /4* 0.066 334 1.02 1.26

P3A/ 1 0.086 116 N=O 1.07 1.07


P3A/2 0.090 113 0.96 0.95
P3B/l 0.092 93 N=O 0.89 0.90
Phase II Hot- P3B/2 0.090 94 0.98 0.98
pressed P4Ail 0.077 83 N=O 1.06 1.04
P4A/2 O.ORl 83 1.03 1.02
P4B/1 0.082 73 N=O 1.10 1.09
P4B/2 0.079 74 1.09 1.08
*Stress relieved.

data are given). In the table p,.,r is the experimental buckling pressure, Phr is the buckling
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

pressure predicted by BOSOR 5 for the best-fit perfect domes and Pimr that predicted for the
imperfect domes, using the actual measured shape (taken to be axisymmetric).
The agreement between the experimental buckling pressures in Table 9.5 and the predicted
ones is very good ( -15 to ± 10 percent), except for the Sl domes (where the differences
reach 25 to 30 percent). The values in the last two columns are very close, indicating that
here the inclusion of actual initial imperfections (extended to be axisymmetric) in the cal-
culations did not improve the correlation with experiment and hence that torispherical shells
under external pressure of the type tested appear to be insensitive to initial geometric im-
perfections.
Two types of failure modes were observed in the experiments. All the pressed domes
(denoted P) failed with a single buckle off the axis of symmetry (see Figure 9.135a), whereas
all the spun domes (denoted S) failed axisymmetrically at the crown (see Figure 9.135b).
As can be seen in Table 9.5, the predicted mode shapes were, however, all axisymmetric
except one (for the perfect S 1 the mode was N = 8 and for the imperfect ones axisymmetric
N = 0). Substantial calculations were carried out with BOSOR 5 in [9.364], which showed
that even small imperfection magnitudes changed the failure mode, but here all the predicted
modes were axisymmetric, while the experimentally observed ones for the P domes were
not (see Figure 9.135a). Probably the domes failed in this manner due to asymmetric or

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
772 Shell Buckling Experiments

( (1 )

Figure 9.135 Liverpool tol'ispherical domes-typical fai lure modes


under external pressure (from [9.3641. courtesy of Pro·
fessor G.D. Galletly): (a) t)•pical hOt·prcsscd head ( Pl.
with rl 0 = 0. 16. RJt ~ 120 and RJ 0 - 0.9). (b) cold-
spun head (S2. with riO = 0.07. RJt ~ 330 and RJO
(b) I)

local imperfections in shape and thickness. Direct correlation studi es between measured
asymmetric imperfections and experimental failure modes, however, were only carried out
in other very recent Liverpool Uni versi ty st udies [9.370] and [9.371 ].
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Galletly and his co-workers continued extensive calculations with BOSOR 5 on the
buckling/collapse behavior and on the influence of imperfections on the buckli ng and plastic
collapse of torisphcrical domes under external pressure (9.356], 19.366] and [9.369]- (9.371 ].
They conc luded that the collapse pressures distinctly depended on the geometric parameter
(rl D) and that for perfect shells failure was predicted to occur either by axisymmetric yield-
ing collapse in the knuck le or by bifurcation at the knuckle/spherical cap junction. For
imperfect shells, axisymmetr ic initial geometric imperfections were found to be unimportant
if the failure was by axisymmetric yieldi ng of the knuckle, but if the failure occmTed in the
crown they affected it and lowered the collapse pressure of the dome. Reccmly non-
axisynun etric imperfections were considered usi ng the ABAQUS program [9.370] and
[9.37 I].
Concurrently, the Liverpool group carried out three series of careful experiments. one on
near-perfect machined steel domes [9.356], one on as-pressed and petal-welded steel domes
[9.370], and one on a large number of ABS plastic injection-molded models [9.37 1], all of
which deserve some discussion. They represent the modern approach of combined comple-
mentary experi mental and numerical investigations that yield the most reliable resu lts.
The study on near-perfect machi ned steel domes [9.356[ included eight 200 mm diameter
mild-steel models. Six specimens had (R,ID) = 1.0. a nomina.! (R,It) = 300 and (riD) =
0.2, 0.3 , 0.4, 0.42, 0.45 and 0.5 (a hemisphere). while the two remaining shells with
(RJD) = 0.55 and 1.25, (R,It) = 142 and 378 and (r/D) = 0.3 and 0.425. respecti vely,
corresponded to maximum and minimum buckling strength.
The specimens in these tests were machined from 245 mm (1 0 in .) di ameter bill ets of
mild steel (BS 4360 Gr. 43A). The domes and their fabrication method were identical to
those of another Liveq>ool test program on intemally pressurized torispherical shells [9.372],
wh ich is discussed below. The machine tool used was a Butler 500 lathe that had been
retrofined with a NUM 760T CNC control system. which was linked by wire to a micro-
VAX computer. The machining operation was carried out in six steps:

I.The intemal profile of the torisphere was drilled and bored, leaving 3 mm excess ma-
terial.
2. The workpiece was heat treated for stress rel ief.
3. The fi nal machi ning of the internal profile was completed.
4. A steel mandrel to fit the internal profile was machined.
5. The workpiece was mounted on the mandrel and the outside profile was machined, using
a Teledictor ultrasonic thickness gage (with a resolution of 0.00 I mm) to monitor the
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Toroidal Shells, Torispherical Shells, Buckling under Internal Pressure 773

thickness during the final machining operation and correcting the thickness errors with
the NUM system.
6. The finished torispherical shell was stress-relieved under vacuum conditions.

The clearances between mandrel and the workpiece were critical. A radial clearance be-
tween them of 0.025 mm (about 5-9 percent of the shell thickness) was found to allow easy
assembly and yet provide sufficient support during machining. The axial clearance between
the mating faces of torisphere and mandrel determined whether the workpiece became loose
during the operation or distorted upon removal of the finished shell. Best results were ob-
tained with a clearance of 0.040 mm.
The thickness of all the machined domes was measured with an ultrasonic thickness gage
at 6-20 points along each of 8-16 equally spaced meridians. The actual thicknesses were
found to differ from their respective average values by less than 4 percent, except at the
edges of the domes, where the difference rose to + 7 percent in some specimens. Detailed
measurement of the radii of curvature on one dome with a coordinate measuring machine
showed differences of less than 0.15 percent from the specified ones. The accuracy of the
Liverpool torispherical domes therefore approached that of the "classical" U.S. Navy DTMB
1963 near-perfect spherical caps [9.323], which were smaller but of similar radius-to-
thickness ratios and were machined in a similar manner, though Krenzke's aluminum alloy
hemispheres were apparently still more accurate (by about half an order of magnitude).
The mechanical properties of the mild steel were obtained from tests on 15 round and flat
tensile specimens taken from the long steel billet from which the domes were machined.
Most of the specimens were cut in the direction perpendicular to the axis of the billet
(corresponding to the hoop direction of the shells), but some were also cut in the longitudinal
direction.
For testing, each dome was bolted to a steel end plate. A small hole was drilled in the
end plate for a bleed pipe, through which water could be introduced to the interior of the
dome prior to the test. The amount of water issuing from the bleed pipe served as a measure
of the change in volume during the external pressure tests. Since, however, not much water
was expelled before the domes failed, no useful pressure versus change-of-volume curves
were obtained in this series of tests.
The domes with end plates were inserted into a vertical pressure chamber whose pressure
was controlled by a manual hydraulic pump, measured with a high-resolution diaphragm-
type pressure transducer and read on a strain indicator.
The agreement between the experimental collapse pressures of the near-perfect specimens
and the predictions of BOSOR 5 for perfect domes was very good, (Pexr/ PBosoR:J = 0.96-
1.05. The small differences (as well as the test results, which slightly exceeded the predic-
tions) can probably be explained by variations in mechanical properties (though in Table 5
of [9.372] the measured increases were slightly smaller). The measured small variations in
the dome thickness along the meridians were also included in the BOSOR 5 predictions, but
their effect on the buckling pressures was found to be less than l percent. The experimental
results therefore confirm the predicted dependence of the critical pressure on (r/ D), see
Figure 9.136, and in particular the predicted dip in the buckling strength around (rl D) =
0.45.
There was also fairly good correlation in failure modes. Six of the domes were predicted
(by BOSOR 5) to fail in an axisymmetric manner and did so in the tests (see Figure 9.137a),
whereas for the two domes for which edge buckling was predicted, with n = 17 and 12
respectively, a single asymmetric buckle was observed in the experiments (see for example
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 9.l37b).
The second experimental study, on as-pressed and petal-welded steel domes [9.370], in-
cluded 16 torispherical end closures and 2 hemispherical ones. The prime motivation for this
investigation was the uncertainty about the influence of residual stresses and geometrical
imperfections caused by the welding of crown and petaled heads (a common method of
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
774 Shell Buckling Experiments

HEMISPHERE @

¥· lOO
• • TESTA~TS

- TM[ORE~LLY-PREDICTED
( FOR I · ~0}

0.5

0 .2 1-...1....----JL-----'---...1....~---,-L...~
0.0 0.0 5 0.15 0.25 0.35 0.45 !
0

Figure 9.136 Liverpool near-perfect torispherical domes-six experimental resulls versus the BOSOR
5 predicted e<>llapse curve; (RJ D) = I, (R,/t) = 300 and 0.05 :s (rl D) s 0.5 (from
[9.356))

fabrication for large end closures) on their collapse under externally pressure. The specimens
were designed to fi t into the Li verpool hyperbaric chamber (i.e. D < 900 mm) and roughly
represented 1/10 scale models of submarine hull end closures. Six different geometries were
chosen. with two (RJ D) ratios "" 1.0 and 0.8 and (r/ D) values ranging between 0.06 and
018
To simulate the petali ng procedure of typical full-size submarine end closu res (which
would be fabricated by welding together eight individually pressed petals and a spherical
cap) without the need for a costly set of male and female dies for each torispheri cal shape,
the as-pressed heads were slit along eight meridians and along the spherical cap equator and
then rejoined by welding. S ince only the effects of the weldi ng were studied, this simpler
simulation was considered to be adequate. Partial slilling, in which the slit petals and cap
remain attached at corner poi nts, was adopted si nce it eliminated expensive jigging for reas-
sembly without significantly affecting the weld effects.
Extensive dimensional surveys were performed on each head, both thickness measure-
ments at 2300- 2500 locations with an ultrasonic thickness gage, and meridional radi i of
curvature measurements at 300-500 locations with a special dome-measuring rig (sec Figure
9. 138 and [9.373]). The dimensional data were processed to provide contour plots for thick-
ness variations and radial variations from best-fit profi les. Measured thickness variations (see
[9.373]) were found to be usually less than 5 percent of the average (except that in some of
the petal-welded domes local variations of up to 12 percent were observed); and the worst
meridian deviation for all the torispheres (the "worst" deviation being defi ned as the max-
imum inward radial de\•iation from the best fit within the crown segment) was less than 0.9
percent of the average crown radius of curvature.
For testing the specimens were welded to a 76-mm thick circular fiat plate insetted into
the Liverpool hyperbaric chamber, filled with insulating oil and connected to a vent via one

..-.-. . --.- ....--


(0 )
- ....-. -- - -· ,;"

Figure 9.137 Liverpool near-perfect torispherical shells-six


domes collapse patlems in extcmal pressure test-
ing (from [9.356]. courtesy of Professor O.D.
Oalletly): (a) six domes that failed axisymmetri -
cally as predicted. (b) a dome that failed in one
asymmetrical buckle instead of the I2 waves pre-
--`,`,`````,`,````,,

Copyright Wiley
(b) dicted
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Toroidal Shells, Torispherical Shells, Buckling under Internal Pressure 775

Figure 9.138 Liverpool externally pressurized torispherical and hemispherical domes-dome measur-
ing rig for large specimens that fit in hyperbaric chamber (from [9.373])

of the chamber instrumentation ports. By measuring the volume of oil expelled as the external
pressure was increased, plots of pressure versus volume change were produced (which here
were well defined, contrary to the experience in the tests of the near-perfect domes just
discussed). Failure of the specimens was usually accompanied by a dull bang and always
by a sudden pressure drop.
The experimental collapse pressures were compared with BOSOR 5 predictions both for
perfect shells Phr and for imperfect shells with axisymmetric imperfections based on the
measured radial deviations, Pimr· The ratios (Pcxr/Phr) for perfect shell predictions varied
from 0.76 to 1.13 (where the welded specimen El 17W, whose experimental pressure ex-
ceeded the perfect shell predictions by 13 percent, raised some doubts). Though the corre-
lation was generally good, it was decided to re-calculate Phr of El 17W with a higher stress-
strain curve obtained from another material test specimen (matched to another dome), instead
of that obtained from the two material specimens matched originally to El 17W (see [9.372]).
That raised Phr significantly and changed (Pcx/ Phr) to 0.96. However it was also decided to
test four additional as-pressed heads, two each of the two domes that yielded (Pcxp/ph 1) >
1. The resulting ratios for these four tests were (Pcx/ Phr) = 0.82-0.99.
For imperfect shell predictions, two sets of calculations were made with BOSOR 5 for
each specimen: one employing the average of the measured radial deviations as an equivalent
axisymmetric imperfection, and one using the worst radial deviation (maximum inward radial
deviation in the crown region as the "equivalent" axisymmetric imperfection. The difference
between these two imperfect predictions was small, the ratios of the worst to average ranging
from 0.93 to 1.03. The comparison of imperfect and perfect shell predictions was not very
significant and somewhat erratic, the ratios of the worst to perfect ranging from 0.93 to 1.19.
The inclusion of axisymmetric geometrical imperfections in the predictions did not bring
about a great improvement in the correlation with experiment, the ratios being (p cxjPimp) =
0.74-1.04, using the worst equivalent axisymmetric imperfections (as compared to the orig-
inal Pcx/Phr = 0.76-1.13).
In order to improve the correlation between theory and experiment, it was decided to
analyze some of the imperfect domes by a large 2D program. The finite element ABAQUS
code [9.374], mounted on the CYBER 205 at the University of Manchester Regional Com-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

puting Centre, was chosen. For the 2D analysis the actual as-measured coordinates of the
Copyright Wiley
complete domes were used as inputs. The computations
Provided by IHS Markit under license with WILEY
were quite costly, the time of a
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
776 Shell Buckling Experiments

typical ABAQUS run on the CYBER 205 being two !0 three hours. The results of the
ABAQUS 20 analyses P., for the eight domes considered agreed well with the experimental
collapse pressures. the ratios being (p<,,Jp,,) = 0.88- 1.12. a signifi cam improvement over
the correlation with BOSOR 5. Earlier I 0 axisymmetric calcu lations with ABAQUS for
perfect best -fit domes were in excel lent agreement with the BOSOR 5 results. However,
there was considerable scatter in the intluence of imperfections on the predicted collapse
pressures, and in some cases the imperfections increased the critical pressures, a possibility
already noted in Section 9. 10 in connection with the spherical caps of Sunakawa and lchida
and Yamada and Yamada 19. 127], [9.2 11] and [9.322].
As in the near-perfect specimen s of the previous test series. BOSOR 5 also predicted
failure by ax isynunetric collapse. The 20 ABAQUS analysis for the imperfect domes. how-
ever. yielded an asymmetric fi nal failure pattern (see for example Figure 9. 139a for a petal-
we lded dome E/20W) similar to the experimentally observed ones (see for example Fi gure
9.139b. showing the failed dome £/l7W). It appears that if one wants to include geometrical
imperfection in the analysis, the actual and (usuall y non-ax isynunetric) measured imperfec-
tions have to be considered to obtain a good prediction, as has also been found in extensive
studi es on cylindrical shells. which are discussed in Chapter 10.
The third swdy on ABS plastic models 19.37 1] focused on the influence of localized
geometrical imperfection on their clastic buckling under external pressure. All the ISO mod-
els tested had the same basic geometry, (RJ D ) = I. (riD) = 0. 15 and (RJ r) = 200. They
were manu factured by an injection-molding technique. in which granular plastic is fed into
one end of a cylinder. heated and plastici zed, and then forced out at the other end of the
cylinder into a cooler mold that is clamped c lose under pressure. The melt cools in the mold
and cures. and finally the molding (here the dome) is ejected. The specimens were made in
five batches of 30, one near-perfect and four with asymmetric local imperfections, asym-
metric 12• off-axis Aat patches (local increase in radius of curvature) of four different mag-
nitudes (l>ofr) = 0. 10. 0.3 1. 0.61 and 1.02. The local Aat patches were produced by di fferent
off-axis imperfection plugs (male and female) fitted into the molding inserts.
The thickness of each torisphere was measured with a gage that penni ts automatic reten-
tion of the thi ckness reading between two balls at 88 points and at 12 additional points on
and adjacent to the Hat patch. The radi us of the spherical crown was also measured at about
40 points on each shell to check for any effects of thermal shrinkage. The mechanical
properties of the ABS plastic specimens were obtained from tests on Aat coupons made by
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

the same injection-molding process as the domes. The test models produced were quite
consistent. with a standard deviation of thickness of less than 3.2 percent, (RJ D) of less
than 2 percent. and material properties of less than 2.5 percent (except v, wh ich had twice
the standard deviat ion).
For testing. the domes were clamped onto an internal mandrel attached to a rigid base
and the cavity under the torisphere was evacuated in increments by a vacuum pump. Fai lure

(C)

Figure 9.139 Live1pool petal-welded steel domes-failure modes (from


[9.370]): (a) mode predicted by 2-D ABAQUS analysis
for head E/20W. (b) experimental failure mode for head
(b ) E/l7W
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Toroidal Shells, Torispherical Shells, Buckling under Internal Pressure 777

was sudden in all cases and the buckling pressure could always be readily determined. The
bifurcation pressures for each of the 30 near-perfect domes were first calculated with Bush-
nell's elastic 1-D code BOSOR 4 [2.62]. Comparison with experimental buckling pressures
yielded (p"'P/PH) ratios from 0.98 to 1.08. The experimental buckling modes of all the domes
had four lobes, whereas the predicted ones had six waves (similar observations that the
experimental buckling mode has a smaller number of circumferential waves than the pre-
dicted ones are quite common in cylindrical shells).
BOSOR 4 was then employed to calculate the buckling pressures of the four batches of
imperfect domes, where the asymmetric off-axis imperfections were modeled as axisym-
metric located at the apex. Again the agreement between experiment and prediction was
close for most of the shells tested.
The experimental data of each of the batches were then reduced to averaged models (with
average geometries and average experimental buckling pressure), and these were analyzed
with the 2-D FE ABAQUS program. The predicted buckling pressures with ABAQUS were
very close to those predicted by BOSOR 4 with the equivalent axisymmetric imperfections.
The difference was smaller for the larger amplitudes of imperfection (within 1.0 percent for
(8/t) > 0.3) and slightly larger for the near-perfect batch and small imperfection batch (3.6
and 2.1 percent, respectively). The larger-amplitude imperfections, however, caused very
significant decreases in buckling pressures. For example, with ( 8/ t) = 1.02 the buckling
pressure was only about 40 percent of that of the near-perfect domes in both the experiments
and the predictions. For the type of fiat patch imperfections investigated in this test program,
the small asymmetry of 12o off-axis seemed to be negligible. Hence the much simpler cal-
culations of 1-D programs like BOSOR, with equivalent axisymmetric imperfections, may
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

indeed suffice in many cases, but further study is required to bound the relevant types of
imperfections and shells.

9. 11.3 Buckling under Internal Pressure


As pointed out in Chapter 2, buckling can occur in some shells when they are subjected to
internal pressure. Such shells contract in certain regions when internal pressure is applied,
and this inward movement causes compressive stresses that may lead to buckling. As dis-
cussed in Subsection 2.1.12d, application of the membrane equations of an axisymmetric
shell of revolution to a torispherical end closure of a pressure vessel (with both radii of
curvature positive) shows that the hoop stress can be positive or negative, depending on the
ratio of the radii of curvature (R,/ R,,). From Eq. (2.298) it was evident that N 11 will be
negative (compressive) when (R,/ R",) > 2. Depending on the geometry of the shell (primarily
its thickness) and its material, these large compressive hoop stresses may then cause asym-
metric buckling (see Figures 2.51 and 9.140).
Such buckling phenomena were first observed and studied in the late fifties and early
sixties. A large torispherical end closure of a 15 m diameter pressure vessel built for the
Tidewater Oil Co. in Avon, California, failed under internal pressure in 1956 while under-
going its hydrostatic proof test. This motivated elastic analyses by Galletly [9.375] and
[2.58], which showed that the direct hoop stress in the knuckle was compressive and ex-
ceeded the yield point of the material at a number of locations. In his 1959 paper [2.58]
Galletly also pointed out the possibility of elastic buckling for tori spherical heads made from
a material with a relatively high yield point.
An earlier report of buckling failure of a pressure vessel with stainless steel torispherical
ends by Meesters and Slaaf [9.376], apparently went unnoticed. The 1961 buckling failure
under internal pressure of a 20 in. diameter carbon steel dished end (reported in [9.377])
was better known. The buckling failure during proof testing of two 27 in. diameter aluminum
torispherical bulkheads of an experimental design for the U.S. Jupiter IRBM missile (reported
in [9.123]), which occurred at about the same time, left a strong impression on the aerospace
industry and encouraged a series of experiments on rigid PVC models [9.123] and the first
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
778 Shell Buckling Experiments

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Figure 9.140 Circumferential buckling due to internal pressure in steel torispherical shells: (a) Not-
tingham University 2-m diameter stainless steel shell-test rig, (b) closeup of buckled
knuckle (courtesy of Professor P. Stanley). (c) CEA Saclay, France. large torispherical
head with Rlt = 200. tested under internal pressure-buckled knuckle (counesy of Dr. A.
Combescurc)

proper (though linear) buckli ng analysis by Mescall [9.378] at the U.S . Army Materials
Engineering Laboratory, Watertown, Massachusens.
As ment ioned in Section 9.3, Adachi and Benicek at Watertown fabricated their PVC
torispherical specimens by a simi lar thermal vacuu m technique to that employed by King
and Adam [9. 122], di scussed there in detail. PVC was also chosen because of its high
(a) E) ratio"" 0.01, slightly higher than aerospace grade aluminum alloy. They too obtained
accurate shells (with thickness variations in the knuckle region of less than 3 percent of the
average) at low cost, permitting them to test 50 models covering four geometries. including
that of the failed Jupi ter bulkheads, with (r/1) "" 60-500. Their test setup was fairly si mple:
the torispherical shells were clamped with clamping rings and gaskets by 6 C-clamps to a
thick aluminum base plate. Pressurization was accomplished by introducing water continu-
ously at a controlled low rate (so as to preclude strai n rate effects in the PVC) from a
pressure accumulator, this control being e ffected by a metering valve. T he applied pressure
was measured by a pressure transducer connected through the base plate and monitored by
an osci llograph recorder. A central vent that reached the top underneath the dome was
provided to eliminate any trapped air. Due to the slow pressurization, the pattern and se-
quence o f buckles could be observed visually. The occutTence of each buckle was noted and
its shape marked directl y on the bulkJ1ead. But though the test rig was sim ple, it yielded
consistent results, with a maximum scaner of approximately ± I0 percent, and it was the
first systematic experimental evidence of circumferential many-wave buckling in the knuckle
region of torispherieal shells subj ected to internal pressure.
In the three decades that followed, the buckling and collapse of thin intemally pressurized
torispherical shells has been extensively studied, both ex perimentally and analytically (see
for example Gallctly's surveys in [9.379J- [9.J81J or 12.59)-[2.61]. [9.163)-[9. 166], [9.372],
and [9.383]- [9.393]). Most of the experimental investigations have been carried out at five
research centers: the Division of Structural Engineering of the University of Manchester
Institute of Science and Technology (UMIST), the Depa1tment of Mechanical Engineering,
Universi ty of Li verpool, and the Department of Mechanical Engineering, University of Not-
tingham, all three in the U.K.; the Centre d'Etudes Nucleaires, Saclay, France; and the
Chicago Bridge and Iron Co. (CBl), Pl ainfield. Illinois. Some isolated tests were pe1formed
at other centers (see for example [9.393]). Some salient poi nts of these studies are now
briefly discussed.
As in the case of external pressure loading. two types of model torispherical shells were
tested
Copyright Wiley in these studies: small machined near-perfect models and larger fabri cated ones (full
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Toroidal Shells, Torispherical Shells, Buckling under Internal Pressure 779

size or nearly so) of production quality. The latter specimens were manufactured by one of
the two methods usually employed in pressure vessels (see [9.389]):
1. Pressed and spun. In this method a flat circular sheet is first pressed into the required
crown radius R, in a concave die and hammered in a machine hammer to a smooth
surface finish. A knuckle is then formed on the dome by turning in the rim while rotating
the dome about its axis. A fixed inner roller acts as a former and the outer roller can
move meridionally. There is considerable thinning and work-hardening in the knuckle
region.
2. Crown and segment (sometimes also called petal-welded). Here the crown of the dome
is pressed separately as a spherical segment and hammer-finished. The knuckle of the
dome is pressed from a frustrum of a cone, made from four or more equal conical
segments (welded together before pressing). Then the knuckle and crown are welded
together. Alternatively, the petals can be pressed into shape separately and then welded
together with the crown. Finally the knuckle is usually hammer-finished. This method
is more expensive than the pressed-and-spun method, but the thinning and work-
hardening incurred are less severe.
The Manchester UMIST experiments [9.165] and [9.166] included 22 small 135 mm
diameter aluminum alloy specimens, machined from solid billets of 152 mm diameter. The
relevant geometries of the domes were (RJ D) = 1.0, (Dit) = 53-530 and (r/ D) = 0.057-
0.227. Though considerable care was taken in the machining (the final machining was carried
out on a numerically controlled lathe and for the final work on the outer profile the dome
was filled with a low-melting alloy, Cerrobend, for stiffness), substantial variations in thick-
ness from the average were measured in the knuckle region. The thickness was measured
for each specimen at more than 400 points, and variations of about ± 10 percent were found
in the neighborhood of the first knuckle, in a few of the domes even up to ± 20 percent.
Indeed, in the tests the first knuckle always occurred at the thinnest meridian of each dome.
Better accuracy in thickness was obtained in the later Liverpool near-perfect machined mild-
steel domes of similar size [9.356] and [9.372] that were discussed in detail in Subsection
9.11.2.
The UMIST tests are, however, best known for their careful radial displacement measure-
ments and their rotating probes. The test rig shown in Figure 9.141 was developed from an

CENTRAL DISPLACEMENT

'o' RING SEALS


--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 9.141 Manchester UMIST test rig for torispherical domes under internal pressure, showing the
Copyright Wiley
radial displacement transducers and theLicensee=McDermott
Provided by IHS Markit under license with WILEY rotating displacement probe User=G,
Inc - India/8215328006, (fromBoopathi
[9.165])
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
780 Shell Buckling Experiments

earlier one employed at UMIST in a study of the plastic behavior of oblique nozzles in
spherical pressure vessels [9.394 ]. The pressurizing medium was oil and was applied using
a simple hand pump via a stop valve. An air vent tube was provided with the oil supply
pipe.
Radial displacement along a meridian was measured by an array of Mercer linear variable
displacement transducers (LVDT's) type 364, which were arranged normal to the surface
along one meridian. A typical arrangement of 9 LVDT's is shown in the figure, but arrays
of up to 17 transducers were used. The output from the LVDT's was channeled through a
Mercer 18-way probe selector and then read using a Mercer gage unit. Some LVDT's were
placed along slightly shifted meridians to permit a closer pitch, but this offset was neglected
since the deflection behavior was assumed to be rotationally symmetric.
In order to detect the buckling of the toroidal knuckle, a rotating probe arrangement was
used. The rotating probe consisted of an LVDT, or other displacement transducers, fixed to
an arm that could be rotated about the axis of the specimen through 310° of the circumfer-
ence. Its position around the circumference was determined by a rotary slidewire potenti-
ometer (also shown in Figure 9.141). The signals from the LVDT and the potentiometer
were fed to an X-Y recorder and thus the development of the buckles could be monitored.
In later tests, the rotating probe had two LVDT's, one at the sphere/torus (crown/knuckle)
junction and the other at the middle of the knuckle, both normal to the surface and probing
simultaneously, with their outputs connected to the two Y terminals of a two-pen X-Y re-
corder. The two LVDT's were calibrated in situ with feeler gages, while the angular position

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
of the arm was calibrated from a series of lines marked out earlier for positioning of the
strain gages.
During a test, all the Mercer LVDT's were set to zero at zero pressure. A scan was then
taken with the rotating probe, giving two zero pressure curves in the knuckle region on the
X-Y recorder. As the pressure was increased, similar deflection readings and circumferential
scans were taken after each increment of pressure. The thin specimens with (D/t) 530 =
(three of the first test series and eight of the second test series) failed by buckling of the
toroidal segment (the knuckle). For all these shells the rotating probe proved to be a very
sensitive device for detecting the buckles, and it was possible to detect incipient buckles
well before they were visible by the unaided eye. In Figure 9.142 plots of radial deflection
round the circumference (obtained with this probe) are presented for a typical thin dome 4B,
with (D/t) = 530, at various internal pressures. The detection of the first buckle at p = 378
kN/m 2 is shown by the arrows. The buckles developed fairly rapidly, all in an outward
direction (except two in one specimen), up to 12 waves (or 13 in one specimen). It appeared
that the elastic buckling mode initially had a very short circumferential wavelength (as pre-
dicted theoretically), but as the pressure increased and therefore the circumferential com-

ANGULM POSIT'ON, deQ

330 270 210 1~0 90 30 330 270 210 150 90 30 330 270 210 150 90 30
~

E -o.so
E 861KNim2
-o.?s

...~
-t.OO

......
..J -t.2~

~
448 2
0
-tso KN/m f\1\ lA
fi -1.?5
!II
i:
-2.00 WAVB.ENGTH 2b
-z.~

Figure 9.142 Manchester UMIST machined torispherical domes under internal pressure--development
of buckles with pressure for specimen 4B, with (DI t) =
530, the arrows showing the
first buckle (from [9.165])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Toroidal Shells, Torispherical Shells, Buckling under lntemal Pressure 781

prc>'i'e ~tress. one buckle became plastic and dc,·eloped a large radial amplitude. to be
followed by others. Finally. all the buckles were deep and plastic. but with a much >maller
circumferential wave number than the initial e l a~ti c one;..
The rather successful rotat ing probe for incipient buckle detection was recently further
developed hy Galletly and his co-workers in their Liverpool experiments on ncar-perfect
machi ned torispherical shells ;.objected 10 internal pre~sure [9.372]. The Li verpool test set·
up is 'hown in Figure 9.143a. The domes were mounted with a clamp ring onto a base plate
thnt could he rotated about a horitontal axis" hile maintaining a pressure supply to the dome
\"ia a rotating seal. A geared motor with a control unit pro' idcd a rotational speed or about
3 rpm. Here the stationary tran'>dueers. five Sensonic LVDT"s and one Welwyn '>train gage
displacement transducer. monitored the surface or each torispherical shell as it was rotated
whi le '>objected to internal pressure. As seen in Figure 9. 143a and b. the LVDT's were
all ached 10 a carrier arm in location; that covered the cap/ knuckle transition region, whereas
the Welwyn transducer was held by a single magnetic st:tnd. The LVDT's logged via an AD
convener on an Opu~ PC. whi le the data from the ~trai n gage displacement transducer were
collected with a strai n indicator onto a time-base cha11 recorder. In the tests, the domes were
rotated at each pressure increment. while the LVDT data were collected on the PC and chan
recorder trace~ were taken of that from the Weh') n tran\ducer and some of the LVDT's.
l n~tead or the rotating probe-. u;ed at UM1ST. the rotating head and stationary probes were
employed for the detection of the threshold buc~ling pressure and exami nation of the de-
velopment of the buckling w<~ves.
From the recorded radia l dii-.placcments the init ial buckli ng pressure, or threshold buckling
prcs~u rc. was determi ned by two methods. one usi ng the radial dellection data stored on the
PC and the other using the chart records of the stra in gage transducer.
In the fi rst method the r:Jdial dellection data. mea~ured by the five LVDT's at every 1.8°
Clrcumferentially and stored on the PC. were reduced :after the tests and presented as ploL~
of the dellected shell ~urface. The first step in the dma reduction at each pre~~ure level was
to find the mean deHection for each probe. Thi~ mean repre\ented the average contraction

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(O)

(b)

Fi~:ure 9. 143 Liverpool ncnr-pcrfccl lorbphclical shell• under internal pressurc-IC\1 sclup (fo·om
[9.372]. couo·1csy of Profc\Snr G.D. Gallealy): (al 'ide view of 1es1 rig. (b) view of the
probes
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
782 Shell Buckling Experiments

(or expansion) of the shell and was then subtracted from the corresponding 200 radial de-
flection measurements. The maximum deviation from the mean was also deduced for each
probe and was plotted against the internal pressure. Such a plot of the growth of the maxi-
mum deviation (normalized by the thickness t, as recorded by LVDT C (which was located
slightly above the middle of the knuckle) is shown in Figure 9.144 for all the domes tested
(domes 1 to 6, with Dlt = 400-800, RJD = 1.0 and riD = 0.15 and 0.20). Up to the
threshold buckling pressure (denoted by triangles in the figure), the growth is almost linear,
but once the threshold is exceeded the growth in the maximum deviation (from the mean)
suddenly accelerates. The visible buckles appeared considerably beyond the initial buckling
pressure. Additional similar plots were presented in [9.372], as well as circumferential traces
for one dome, recorded by three LVDT's at four pressure levels, which showed that the
development of waviness (which characterizes bifurcation buckling) correlated well with the
threshold pressure.
In the second method the output from the strain gage displacement transducer produced
a trace on a chart recorder that was examined after the test. A typical trace showing the
development of buckles is presented in Figure 9.145 for dome 2 (with additional ones shown
in [9.372]). Again the initial buckling pressures were always lower than the "visible buckles"
pressure.
Most of the torispherical shells subjected to internal pressure that were tested at Man-
chester and Liverpool failed by elastic-plastic buckling or plastic buckling. Their buckling
behavior will therefore be discussed in more detail in Chapter 16. It should, however, be
reiterated here that internally pressurized torispherical shells can fail by axisymmetric col-
lapse or by buckling, which involves many circumferential waves in the buckling region,
and that this buckling mode will be elastic only for very thin shells and elastic-plastic for
most practical pressure vessel dished-end geometries.
Turning now to the larger fabricated shells, the objective of the Nottingham experiments
[9.388] and [9.389], carried out in the seventies, was to provide data for designers by testing
three series of typical full-size stainless steel torispherical ends of routine production quality.
The experiments included 17 torisphericai shells divided into three groups: one of six shells
of nominal diameter 54 in. (1372 mm) and nominal (D/t) = 422, one of seven shells with
D = 81 in. (2057 mm) and (Dit) = 633, and one of four shells with D = 108 in. (2743
mm) and (D/t) = 840. All groups included both pressed-and-spun domes (about two-thirds
of the specimens) and crown-and-segment domes. The knuckle radius ratio range was

15

Probe C
2

A ~~~:r:o BUCKLING
z LO 5
Q
!<( e r:~~E ~~;gsEYE)
>
~

Pressure
N/mm2

Figure 9.144 Liverpool near-perfect torispherical shells under internal pressure-local maximum de-
flection (divided by the shell thickness) versus internal pressure, as recorded by probe C
(located not far from the middle of the knuckle). The threshold buckling pressures for
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

domes 1 to 6, (Dit) = 400-800, are denoted by triangles (from [9.372])


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Toroidal Shells, Torispherical Shells, Buckling under Internal Pressure 783

p = I. 379 (200)
~ II
ru ~
N/mm2 (lbf/in2)
lfl

~
p = 1.069 (155)
... 1. wl
I"""V' 1.11 BUCKLES VISIBLE
A:r 1.035 N/mm2

p =0.931 (135)
'"'w ~
p = 0862 (125)
"""'"..,..- p = 0.758 ( 110)

p = 0.689 (100)

0 7T/2 7T 37T/2 27T 8


Figure 9.145 Liverpool near-perfect torispherical shells under internal pressure-chart recorder traces
showing the development of buckles in dome 2, (Dit) = 500, (RJD) = 1.0 and
(rl d) = 0.15 (from [9.372])

(r/ D) = 0.056-0.167, with more than half the specimens having (r/ D) > 0.074. In most of
the shells (RJ D) = 1, except for five domes in which (RJ D) = 0.722-0.833. The length
of the cylindrical portions of the ends varied, L = 10-48 in. The material for all the shells
was a high-proof stress austenitic stainless steel (Silver Fox Hi-proof 304, produced by the
British Steel Corporation).
Both thickness and curvature were carefully measured at over 20 locations on eight me-
ridians on each dome, with an ultrasonic probe and a three-legged out-of-plane measurement
device, respectively. For strain measurements about 100 pairs of 50° foil gages were bonded
to each shell, the 200 interior gage leads being brought out via a special sealed pole attach-
ment. In the pressed-and-spun ends the thickness measurements revealed the considerable
thinning (up to 25 percent) in the knuckle region due to the cold-forming operation, which
depended on the geometry and the skill and consistency of workmanship of the operator.
However, since the maximum circumferential stress occurred in the knuckle near the knuckle/
crown junction, where the thinning was relatively small, the effect of the thinning on the
buckling pressure was probably less severe. In the crown-and-segment welded domes, the
thinning measured was much smaller, not exceeding 3 percent of the average thickness.
Significant variations from the nominal meridional curvatures were measured in both the
crown and knuckle, which appeared to be strongly operator-dependent. The mechanical prop-
erties were obtained from the plate material test specimens, both in as-received condition
and after having been rolled down to 70 percent of its original thickness (simulating the
pressing and spinning operation).
For testing, one end for each diameter was used as a base end, the others being welded
in succession to the base end, cylinder to cylinder, to form a closed pressure vessel (the
joined cylindrical portions of the ends becoming the cylindrical shell of the vessel). The
base ends were of crown-and-segment construction and were expected to be the strongest
ends. The vessel was supported, with the axis vertical, by means of an angle- or channel-
section ring welded to the cylindrical portion of the base end. After attachment of internal
and external strain gages and leads in two stages and water proofing, the vessels were filled
with water and tested at internal pressure increments until buckling occurred.
The pressurizing and strain gage bridge systems are shown schematically in Figure 9.146.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

In the pressurizing system (b), the small-capacity pump built up the pressure, with the ac-
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
784 Shell Buckling Experiments

100 DUMMY GAGES ~ 100 OUMfiiiiY GAGES


FOft INN!.ft SUftf'ACI~ F~ OUTE.ft SUAf'AC
STAAIN GAGES STIIIAIN GAGES

200
CHANNEL
SCAHN[I'I

VESSEL

0~60ibrtln2
~ESSURE
REGU4.ATOR

VESSEL FLL.ED AND


DRAINED THROUGH
THIS VALVE
(b)

Figure 9.146 Nottingham large fabricated tori spherical shells subjected to internal pressure-schematic
diagram of (a) strain gage bridge recording system and (b) pressurizing system (from
[9.388])

cumulator smoothing out pressure fluctuations, and the pressure regulator was then adjusted
to bring the vessel pressure to the required value. When significant effects of minor leaks
or creep deformations were noted, the vessel pressure was easily kept constant by adjusting
the regulator. In the data logging system (a) switching was performed by four 50-channel
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Solartron scanner head units connected to a Solartron scanner controller, with a scanning
rate of 8 gages/ sec. The strain gage outputs were processed by a PDP 8 mini-computer,
which easily accounted for zero pressure readings, and therefore bridge balancing was not
necessary.
In the tests, the vessel pressure was slowly increased to a desired value and held constant.
The strain gages were scanned with scans completed in less than half a minute) at regular
intervals (typically 20-30 minutes), till the strains had settled, while other gages were also
read simultaneously and both ends were carefully inspected for any signs of buckles. The
pressure was increased in steps till the onset of buckling, and then slowly beyond as the
buckles developed. Each test lasted for several days, with different overnight pressure levels,
as detailed in [9.388] and [9.389], where other details of the tests are also presented.
Buckling occurred in the knuckle of each end closure over a range of pressures. After the
first buckle (or sometimes several) had formed, there was a decrease in vessel pressure, but
it was possible to continue increasing the internal pressure, with additional buckles appearing
periodically. A typical pressure-deflection curve for thin internally pressurized torispherical
shells (Figure 9.147, from [9.387)) demonstrates this general behavior. One should note the
stable postbuckling behavior in the figure, because of which the buckling of thin dished ends
under internal pressure is not sensitive to small geometric imperfections.
There were two distinct types of buckle in the Nottingham tests: one, designated GO
(gradual outward), which developed gradually as an outward bulge elongated in the merid-
ional direction; and the other, the snap buckle, designated SI (snap inwards), in which the
major deformation occurred extremely rapidly as a deep inward indentation along a meridian.
Both types of buckle were preceded by the formation of a series of just-perceptible ripples
in the knuckle region that grew with increasing pressure. For the GO type, one of the ripples
would develop, usually within a few minutes, into a stable outward buckle, with no noise;
whereas the SI type buckles developed with a resounding noise and vibration of the sup-
porting frame. Usually about five to eight, and sometimes up to nine, final circumferential
buckles were observed in most domes.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Toroidal Shells, Torispherical Shells, Buckling under Internal Pressure 785

1.0

(j)9

0.8 ~OCCURRENCE OF
FIRST BUCKLE
0n. 0.7
~
w 0.6
a:
:::J
en END OF LINEAR PORTION OF
en 0.5
IAJ PRESSURE CROWN
a:
n. DISPLACEMENT CURVE
.J 0.4
<(
z

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
a:
w Qj,

~
0.2

0.1

0 5 10 15
AXIAL DISPLACEMENT OF CROWN (mm)

Figure 9.147 Typical pressure versus crown displacement curve of a thin torispherical dome subjected
to internal pressure (from [9.387])

A notable aspect of the buckling behavior was the time delay between the attainment of
the initial pressure and the formation of a buckle. Though usually the delay amounted to
less than a few minutes, longer times of up to three hours were sometimes observed. This
effect was apparently associated with time-dependent plasticity (or "cold creep") phenomena
noted at stress levels below yield.
Another extensive experimental program concerned with large fabricated torispherical
domes subjected to internal pressure was carried out at the Saclay nuclear research center
in France [9.390]. Sixteen 500-mm diameter torispherical heads, with nominal (RJ D) =
1.0-1.1, (D/t) = 333-1000 and (r/D) = 0.04-0.10, were made by cold-spinning carbon
steel plates. The thickness of the specimens was measured by an ultrasonic probe at many
locations, and though only small circumferential variations were noted (less than 5 percent
of the average thickness), there were the typical significant meridional variations associated
with the spinning process (for most shells these were less than 12 percent of the average,
but for some Mmax was larger, in one shell up to 22 percent).
The buckling behavior (see Figure 9.140c) was roughly similar to that observed in the
Nottingham tests, with stable postbuckling and attainment of pressures up to twice the buck-
ling pressures. The buckled knuckles of the Nottingham and Saclay shells exhibited similar
patterns. To obtain a better measurement of the initiation of buckling, a rotary probe, some-
what similar to that used by Kirk and Gill [9.165]. was employed. The rotating probe records
showed that buckling was again preceded by low-amplitude waviness, with 20-40 circum-
ferential waves, which developed, as the pressures continued to increase, into about 6 visible
waves.
A third example of experiments on large fabricated torispherical domes under internal
pressure is the tests carried out in the eighties at the Chicago Bridge and Iron Co. by Miller
and his co-workers [3.391] and [9.392]. Two 192 in. (4877 mm) diameter welded steel
domes, with (RJD) = 0.9, (Dit) = 980 and 711, respectively, and (riD = 0.17), were
fabricated using methods, workmanship, inspection techniques and equipment similar to
those employed on full-scale shells representative of containment vessels and other large-
diameter pressure vessels.
A salient feature of the CBI tests was the specially designed geometry measuring frame
shown in Figure 9.148. The frame included a precision-machined rotating arm for mounting
and referencing the displacement transducers (LVDT's) and a support truss with a reduction
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
786 Shell Buckling Experiments

COMPUTER CONTROLLED REDUCTION GEAR


~-------
~TEPPING MOTOR

GEOMETRY M~ASURING
J BEARING
DEVICE
--~-

TEST MODEL

~0 TON TEST FLOOR

Figure 9.148 Test setup for torispherical head models at the Chicago Bridge and Iron Co., showing
the special geometry measuring device (from [9.392], courtesy of Dr. C.D. Miller)

gear and a stepping motor for computer control of the measuring arm location. As the arm
was rotated around the shell, the transducers measured the shape of the shell relative to the
machined surface of the arm. Any change in the radial dimensions could therefore be pre-
sented as a contour of the actual shape or as a change in shape from some previous mea-
surement. Fifteen LVDT's were mounted on the arm, as shown in the figure, to measure the
initial shape and monitor its change during the test. Shape measurements were taken at 3o
increments around the entire model at predetermined pressure values, each scan yielding
1680 readings.
The initial shape measurements indicated that model 1 had surface deviations (of the
sphere and knuckle sections) up to 1 percent of the spherical radius inward and up to 0.5
percent outward, as well as noticeable flattening at the welds. The shape of model 2 was
better, with maximum surface deviations of 0.4 percent inward and 0.5 percent outward.
Extensive strain gage measurements were also taken with 50 two-element strain gages
installed on each dome. Pressures, strains and shape changes were recorded on an HP 3497 A
data acquisition system with two PCs.
The buckling behavior was again essentially similar to that shown in Figure 9.147 and
observed in the Nottingham and Saclay tests. After the first buckles appeared, the pressure
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

could be increased much further up to three to four times the initial buckling pressure before
rupture occurred, with buckles developing in the knuckle region (each usually accompanied
by a small pressure drop), till 14 buckles had formed around the circumference.
Ellipsoidal shells, which are an alternative type of end closure in pressure vessels, are
analyzed and tested in a similar manner to torispherical shells (see for example [9.162] and
[9.395]) and therefore appear in some studies together.
Here one may ask whether complete circular toroidal shells are also prone to buckling
when subjected to internal pressure. Intuitively one would suspect that buckling could occur,
for the internal pressure that inflates the torus and expands its circular cross-section could
cause a contraction of the region of the internal equator (Din Figure 9.128a), which would
result in compressible stresses and eventually in buckling. But in reality this is not the case,
since the inflated torus expands not only in its cross-section, but also in the distance (b in
Figure 9.128a) between the center of the circular meridian and the axis of revolution. As a
result, the internal equator D remains practically at rest without any displacement and hence
no compressive stresses arise and no buckling occurs.
The coupled deformation of the internally pressurized circular toroidal shell, in which both
the cross-sectional radius rand the distance b grow simultaneously, was predicted by analysis
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Shells Subjected to Transverse Shear Loads 787

and computations (for example [9.397] or computation' with BOSOR 5. (2.63)). But it was
al~o \hown experimentally by Mercier et al. (9.398( for a thin, highly ela.~tic toroidal shell
made of vu lcanized natura l mhher whose original and deformed geometry (with very large
di ~placements) were measured. demonstrating the large increase in b and zero displacement
at D. However. the absence of compressi ve stres~ and therefore no danger of buck ling are
ensured only for toroida l shells of circular cros~-sect ion. whereas in toroidal shells of asym-
metric cross-sections comprc~~ive stresses that could cause buckling may occur (see for
c~amplc [9.399]).
Finally. one should remember that buckling due to internal pressure can occur not only
in tori\pherical or ellip~oidal ~hells. but also in others. For example. cylindrical shells of
elliptic cross-section arc prone to buckling when subjected to internal pre~sure. as was re-
cently found by Albu~ and Ory at the RWTH Aachen whi le working on a tank design concept
for hypersonic aircraft (sec (9.400(). They showed that internal pressure produced in elliptic
cylindrical shells compressive stresses in the meridional direction at the top of the semi-

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
major axis that at a critica l value of the pressure resulted in general instabili ty.
Such a buckl ing failure of an elliptic cylindrical ~hell made of Hostaphan (a polyester
film \imilar to Mylar) i~ shown in Figure 9.149. The cylinder has a semi-axis ratio (alb)=
2. a semi-major axis a = 100 mm and a length of 290 mm. The thin elliptic cylindrical
shell i\ pressurized by a simple htcycle pump. When the critical internal pressure is reached.
fir~t shear buckling occurs in the vicinity of the bulkhead,, But then. with an additional slight
incre:l\c in internal pressure. failure by general in~>tability takes place, as can be clearly seen
in the ligurc.
lienee the designer always has to remember that J'or some types of shells internal pressure
can also be a cause for buckling.

9.12 Shells Subjected to Transverse Shear Loads


Before we close Chapter 9. one other basic loading ca~c. that of horizontal shear. warrants
a brief discussion. This ca\e appear~ primarily in !>Cbmic loading. which will be discussed
in Chapter 19 but has already been mentioned brieny in Section 9.10 in connection with the
Karlsruhe spherical containment shell.
T he earliest experimental investigation of buckling of cylincJrical shells subjected to hor-
i.wntal shear was the tests carried ou t by Lundquist at NACA Langley in the thi rties on
cantilever duralumin cylindc~ in combined transverse shear and bending (9.401), as part of
the NACA extensive studies of the strength of stres,ed-~kin aircraft structures. The shells

Figure 9.149 Buckled Ho,taphan cylindrical shell ol' elliptic


cross-section >ubjcctcd to intemul pressure. the
semi-axis ratio (a/ !J) = 2. the semi-major axis
" = 100 mm.thc length L = 290 nun nnd (a/r) =
286 (from )9.400))
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
788 Shell Buckling Experiments

had (Rit) = 325-1455, and all failed by elastic buckling. Different ratios of bending moment
to shear were applied by placing the loading jack at different distances from the cantilever
shell, and three regions of failure were studied: bending failure, shear failure and the tran-
sition region. For small values of the bendling to shear ratio (MIRV), representing short shells
(where M is the applied bending moment and V the applied transverse shear force), failure
occurred in shear by formation of diagonal shear wrinkles. Lundquist therefore suggested
that the shear strength should be closely related to that under torsion and found that the
critical transverse shear stress was about 1.25 that for torsion for very small (MIRV) ratios.
This idea was later further developed by Yamaki et al. [9.402], as detailed below.
After Lundquist, the study of the buckling of cantilever cylindrical shells under transverse
shear loads was resumed only in the sixties and seventies. A rough theoretical approximate
solution was first obtained by Lu [9.403] for long shells for which bending was dominant.
A better approximation, using Donnell's linearized shell equations, was then presented by
SchrOder [9.404], who showed separate curves for buckling due to transverse shear and for
buckling due to bending. His calculations for simply supported shells demonstrated that for
short shells, say (L/ R) < 3, the buckling stresses for transverse shear are well below those
for bending. A similar solution was developed independently a few years earlier in the then
Soviet Union (see [9.405]).
In the late seventies Yamaki, Naito and Sato carried out a series of precise experiments
on the elastic buckling of cantilever cylindrical shells made from polyester film (the same
Diafoil used in Yamaki's earlier tests discussed in Sections 9.4 and 9.7) under the combined
action of transverse edge loads and hydrostatic pressure [9.402]. Eight cylinders, with
(R It) ~ 400 and (L/ R) = 0.23-2.28, were each tested under various load combinations.
Yamaki et al. also extended Schroder's cakulations to the combined loading cas<:. For trans-
verse shear, they then compared their and Lundquist's earlier experimental results with the-
oretical predictions, including the elastic critical shearing stress for clamped cylinders under
torsion, which they suggested would be a good approximation of the maximum shearing
stress under transverse load (see Figure 9 . 150).
Note that the critical stress under torsion was calculated for a fully clamped cylindrical
shell, whereas the critical shearing stress under transverse load was obtained for simple

EXPERIMENT
0 : YAMAKI ET AI_ (Ref 9.402)
~:;·"FAILURE" }
'V : "FIRST WRINKLE" LUNDQUIST (Ref. 9.401)

THEORY
: kv SCHRODER !Ref. 9.404)
---: k 5. YAMAKI AND KODAMA (Ref. 9.406)
,~

kv =(VL2t7T 3 RD)
k5 =(TL2t27T3 R 2 Dl
D = El/12(1-11 2 )
Z =~L2 t(Rt)

IOILO__J_~~~~~--L-L~~LUI0~3~L_~LL~IQ4
Z=-Ji=;2'!L2tRTl

Figure 9.150 Buckling of cylindrical shells subjected to transverse edge loads-comparison of non-
dimensional experimental results with theoretical predictions, including those for similar
shells under torsion (from [9.402], courtesy of Dr. J. Albus). Note that for transverse
shear simple supports were assumed, whereas for torsion the shell was taken to be fully
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

clamped
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Shells Subjected to Transverse Shear Loads 789

~uppon-.. This explains the di !Terence between the good agreement of lhe two predictions in
Figure 9. 150. and lhe about 25 percem higher critical transverse shear stress for very shon
shells found by Lundq uist, whose approximate cri tical ~tress under torsion also a.ssumed
simple su ppons. The clamping m.sumed by Yamaki et al., which indeed simulmed the test
condit ions mo re correctly, raised the predicted critical torsional stress for very short shells
(say Z < 50) sufficien tly to bridge lhe about 25 percent gap noted by Lundquist. However,
because of the different boundary conditions used, the comparison of Figure 9.150 should
preferably not be applied to 'cry shon shells.
Some earlier experimental studies on lhe buckling of cantilever cylindrical shells under
transverse shear loads carried out in Russia were bricny discussed by Grigoliuk and Kabanov
(9.405], but lhe test resuhs presented included o nly very few shon shells with (L/ R) < I
for which the shear loading wa• dominant.
In the early eighties Gal let ly and Blachut tested three Mylar cylindrical shel ls, like those
st udied by Yamaki et al., with (Rit) = 200-400 and (L/ R) = I [9.407]. Later they initiated
a similar test program on plastic huckling of short steel cyl indrical shell s subjected to trans-
verse edge shear loads [9.408(. Though primari ly aimed at pla.stic buckling. the Liverpool
Univen.ity test setup shown in Figure 9.151 is typical of current practice for elastic and
plao;tic buckling experiments on cylindrical shells ~ubjcctcd to transverse shear loads. The

CENTRE LINE (o)

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(b)

(c )

Fi~:u re 9.151 Li verpool Univcrsi1y 1cs1 setup for short cyli ndrical ~hells subjcc1cd tO trnnsvcr;,c loads
(from J9.40RJ): (a) detail of flange auachmcnl, (b) c)ulline of the test sc1up, where gage
A measures horizomat deflection ancl gage B vcr1icat dcOccLion, (c) a buckled mild s1eel
shell in 1he ICSI Jig
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
790 Shell Buckling Experiments

test shells had heavy flanges attached to their ends (in these Liverpool tests the models were
all made by wrapping steel sheet around a mandrel, welding the longitudinal seam and then
welding the heavy flanges onto the ends of the short cylinder, as shown in Figure 9.151a).
The lower flange was bolted to the rigid base, and the transverse shearing loads were applied
to the upper flange by a ram via a load cell. The links of the load cell permitted vertical
deflection of the upper flange during shear deformation of the shell. The resulting horizontal
and vertical deflections were measured by gages A and B respectively, as indicated in Figure
9.15lb. A stop on the opposite side of the flange prevented excessive postbuckling defor-
mations.
The buckling was plastic for practically all the models. The maximum load attained in
the test was taken as the failure load for all the specimens, the buckles usually occurring
just after the maximum load (see Figure 9.151c). Of the 14 specimens, 5 were duplicates of
the first five models, made to obtain information on the repeatability of the buckling. The
variation in buckling stress between original and duplicate was indeed found to be small,
the largest variation being about 10 percent. This and the second-largest variation of 8 percent
occurred when the initial out-of-circularities of one of the pairs were larger than for the
other shells (8/ R > 0.0005). Unfortunately, no initial imperfection plots nor their detailed
analysis were presented.
At the conclusion of each test (and after plastic buckles had formed) the model was rotated
through 180° and re-tested in order to study the effect of relatively large imperfections on
the buckling load. This second buckling load was found to be, for all but one of the shells,
at least 90 percent of the original one. However, the imperfections introduced in this manner,
representing waves orthogonal to the expected buckling waves, were not of a critical shape
(they were probably primarily of a stabilizing nature), and therefore the conclusion "that
this shell buckling problem is not very imperfection-sensitive" was not justified.
A similar series of experiments on cantilever cylindrical shells subject to transverse shear
load was carried out at the Institute of Industrial Science of the University of Tokyo in the
mid-eighties [9.409]. The models, with (R/t) = 150 and (LIR) = 1.83-4.74, were again
made by wrapping mild steel and aluminum sheets around a mandrel and then welding the
longitudinal seam. Finally, the ends of the cylinders were again welded to thick metal flanges.
The Tokyo test rig was also similar to the Liverpool one, applying the shearing forces directly
to the upper flange. The main purpose of the Tokyo experiments was to examine the influence
of the (LI R) ratio and the material properties on the failure mechanism, but the effect of
cyclic loading was also studied and found to be immaterial.
More recently, the effect of more realistic boundary conditions, i.e. of an elliptical head
on one end of the cylindrical shell instead of heavy flanges on both ends, were studied at
the Central Research Institute of Electric Power Industry in Japan [9.410]. Six models, with
R = 1000 mm, (Rit) = 167 and 250 and (RIL) = 1, made by rolling stainless steel plates
and welding the longitudinal seam, were prepared. Elliptical heads were welded onto one
end of four models with heavy flanges at the other end, while two specimens were prepared
with heavy flanges welded to both edges. The shear buckling loads decreased slightly for
cylinders with elliptical heads, as compared with those with heavy flanges at both ends of
the cylinders. Decreases of 1-12 percent were noted, becoming more significant with in-
crease in (Rit).
The recent and ongoing studies on the buckling of cantilever shells subjected to transverse
shear loads were mostly associated with nuclear and earthquake engineering, primarily in
connection with containment shells and fast breeder reactor vessels (see for example [9.333],
[9.336] and [9.411]), and the interested reader should therefore turn to that literature, as was
already indicated in Section 9.10 in relation to the spherical containment shell discussed
there.

References
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
9.1 Fairbairn, W., On the Resistance of Tubes to Collapse, Report of the British Association for the
Copyright Wiley Advancement of Science-1857, 27, London, Licensee=McDermott
1857, 215-219.
Provided by IHS Markit under license with WILEY Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 791

9.2 Fairbairn, W., On the Resistance of Tubes to Collapse, Phil. Trans. of the Royal Society of London,
148, 1858, 389-413.
9.3 Fairbairn, W., On the Collapse of Glass Globes and Cylinders, Report of the British Association
for the Advancement of Science-1858, 28, London, 1858, 174-176.
9.4 Grashof, F., W. Fairbairn's Versuche iiber den Widerstand von Ri.ihren gegen Zusammendriickung,
Zeitschrift des Vereines deutscher Ingenieure, 3, 1859, 234-243.
9.5 Love, M., Sur la Resistance des Conduits Interieurs a Fumee dans les Chaudieres a Vapeur,
Memoires et Compte Rendu des Traveaux, Societe des Ingenieurs Civils de France, 12, 1859,
471-500.
9.6 Nystrom, J.W., A New Treatise on Steam Engineering, Putnam's Sons, New York, 1876, 106-
107.
9.7 Unwin, W.C., On the Resistance of Flues to Collapse, Proceedings of the institution of Civil
Engineers, 46, (4), 1878, 225-241.
9.8 Belpaire, T., Note sur las Resistance des Tubes Pressees de l'Exterieur, Annales du Genie Civil,
March, 1879, 177-187.
9.9 Wehage, H., Zur Bcrechnung der Flammri.ihre von Dampfkesseln, Dingler's Polytechnisches Jour-
nal, 242, 1881, 236-243.
9.10 Engineering (London), 21, 1876, March 24, 234-238, May 5, 368 and 370, May 12, 392-393,
May 19, 415-416, May 26, p. 441, June 2, 458-459, and June 9, 482-483.
9.11 Roelker, C.R., Experimental Investigation of the Resistance of Flues to Collapse, Van Nostrand's
Engineering Magazine, 24, March 1881, 208-220.
9.12 Clark, O.K., A Manual of Rules, Tables and Data for Mechanical Engineers, Blackie and Son,
London, 1877, 692-696.
9.13 (Editorial), The Strength of Boiler Flues, The Engineer (London), June 1881, 429.
9.14 Clark, O.K., Strength of Furnace Flues, Engineering (London), 33, Sept. 21, 1881, 280.
9.15 Bach, C., Die auf der kaiserlichen Werft in Danzig von 1887-1892 ausgefiihrten Versuche Uber
die Widerstandsfahikeit von Flammrohren, Zeitschrift des Vereines deutscher lngenieure, 38, No.
23, 9 June 1894, 689-696.
9.16 Bach, C., Untersuchungen iiber die Formanderungen und die Anstrengung gewi.ilbter Boden, Zeit-
schrift des Vereines deutscher lngenieure, 43, (51), December 23, 1899, 1585-1593, (52), Decem-
ber 30, 1899, 1613-1625; and Die Widerstandsfahigkeit kugelfi.irmiger Wandungen gegeniiber
ausserem Ueberdruck, Zeitschrift des Vereines deutscher Ingenieure, 46, (10), March 8, 1902,
333-341.
9.17 Carman, A.P., Resistance of Tubes to Collapse, Physical Review, 21, 1905, 381-387.
9.18 Carman, A.P., and Carr, M.L., Resistance of Tubes to Collapse, University of Illinois Engineering
Experiment Station, Bulletin No. 5, June 1906.
9.19 Steward, R.T., Collapsing Pressure of Bessemer Steel Lap-Welded Tubes, Three to Ten Inches in
Diameter, Transactions of the American Society of Mechanical Engineers, 27, 1906, 730-822.
9.20 Bryan, G.H., Application of the Energy Test to the Collapse of Long Thin Pipe under External
Pressure, Proceedings of the Cambridge Philosophical Society, 6, 1888, 287-292.
9.21 Basset, A.B., On the Difficulties of Constructing a Theory of the Collapse of Boiler Tubes,
Philosophical Magazine, 34, (208), September 1892, 221-233.
9.22 Love, A.E.H., Mathematical Theory of Elasticity, vol. 2, 2nd ed., Cambridge University Press,
1906, 308-316.
9.23 Slocum, S.E., The Collapse of Tubes under External Pressure, Engineering (London), 87, January
1909, 35-37.
9.24 Cook, G., The Resistance of Tubes to Collapse, in Report of Committee of the British Association
for the Advancement of Science on Complex Stress Distributions in Engineering Materials, Report
of B.A.A.S. for 1913, London 1914, 213-224.
9.25 Cook, G., The Collapse of Short Tubes by External Pressure, Philosophical Magazine, 27, 6th
series, July 1914, 51-56.
9.26 SouthwelL R.V., On the Collapse of Tubes by External Pressure, Philosophical Magazine, 25, 6th
series, May 1913, 687-698; 26, 6th series, Sept. 1913, 502-511; 29, 6th series, Jan. 1915, 67-
77.
9.27 Von Mises, R., Der kritische Aussendruck zylindrischer Rohre, Zeitschrift des Vereines deutscher
lngenieure, 58, (19), May 1914, 750-755.
9.28 Carman, A.P., The Collapse of Short Thin Tubes, University of Illinois Engineering Experiment
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley Station, Bulletin No. 99, June, 1917.


Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
792 Shell Buckling Experiments

9.29 Southwell, R.V., On the General Theory of Elastic Stability, Philosophical Transactions of the
Royal Society of London, 213, Series A, 1914, 187-244.
9.30 Cook, G., The Collapse of Short Thin Tubes by External Pressure, Philosophical Magazine, 50,
October 1925, 844-848.
9.31 Von Mises, R., Der kritische Aussendruck fUr allseits belastete zylindrischer Rohre, in: Festschrift,
Prof Dr. A. Stodola, Orell Fiissli Verlag, Zurich, 1929, 418-430.
9.32 Sechler, E.E., The Historical Development of Shell Research and Design, in: Thin-Shell Structures,
Theory, Experiments and Design. Y.C. Fung, and E.E. Sechler, eds., Prentice-Hall, Englewood
Cliffs, N.J. 1974, 3-25.
9.33 Saunders, H.E., and Windenburg, D.F.. Strength of Thin Cylindrical Shells under External Pres-
sure, Transactions of the American Society of Mechanical Engineers, 53, (APM-53-17a), 1931,
207-218.
9.34 Saunders, H.E., and Windenburg, D.F., The Use of Models in Determining the Strength of Thin-
Walled Structures, Transactions of the American Society of Mechanical Engineers, 54, (APM-54-
25), 1932, 263-275.
9.35 Windenburg, D.F.. and Trilling, C., Collapse by Instability of Thin Cylindrical Shells under Ex-
ternal Pressure, Transactions of the American Societv of Mechanical Engineers, 56, (APM-56-
20), 1934, 819-825.
9.36 Jasper, M.T., and Sullivan, J.W.W., The Collapsing Strength of Steel Tubes, Transactions of the
American Society of Mechanical Engineers, 53, (APM-53-17b), 1931,219-245.
9.37 Tokugawa, T., Model Experiments on the Elastic Stability of Closed and Cross-Stiffened Circular
Cylinders under Uniform External Pressure. in: Proceedings, World Engineering Congress, Tokyo,
1929, 29, 1931, 219-279.
9.38 Sturm, R.G., A Study of the Collapsing Pressure of Thin-Walled Cylinders, University of Illinois
Engineering Experiment Station. Bulletin No. 329, November 1941.
9.39 Lilly, W.E., The Strength of Columns, in: Proceedings of Institlltion of Mechanical Engineers,
June 1905, Parts 3 and 4, 627-722.
9.40 Lilly. W.E .. The Economic Design of Columns, Transactions of the institution of Civil Engineers
of Ireland (Dublin), 33, March 1906, 67-93.
9.41 Lilly, WE.. The Design of Struts, Engineering (London), 85, Jan. 10, 1908, 37-40.
9.42 Mallock, A., Note on the Instability of Tubes Subjected to End Pressure, and on the Folds in a
Flexible Material, Proceedings of the Royal Society of London, Series A, 81, 1908, 388-393.
9.43 Mason, W., Mild-Steel Tubes in Compression and under Combined Stress, Proceedings of the
Institution of Mechanical Engineers, Dec. 1909. Part 4, 1205-1236.
9.44 Lorenz, R., Die nicht achsensymmetrische Knickung di.innwandiger Hohlzylinder, Physikalische
Zeitschrijt Leibzig, 12, April 1911. 241-260.
9.45 Timoshenko, S., Einige Stabilittsprobleme der Elastizitatstheorie. Zeitschrift fiir Mathematik und
Physik, 58, 1910, 337-385 (survey of papers published in 1907 in the Bulletin of the Polytechnical
Institute in Kiev).
9.46 Timoshenko, S., Buckling of a Cylindrical Shell Under the Action of Uniform Axial Pressure, in:
Bulletin of the Electrotechnical Institute, St. Petersburg, 11, I 9 I 4 (in Russian).
9.47 Dean, WR., On the Theory of Elastic Stability, Proceedings c~f' the Royal Society of London, 107,
Series A, 1925, 734-759.
9.48 Hoff, N.J., Thin Shells in Aerospace Structures, Aeronautics and Astronautics, 5, Feb. 1967, 26-
45.
9.49 Barling, W.R., and Webb, H.A., Design of Aerospace Struts, The Aeronautical Journal (of the
Royal Aeronautical Society). 22, Oct. 1918, 313-329.
9.50 Popplewell. WC., and Carrington, H., The Failure of Short Tubular Struts of High-Tensile Steel,
Proceedings of the Institution of Civil Engineers, 203, 1917, 381-393.
9.51 Robertson, A., The Strength of Tubular :Struts, Proceedings of the Royal Society of London, 121,
Series A, 1928, 558-585.
9.52 Robertson, A., The Strength of Tubular Struts, Reports and Memoranda, R&M No. 1185, British
A.R.C., July 1929.
9.53 Lundquist, E.E., Strength Tests of Thin-Walled Duralumin Cylinders in Compression, NACA
Report No. 473, June 1933.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,

9.54 Donnell, L.H., A New Theory for the Buckling of Thin Cylinders under Axial Compression and
Bending, Transactions of th American Society of' Mechanical Engineers, 56, (AER-56-12), 1934,
795-806.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 793

9.55 Wilson, W.M., and Newmark, N.M., The Strength of Thin Cylindrical Shells as Columns, Uni-
versity of Illinois Engineering Experiment Station, Bulletin No. 255, Feb. 1933.
9.56 Wilson, W.M., Tests of Steel Columns, University of Illinois Engineering Experiment Station,
Bulletin No. 292, April 1937.
9.57 Wilson, W.M., and Olson, E.D., "Tests of Cylindrical Shells," University of Illinois Engineering
Experiment Station, Bulletin No. 331, Sept. 1941.
9.58 Babcock, C.D., Expe1iments in Shell Buckling, in: Thin-Shell Structures, Theory; Experiments
and Design, Y.C. Fung and E.E. Sechler, eds., Prentice-Hall, Englewood Cliffs, N.J., 1974, 345-
369.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
9.59 Chajes. A .. Stability and Collapse Analysis of Axially Compressed Cylindrical Shells, in: Shell
Structures, Stability and Strength, R. Narayanan, ed., Elsevier Applied Science Publishers, Lon-
don. 1985, l-17.
9.60 Tvergaard, V., Buckling Behavior of Plate and Shell Structures, in: Theoretical and Applied Me-
chanics, Proceedings of 14th IUTAM Congress, Delft, 1976, W.T. Koiter, North-Holland, Am-
sterdam, 1976. 233-247.
9.61 Bushnell, D., Buckling of Shells-Pit Fall for Designers, AIAA Journal, 19, (9), 1981, 1183-
1226.
9.62 Simitses, G.J., Buckling and Postbuckling of Imperfect Cylindrical Shells: A Review, Applied
Mechanics Review, 39, (10), 1986, 1517-1524.
9.63 Miller, C.D., Summary of Buckling Tests on Fabricated Steel Cylindrical Shells in USA, in:
Buckling r!f Shells in Ojj~·/wre Structures, J.E. Harding, P.J. Dowling, and N. Agelidis, eds .. Gra-
nada, London, 1982, 429-471.
9.64 Schulz, U., Der Stabilitatsnachweis bei Schalen, Berichte der Versuchsanstalt fur Stahl, Holz und
Steine der Universitiit Fridericiana in Karlsruhe, 4, (2), 1981.
9.65 Hoff. N.J .. The Perplexing Behavior of Thin Circular Cylindrical Shells in Axial Compression,
Israel Journal of Technolog_v, 4, (I), 1966, 1-28.
9.66 Tennyson, R.C., A Note on the Classical Buckling Load of Circular Cylindrical Shells under Axial
Compression, AIAA Journal, 1, (2), 1963, 475-476.
9.67 Evensen, D.A., High Speed Photographic Observation of the Buckling of Thin Cylinders, Exper-
imental Mechanics, 4, (4), 1964, 110-117.
9.68 Tennyson, R.C., Buckling of Circular Cylindrical Shells in Compression, AIAA Journal, 2, (7),
1964, 1351-1353.
9.69 Almroth, B.O., Holmes. A.M.C., and Brush, D.O .. An Experimental Study of the Buckling of
Cylinders under Axial Compression, Experimental Mechanics, 4, (9), 1964, 263-270.
9.70 Esslinger, M., and Meyer-Piening, H.R., Bericht i.iber das Filmen des Beul- und Nachbeulverhal-
tens von di.innwandingen Kreiszylindern, Deutsche Luft- und Raumfahrt DLR Mit!. 69-07, June
1969.
9.71 Tennyson, R.C .. Buckling Modes of Circular Cylindrical Shells under Axial Compression, AIAA
Journal, 7, (8), 1969, 1481-1487.
9.72 Geier, B., Die Hochgeschwindigkeits-Kinematographie als Hilfsmittel bei der Erforschung von
Beulvorgangen bei di.innwandigen Kreiszylinderschalen, Research Film, 7, (2), 1970, 128-135.
9.73 Esslinger, M., Hochgeschwindigkeitsaufnahmen von Beulvorgang di.innwandiger, axialbelasteter
Zylindcr, Der Stahlbau, 39, (3), 1970, 73-76.
9.74 Esslinger, M., and Geier, B., Postbuckling Behavior of Structures, CSIM Courses and Lectures,
(236), Springer-Verlag, Berlin, New York, 1975, 122-124.
9.75 Horton, W.H., and Durham, S.C., The Effect of Restricting Buckle Depth in Circular Cylindrical
Shells Repeatedly Compressed to the Buckling Limit, SUDAER, (174), Stanford University, Stan-
ford. California, Sept. 1963.
9.76 Horton. W.H., and Durham, S.C., Imperfections, a Main Contributor to Scatter in Experimental
Values of Buckling Load, International Journal of Solids and Structures, 1, 1965, 59-72.
9.77 Horton, W.H., and Bailey, S.C., Influence of Test Machine Rigidity on the Buckling Load of
Shells, in: Test Methods for Compression Members, ASTM STP 419, American Society for Testing
and Materials, 1967, 183-213.
9.78 Fung, Y.C., and Sechler, E.E., Instability of Thin Elastic Shells, in: Structural Mechanics, Pro-
ceedings, of the First Symposium on Naval Structural Mechanics, Stanford University 1958. J.N.
Goodier and N.J. Hoff, eds., Pergamon Press, Oxford, 1960, 115-168.
9.79 Yoshimura, Y., On the Mechanism of Buckling of a Circular Cylindrical Shell under Axial Com-
pression, Reports of the Institute of Science and Technology of the University of Tokyo, 5, (5),
1951 (in Japanese).
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
794 Shell Buckling Experiments

9.80 Yoshimura, Y., On the Mechanism of Buckling of a Circular Cylindrical Shell under Axial
Compression, National Advisory Committee on Aeronautics, NACA TM (1390), 1955.
9.81 Horton, W.H., Johnson, R.W., and Hoff, N.J., Experiments with Thin-Walled Circular Cylindrical
Specimens Subjected to Axial Compression, Appendix I to Buckling of Shells, by N.J. Hoff, in:
Proceedings, of an Aerospace Symposium of Distinguished Lecturers in Honor of Dr. Theodore
von Karman on his 80th Anniversary, Institute of Aerospace Sciences, New York, 1961, 58-68.
9.82 Mossakovskii, V.I., and Smelyi, G.N., Experimental Investigation of the Influence of Testing
Machines on the Stability of Unstiffened Cylindrical Shells under Axial Compression, Mekh.
Mashinost, 4, 1963, 162-166.
9.83 Babcock, C.D., The Influence of the Testing Machine on the Buckling of Cylindrical Shells
under Axial Compression, International Journal of Solids and Structures, 3, 1967, 809-817.
9.84 Carlson, R.L., Sendelbeck, R.L., and Hoff, N.J., Experimental Studies of the Buckling of Com-
plete Spherical Shells, Experimental Mechanics, 7, 1967, 281-288.
9.85 Thompson, J.M.T., Making of Thin Metal Shells for Model Stress Analysis, Journal of Me-
chanical Engineering Sciences, 2, (2), 1960, 105-108.
9.86 Babcock, C.D., and Sechler, E.E., The Effect of Initial Imperfections on the Buckling Stress of
Cylindrical Shells, NASA TN D-2005, 1963.
9.87 Sendelbeck, R.L., The Manufacture of Thin Shells by the Electroforming Process, SUDAER
Report No. 185, Stanford University, Stanford, California, 1964.
9.88 Babcock, C.D., The Buckling of Cylindrical Shells with an Initial Imperfection under Compres-
sion Loading, Ph.D. thesis, California I11stute of Technology, Pasadena, California, 1962.
9.89 Carlson, R.L., Sendelbeck, R.L., and Hoff, N.J., An Experimental Study of the Buckling of
Complete Spherical Shells, Stanford University, Department of Aeronautics and Astronautics,
Report SUDAR No. 254, Dec. 1965; also NASA CR-550, August 1966.
9.90 Berke, L., and Carlson, R.L., Experimental Studies of the Postbuckling Behavior of Complete
Spherical Shells, Experimental Mechanics, 8, 1968, 548-533.
9.91 Arbocz, J., Buckling of Conical Shells under Axial Compression, NASA CR-1162. 1968.
9.92 Singer, J., and Bendavid, D., Buckling of Electroformed Conical Shells under Hydrostatic Pres-
sure, A/AA Journal, 6, 1968, 2332-2338.
9.93 Stuart, F.R., Goto, J.T., and Sechler, E.E., The Buckling of Thin-Walled Circular Cylinders under
Axial Compression and Bending, NASA CR-1160, 1968.
9.94 Horton, W.H., and Craig, J.I., Experimental Studies of the Effect of General Imperfections on
the Elastic Stability of Thin Shells, Israel Journal of Technology, 7, 1969, 91-103.
9.95 Sendelbeck, R.L., and Singer, J., Further Experimental Studies of Buckling of Electroformed
Conical Shells, A/AA Journal, 8, (8), 1970, 1532-1534.
9.96 Parmeter, R.R., Shell Deformation Studies Using Holographic Interferometry, in: Thin-Shell
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Structures, Theory, Experiments and Design, Y.C. Fung and E.E. Sechler, eds., Prentice-Hall,
Englewood Cliffs, N.J., 1974, 371-399.
9.97 Raju, B., and Chandra, R., Some Experiments on Buckling of Electroformed Nickel Shell Spec-
imens, Journal of the Aeronautical Society of India, 28, (1), February 1976, 59-67.
9.98 Morton, J., Murray, P.R., and Ruiz, C., Effect of Imperfections on the Buckling Pressure of
Complete Spherical Shells, in: Stability Problems in Engineering Structures and Components,
T.H. Richards and P. Stanley, eds., Applied Science Publishers, London, 1978. 137-157.
9.99 Waeckel, N., Kabore, A., Cousin, M., and Jullien, J.F., Buckling of Cylindrical Shells, presented
at the 8th International Conference on Structural Mechanics in Reactor Technology (SMIRT),
Brussels, August 1985.
9.100 Waeckel, N., Imperfections geometriques initiales et instabilites de structures minces, doctoral
thesis, INSA LYON/University of Lyon, 1984.
9.101 Debbaneh, N., Jullien, J.F., Reynouard, J.M., and Waeckel, N., Corrugated Cylindrical Shells,
in: Proceedings of ECCS Colloquium on Stability of Plate and Shell Structures, Ghent University,
April 6-8, 1987, P. Dubas and D. Vandepitte, eds., 477-484.
9.102 Jullien, J.F., Procedure de realisation des coques raidies, Contract CRYOSPACE 89, INSA Report
IN-OR/201.113/001, March 15, 1990.
9.103 Weingarten, V.I., Morgan, E.J., and Seide, P., Elastic Stability of Thin-Walled Cylindrical and
Conical Shells under Axial Compression, AIAA Journal, 3, (3), 1965, 500-505.
9.104 Weingarten, V.I., Morgan, E.J., and Seide, P., Elastic Stability of Thin-Walled Cylindrical and
Conical Shells under Combined Intema:t Pressure and Axial Compression, AIAA Journal, 3, (6),
1965, 1118-1125.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 795

9.105 Thielemann, W.F., On the Postbuckling Behavior of Thin Cylindrical Shells, in: Proceedings of
NASA Symposium on Stability of Shell Structures, NASA TN-Dl510, December 1962, 203-216.
9.106 MacCalden, P.B., and Matthiesen, R.B., Combination Torsion and Axial Compression Tests on
Conical Shells, AIAA Journal, 5, 1967, 305-309.
9.107 de Neufville, R.L., and Connor, J.J., Postbuckling Behavior of Thin Cylinders, ASCE Journal of
Engineering Mechanics, 94, 1968, 585-603.
9.108 Ishai, 0., Weller, T., and Singer, J., Anisotropy of Mylar A Sheets, ASTM Journal of Materials,
3, (2), 1968, 337-351.
9.109 Yamaki, N., and Otomo, K., Experiments on the Postbuckling Behavior of Circular Cylindrical
Shells under Hydrostatic Pressure, Experimental Mechanics, 13, (9), July 1973, 299-304.
9.110 Yamaki, N., Otomo, K., and Matsuda, K., Experiments on the Postbuckling Behavior of Circular
Cylindrical Shells under Compression, Experimental Mechanics, 15, (1), January 1975, 23-28.
9.111 Yamaki, N., Experiments on the Postbuckling Behavior of Circular Cylindrical Shells under
Torsion, in: Buckling of Structures, Proceedings, IUTAM Symposium, Harvard University, Cam-
bridge, MA, USA, June 17-21, 1974, B. Budiansky, ed., Springer-Verlag, Berlin, 1976, 312-
330.
9.112 Esslinger, M., and Geier, B., Calculated Postbuckling Loads as Lower Limits for the Buckling
Loads of Thin Walled Circular Cylinder, in Buckling of Structures, Proceedings, IUTAM Sym-
posium, Harvard University, Cambridge, MA, June 17-21, 1974, B. Budiansky, ed., Springer-
Verlag, Berlin, 1976, 274-290.
9.113 Esslinger, M., Klein, H., and Hall, R., Beulen und Nachbeulen dlinwandiger Schalen isotroper
Kreiszylindcr unter Aussendruck und Axiallast, Film B 1183, Institut flir den Wissenschaftlichen
Film, Gottingen, 1975.
9.114 Esslinger, M., Geier, B., and Heidemann, U., Comments on the Paper-Some Complements to
the ECCS Design Code Concerning Isotropic Cylinders, DFVLR, IB 152-77/06, April 1967.
9.115 Geier, B., and Heidemann, U., Experimental Investigations on the Buckling of Thin-Walled
Isotropic Cylinders Subjected to External Hydrostatic Pressure, DFVLR, DLR-FB 77-46, 1977.
9.116 Galletly, G.D., and Pemsing, K., Interactive Buckling Tests on Cylindrical Shells Subjected to
Axial Compression and External Pressure-a Comparison of Experiment, Theory and Various
Codes, Proceedings, Institution of Mechanical Engineers, 199, (C4), 1985, 259-280.
9.117 Tooth, A.S., and Fernandez, J.A., A Study of the Buckling Behavior of Horizontally Supported
Thin-Walled Cylindrical Storage Vessels Which Contain Fluid, in: Stability Problems in Engi-
neering Structures and Components, T.H. Richards and P. Stanley, eds., Applied Science Pub-
lishers, London, 1971, 315-340.
9.118 Toh, S.L., and Spence, J., Some Results on Thin Orthotropic Elliptical Cylindrical Pipes Due to
Combined Bending and Pressure Loading, in: Stability Problems in Engineering Structures and
Components, T.H. Richards and P. Stanley, eds., Applied Science Publishers, London, 1979, 57-
73.
9.119 Toda, S., and Komatsu, K., Buckling of Cantilever Cones Under External Pressure, Experimental
Mechanics, 17, (12), December 1977, 477-480.
9.120 Toda, S., Buckling of Cylindrical Shells with Circular Cutouts under Axial Compression, Na-
tional Aerospace Laboratory, NAL TR-560, Tokyo, 1979 (in Japanese).
9.121 Hobbs, R.E., and Galletly, G.D., The Buckling of a Cylindrical Tank with Varying Wall Thick-
ness, in: Steel Plated Structures, P.J. Dowling, J.E. Harding, and P.A. Frieze, eds., Crosby Lock-
wood Staples, London, 1977, 889-899.
9.122 Adam, H.P., and King, P.A., Experimental Investigation of the Stability of Monocoque Domes
Subject to External Pressure, Experimental Mechanics, 5, 1965, 313-320.
9.123 Adachi, J., and Benicek, M., Buckling of Torispherical Shells under Internal Pressure, Experi-
mental Mechanics, 4, 1964, 217-222.
9.124 Wang, L.R.L., Rodriguez-Agrait, L., and Litle, W.A., Effect of Boundary Conditions on Shell
Buckling, ASCE Journal of Engineering Mechanics, 92, (EM6), December 1966, 101-116.
9.125 Tillman, S.C., On the Buckling Behavior of Shallow Spherical Caps under a Uniform Pressure
Load, International Journal of Solids and Structures, 6, 1970, 37-52.
9.126 Kruger, D.S., and Glockner, P.O., Experiments on the Stability of Spherical and Paraboidal Shells,
Experimental Mechanics, 11, 1971, 254-262.
9.127 Sunakawa, M., and Ichida, K., A High Precision Experiment on the Buckling of Spherical Caps
Subjected to External Pressure, Report 508, Institute of Space and Aeronautical Science, Uni-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

versity of Tokyo, March 1974.


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
796 Shell Buckling Experiments

9.128 Pedersen, P.T., and Jensen, J.J., Correlation between Experimental Buckling Loads and Theo-
retical Predictions for Spherical Cargo Tanks for LNG, in: Steel Plated Structures, An Interna-
tional Symposium, P.J. Dowling, J.E. Harding, and P.A. Frieze, eds., Crosby Lockwood Staples,
London, 1977, 638-657.
9.129 Glockner, P.G., and Tawadros, K.Z., Experiments on Free Vibration of Shells of Revolution,
Experimental Mechanics, 13, 1973, 411-421.
9.130 Tennyson, R.C., An Experimental Investigation of the Buckling of Circular Cylindrical Shells in
Axial Compression Using the Photoelastic Technique, University of Toronto, Institute of Aero-
space Studies, UTIAS Report No. 102, November 1964.
9.131 Tennyson. R.C., Photoelastic Circular Cylinders in Axial Compression, in: Test Methods for
Compression Members, American Society for Testing Materials (ASTM), STP-419, 1967, 31-
46.
9.132 Tennyson, R.C., Booton, M., and Chan, K.H., Buckling of Short Cylinders under Combined
Loading, Journal of Applied Mechanics, 45, (3), Sept. 1978, 574-578.
9.133 Gorman, D., and Evan-Iwanovski, R.M., Photoelastic Analysis of Prebuckling Deformations of
Cylindrical Shells, AIAA Journal, 3, (10), 1965, 1956-1958.
9.134 Veranda, D.R., and Weingarten, V.I., Stability of Hyperboloidal Shells, Journal of the Structural
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Division, ASCE, 101, (ST7), July 1975, 1585-1602.


9.135 Mungan, I., Buckling Stress States of Hyperboloidal Shells, Journal of the Structural Division,
ASCE, 102, (STlO), October 1976, 2005-2020.
9.136 Wurm, P., Modellversuche zum Studium des Beulverhaltens doppelt positiv gekri.immter Schalen,
Der Bauingenieur, 15, (9), Sept. 1975, 397-403.
9.137 Morton, J., Murray, P.R., and Ruiz, C., Buckling of Spherical Storage Tanks under Vertical and
Lateral Acceleration, in: Stability Problems in Engineering Structures and Components, T.H.
Richards and P. Stanley, eds., Applied Science Publishers, London, 1978, 341-354.
9.138 Ross, C.T.F., and Mackney, M.D.A., Deformation and Stability Studies of Thin-Walled Domes
under Uniform External Pressure, Journal of Strain Analysis, 18, (3), 1983, 167-172.
9.139 Ruhwedel, J., Experimentelle Beuluntersuchungen von Ki.ihlturmschalen unter Windbelastung,
Institut fi.ir Konstruktiven Ingenieurbau, Ruhr University Bochum, Technical Report 86-6, August
1986.
9.140 Knapp, R.H., Pseudo-Cylindrical Shells: A New Concept for Undersea Structures, Journal of
Engineering for Industry, ASME, 99, (2), May 1977, 485-492.
9.141 Knapp, R.H., Experimental Investigation of Prebuckled Cylinders under External Pressure, Jour-
nal of Engineering for Industry, ASME, 101, May 1979, 178-184.
9.142 Burns, J.J., Experimental Buckling of Thin Shells of Revolution, ASCE Journal of Engineering
Mechanics, 90, (EM3), 1964, 171-193.
9.143 Penning, F.A., Experimental Buckling Modes of Clamped Shallow Shells under Concentrated
Load, Journal of Applied Mechanics, 3.3, 1966, 297-304.
9.144 Penning, F.A., and Thurston, G.A., The Stability of Spherical Shells under Concentrated Load,
NASA Contractor Report CR-265, July 1965.
9.145 Homewood, P.H., Brine, A.C., and Johnson, A.E., Jr., Experimental Investigation of the Buckling
Instability of Monocoque Shells, Experimental Mechanics, 1, 1961, 88-96.
9.146 Weingarten, V.J., Vibration of Conical Shells Subjected to Torsion, ASCE Journal of Engineering
Mechanics, 94, EMl, 1968, 47-56.
9.147 Dumesnil, C.E., and Nevill, G.E., Jr., Deformation of Spherical Caps Impacted into Water, AIAA
Journal, 5, 1967, 1043-1045.
9.148 Guist, L.R., Buckling Load of Thin Circular Cylindrical Shells Formed by Plastic Expansion,
NASA TN D-6322, April 1971.
9.149 Weller, T., and Singer, J., Experimental Studies on Buckling of Ring Stiffened Conical Shells
under Axial Compression, Experimental Mechanics, 10, 1970, 449-457.
9.150 Krenzke, M.A., Tests of Machined Deep Spherical Shells under External Hydrostatic Pressure,
Report 1601, Navy Department, David Taylor Model Basin, 1962.
9.151 Krenzke, M.A., and Kiernan, T.J., Elastic Stability of Near Perfect Shallow Spherical Shells,
AIAA Journal, 1, 1963, 2855-2857; erratum, AIAA Journal, 2, 1964, p. 784.
9.152 Krenzke, M.A., and Kiernan, T.J., The Effect of Initial Imperfections on the Collapse Strength
of Deep Spherical Shells, Report 1757, Navy Department, David Taylor Model Basin, 1965.
9.153 Nishida, K., The Inelastic Buckling of Machined High Strength Steel Hemispheres with Ideal
Boundaries, Report 2090, Navy Department, David Taylor Model Basin, 1965.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 797

9.154 Blachut, J., Galletly, G.D., and Moreton, D.N., Buckling of Near-Perfect Steel Torispherical and
Hemispherical Shells Subjected to Pressure, AIAA Journal, 28, (11 ), 1990, 1971-1975.
9.155 Steinhardt, 0., and Schultz, U., Zum Beulverhalten von Kreiszylinderschalen, Schweizerische
Bauzeitung, 89, (1), January 1971, 1-14.
9.156 Strenkowski, J.S., and Witmer, E.A., Buckling and Postbuckling Studies of Torsion-Loaded Thin-
Walled Cylindrical Shells with Cutouts, Report AMMRC CTR 73-39, Army Materials and Me-
chanics Research Center, Watertown, Mass., October 1973.
9.157 Almroth, B.O., and Holmes, A.M.C., Buckling of Shells with Cutouts, Experiment and Analysis,
International Journal of Solids and Structures, 8, 1972, 1057-1071.
9.158 Galletly, G.D .. Aylward, R.W., and Bushnell, D., An Experimental and Theoretical Investigation
of Elastic and Elastic-Plastic Asymmetric Buckling of Cylinder-Cone Combinations Subjected
to Uniform External Pressure, Ingenieur Archiv, 43, 1974, 345-358.
9.159 Aylward, R.W., Galletly, G.D., and Moffat, D.G., Buckling under External Pressure of Cylinders
with Toriconical or Pierced Torispherical Ends-A Comparison of Experiment and Theory, Jour-
nal Mechanical Engineering Science, 17, 1975, 11-18.
9.160 Galletly, G.D., and Aylward, R.W., The Influence of End Closure Shape on the Buckling of
Cylinders under External Pressure, Transactions of Royal Institution of Naval Architects, 117,
1975, 255-264.
9.161 Galletly, G.D., On the Buckling of Shallow Spherical Caps Subjected to Uniform External Pres-
sure, AIAA Journal, 14, 1976, 1331-1333.
9.162 Galletly, G.D., Elastic and Elastic-Plastic Buckling of Internally Pressurized 2:1 Ellipsoidal
Shells, Trans. ASME Journal of Pressure Technology, 100, Nov. 1978, 335-343.
9.163 Galletly, G.D., Buckling and Collapse of Thin Internally Pressurized Dished Ends, Proceedings,
Institution of Civil Engineers, 67, (2), Sept. 1979, 607-626.
9.164 Galletly, G.D., On Elastic and Elastic-Plastic Asymmetric Buckling of Pressurized Combinations
of Thin Shells, Der Stahlbau, 48, (11), 1979, 340-346.
9.165 Kirk, A., and Gill, S.S., The Failure of Torispherical Ends of Pressure Vessels Due to Instability
and Plastic Deformation-An Experimental Investigation, International Journal of Mechanical
Sciences, 17, 1975, 525-544.
9.166 Patel, P.R., and Gill, S.S., Experiments on the Buckling under Internal Pressure of Thin Tori-
spherical Ends of Cylindrical Pressure Vessels, International Journal of Mechanical Sciences,
20, 1978, 159-175.
9.167 Verduyn, W.D., and Elishakoff, I., A Testing Machine for Statistical Analysis of Small Imperfect
Shells, in: Proceedings of the 7th International Conference on Experimental Stress Analysis,
Haifa, Israel, August 23-27, 1982, A. Betser, ed., Ayalon Press, Haifa, 1982, 545-557.
9.168 Harris, L., Suer, H.S., Skene, W.T., and Benjamin, R.J., The Stability of Thin-Walled Unstiffened
Circular Cylinders Under Axial Compression Including the Effect of Internal Pressure, Journal
of the Aeronautical Sciences, 24, (8), August 1957, 587-606.
9.169 Miller, C.D., Buckling of Axially Compressed Cylinders, Journal of the Structural Division,
ASCE, 103, (ST3), March 1977, 695-721.
9.170 Ostapenko, A., Local Buckling of Welded Tubular Columns, in: Stability of Structures under
Static and Dynamic Loads, American Society of Civil Engineers (ASCE), New York, 1977, 367-
374.
9.171 Ostapenko, A., and Gunzelman, S.X., Local Buckling of Tubular Steel Columns, in: Methods of
Structural Analysis, 3, W. Saul and A. Peyrot, eds., American Society of Civil Engineers (ASCE),
New York, 1976, 549-568.
9.172 Ostapenko, A., and Gunzelman, S.X., Local Buckling Tests on Three Large- Diameter Tubular
Columns, in: Recent Research and Developments in Cold Formed Steel Structures, Department
of Civil Engineering, University of Missouri/Rolla, June 1978, 409-463.
9.173 Marzullo, M.A., and Ostapenko, A., Tests on Two High-Strength Short Tubular Columns, OTC
Paper 2086, lOth Offshore Technology Conference, Houston, Texas, May 8-11, 1978.
9.174 Ostapenko, A.. and Grimm, D.F., Local Buckling of Cylindrical Tubular Columns Made of A-
36 SteeL Lehigh University, Fritz Engineering Laboratory Report (450.7), February 1980.
9.175 Smith, C.S., Kirkwood, W., and Swan, J.W., Buckling Strength and Post- Collapse Behavior of
Tubular Bracing Members Including Damage Effects, in: Boss '79, Proceedings, 2nd Interna-
--`,`,`````,`,````,,``,``,,`-`-`

tional Conference on Behavior of Off-Shore Structures, Imperial College, London, 1979, 2, 303-
326.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
798 Shell Buckling Experiments

9.176 Smith, C.S., Strength and Stiffness of Damaged Tubular Beam Columns, in: Buckling of Shells
in Offshore Structures, J.E. Harding, P.J. Dowling, and N. Agelidis, eds., Granada, London, 1982,
1-23.
9.177 Frum, Y., and Baruch, M., Buckling of Cylindrical Shells Heated along Two Opposite Generator
Combined with Axial Compression, Experimental Mechanics, 16, (4), 1976, 133-139.
9.178 Ari-Gur, J., Baruch, M., and Singer, J., Buckling of Cylindrical Shells under Combined Axial
Preload Nonuniform Heating and Torque, Experimental Mechanics, 19, (11), 1979,406-410.
9.179 Belov, V.K., Experimental Study of the Stability of Heated and Loaded Shells, Aviatsionnaia
Tekhnika, 21, (2), 1978, 5-8 (in Russian).
9.180 Jung, 0., Der Einfluss von Uingsversteiften Ausschnitten auf das Beulverhalten von Zylinder-
schalen, in: Proceedings, Hans Ebner Geddchtnis Kolloquim in Aachen, October 1977, Mitteilung
aus dem Institut fUr Leichtbau, RWTH Aachen, April 1978, 240-267.
9.181 Mossakovskii, V.I., Andreev, L.V., Obodan, N.I., and Patsink. A.G., On the Local Stability of
Cylindrical Shells Loaded by a Concentrate Force, Akademia Nauk SSR, Doklady, 225, November
1975, 517-519 (in Russian).
9.182 Odland, J., and Didriksen, T., Buckling Experiments with Spherical Shell Segments--Part I, Test
Equipment, Report No. 77-442, Det norske Yeritas, 1977.
9.183 Odland, J., and Steen, E., Buckling Experiments with Spherical Shell Segments-Part II, Testing
and Measurements, Report No. 77-532, Det norske Yeritas, 1977.
9.184 Harris, L.A., Suer, H.S., and Skene, W.T., The Effect of Internal Pressure on the Buckling Stress
of Thin-Walled Circular Cylinders under Combined Axial Compression and Torsion, Journal of
the Aeronautical Sciences, 25, (2), February 1958, 142-143.
9.185 Suer, H.S., Harris, L.A., Skene, W.T., and Benjamin, R.J., The Bending Stability of Thin-Walled
Unstiffened Circular Cylinders Including the Effects of Internal Pressure, Journal of the Aero-
nautical Sciences, 25, (5), May 1958, 281-287.
9.186 Suer, H.S., and Harris, L.A., The Stability of Thin-Walled Cylinders under Combined Torsion
and External Lateral or Hydrostatic Pressure, Journal of Applied Mechanics, 26, (1), March
1959, 138-140.
9.187 Harris, L.A., Suer, H.S., and Skene, W.T, Model Investigations of Unstiffened and Stiffened
Circular Shells, Experimental Mechanics, 1, (7), July 1961, 1-9.
9.188 Alnuoth, B.O., Burns, A.B., and Pittner, E.Y., Design Criteria for Axially Loaded Cylindrical
Shells, Journal of Spacecraft and Rockets, 7, (6), June 1970, 714-720.
9.189 Thielemann, W.F., and Esslinger, M.E., On the Postbuckling Behavior of Thin-Walled, Axially
Compressed Circular Cylinders of Finite Length, in: Proceedings-Symposium on The Theory
of Shells, University of Houston, Texas. 1967, 433-479.
9.190 Yamaki, N., and Tani, J., Postbuckling Behavior of Circular Cylindrical Shells under Hydrostatic
Pressure, ZAMM, 54, 1974, 709-714.
9.191 Yamaki, N., and Kodama, S., Postbuckling Behavior of Circular Cylindrical Shells under Com-
pression, International Journal of Non-Linear Mechanics, 11, 1976, 99-111.
9.192 Sendelbeck, R.S., and Hoff, N.J., Loading Rig in Which Axially Compressed Thin Cylindrical
Shells Buckle Near Theoretical Values, Experimental Mechanics, 12, (8), 1972, 372-376.
9.193 Okubo, S., Wilson, P.E., and Whittier, J.:S., Influence of Concentrated Lateral Loads on the Elastic
Stability of Cylinders in Bending, Experimental Mechanics, 10, (9), 1970, 384-389.
9.194 Arbocz, J., and Babcock, C.D., The Effect of General Imperfections on the Buckling of Cylin-
drical Shells, Journal of Applied Mechanics, 36, Series E, (1), 1969, 28-38.
9.195 Tennyson, R.C., and Muggeridge, D.B., Buckling of Axisymmetric Imperfect Circular Cylindri-
cal Shells under Axial Compression, AlAA Journal, 7, (11), Nov. 1969, 2127-2131.
9.196 Hoff, N.J., Some Recent Studies of the Buckling of Thin Shells, Aeronautical Journal of the
Royal Aeronautical Society, 73, (708), December 1969, 1057-1070.
9.197 Klein, H., Beulanlagen im Institut fUr Structurmechanik, DLR Braunschweig, Internal Report IB
131-89/35, December 1989.
9.198 Klein, H., General about Buckling Tests with Thin-Walled Shells, DLR-Mitteilung 89-13, Braun-
schweig, April 1989.
9.199 Limam, A., Jullien, J.F., Greco, E., and Lestrat, D., Buckling of Thin-Walled Cylinders Under
Axial Compression and Internal Pressure, in: Buckling of Shell Structures on Land, in the Sea
and in the Air, J.F. Jullien, ed., Elsevier Applied Science, London and New York, 1991, 359-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

369.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 799

9.200 Aflak, W.F., Flambage plastique de coques cylindriques sous compression axiale. Influence des
imperfections geometriques et des imperfections de conditions aux limites, doctoral thesis, INSA,
Lyon, July 1988.
9.201 Singer, J ., and Eckstein, A., Experimental Investigations of the Instability of Conical Shells under
External Pressure, Bulletin of the Research Council of Israel, llC, (1 ), April, 1962, 97-22.
9.202 Tenerelli, D.J., and Horton, W.H., An Experimental Study of the Local Buckling of Ring-
Stiffened Cylinders Subject to Axial Compression, Israel Journal of Technology, 7, (1-2), 1969,
181-194.
9.203 Feinstein, G., Erickson, B., and Kempner, J., Stability of Oval Cylindrical Shells, Experimental
Mechanics, 11, 1971, 514-520.
9.204 Kempner, J., and Chen, Y.-N., Buckling and Postbuckling of an Axially Compressed Oval Cy-
lindrical Shell, in: Proceedings-Symposium on the Theory of Shells to Honor Lloyd Hamilton
Donnell, University of Houston, Texas, D. Muster, ed., McCutchan Publishing Corporation, 1967,
141-183.
9.205 Kempner, J., and Chen, Y.-N., Postbuckling of an Axially Compressed Oval Cylindrical Shell,
in: Applied Mechanics, Proceedings of the 12th International Congress of Applied Mechanics,
Stanford University, 1968, M. Hetenyi and W.G. Vincenti, eds., Springer-Verlag, Berlin, 1969,
246-256.
9.206 Singer, J., The Influence of Stiffener Geometry and Spacing on the Buckling of Axially Com-
pressed Cylindrical and Conical Shells, TAE Report 68, Department of Aeronautical Engineering,
Technion, Haifa, Israel, October 1967.
9.207 Zukas, J.A., Nicholas, T., Swift, H.F., Greszcuk, L.B., and Curran, D.R., Impact Dynamics, John
Wiley & Sons, New York, 1982, 241-275.
9.208 Tennyson, R.C., Tulk, J.D., and Ricciatti, R., Analysis of the Collapse of Cylindrical Shells
Using High-Speed Photography, Journal of the Society of Motion Picture and Television Engi-
neers, 80, (6), 1971, 477-481.
9.209 Zimcik, D.G., and Tennyson, R.C., Stability of Circular Cylindrical Shells under Transient Axial
Impulse Loading, AIAA Journal, 18, (6), 1980, 691-699.
9.210 Lindberg, H.E., and Herbert, R.E., Dynamic Buckling of a Thin Cylindrical Shell under Axial
Impact, Journal ofApplied Mechanics, 33, 1966, 105-112.
9.211 Yamada, M., and Yamada, S., Agreement between Theory and Experiment on Large Deflection
Behavior of Clamped Shallow Spherical Shells under External Pressure, in: Collapse: The Buck-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

ling of Structures in Theory and Practice, J.M.T. Thompson, and G.W. Hunt, eds., Cambridge
University Press, 1983, 431-441.
9.212 Singer, J., The Influence of Stiffener Geometry and Spacing on the Buckling of Axially Com-
pressed Cylindrical and Conical Shells, in: Theory of Thin Shells, Proceedings of 2nd IUTAM
Symposium on Theory of Thin Shells, Copenhagen, September 1967, F.E. Niordson, ed., Springer-
Verlag, Berlin, 1969, 234-263.
9.213 Horton, W.H., Nassar, E.M., and Singhal, M.K., Determination of the Critical Loads of Shells
by Nondestructive Methods, Experimental Mechanics, 17, (4), 1977, 154-160.
9.214 Galletly, G.D., and Reynolds, T.C., A Simple Extension of Southwell's Method for Determining
the Elastic General Instability Pressure for Ring-Stiffened Cylinders Subject to External Hydro-
static Pressure, Proceedings of the Society for Experimental Stress Analysis, 13, (2), 1956, 141-
152.
9.215 Horton, W.H., and Cundari, F.L., On the Applicability of the Southwell Plot to the Interpretation
of Test Data from Instability Studies of Shell Bodies, in: Proceedings of the 8th AIAA/ASME
Structures, Structural Dynamics and Materials Conference, Palm Springs, California, 1967,651-
660.
9.216 Bank, M.H., The Effect of Fiber Direction on the Stability of Single-Layer Resin Impregnated
Glass Cloth Cylinders under Torsion, Engineer's thesis, Stanford University, Stanford, California,
1966.
9.217 Horton, W.H., On the Elastic Stability of Shells, Final Report NASA Grant NRG 11-002-086,
Georgia Institute of Technology, Atlanta, GA, June 1976.
9.218 Lebouc, P.J.J., Instability of Spherical Caps and Complete Spheres, U.S. Army Air Mobility
Research and Development Laboratory, Fort Eustis, Virginia, USAAVLABS TR 69-61, Septem-
ber 1971.
9.219 Craig, .T.I., and Duggan, M.F., Nondestructive Shell-Stability Examination by a Combined-
Loading Technique, Experimental Mechanics, 13, (9), 1973, 381-387.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
800 Shell Buckling Experiments

9.220 Roorda, J .. Concepts in Elastic Structural Stability, in: Mechanics Today, S. Nemat-Nasser, ed.,
1, 1972, 322-372.
9.221 Arbocz, J., and Babcock, C. D., Experimental Investigation of the Effect of General Imperfections
on the Buckling of Cylindrical Shells, NASA CR-1163. 1968.
9.222 Weller, T., Singer, J., and Nachmani, S., Recent Experimental Studies on the Buckling of
Stringer-Stiffened Cylindrical Shells, TAE Report 100, Department of Aeronautical Engineering,
Technion, Haifa, Israel, April 1970.
9.223 Kaplan, A., and Fung, Y.C., A Non-Linear Theory of Bending and Buckling of Thin Elastic
Shallow Spherical Shells, NACA TN-3212, 1954.
9.224 Kaplan, A., Buckling of Spherical Shells, in: Thin-Shell Structures: Theory, Experiment and
Design, Y.C. Fung and E.E. Sechler, eds., Prentice-Hall, Englewood Cliffs, N.J., 1974, 247-288.
9.225 Thompson, J.M.T., The Elastic Instability of a Complete Spherical Shell, Aeronautical Quarterly,
13, 1962, 189-201.
9.226 Ashwell, D.G., On the Large Deflection of a Spherical Shell with an Inward Point Load, in:
Proceedings of JUTAM Symposium on the Theory of Thin Elastic Shells, W.T. Koiter, ed., North-
Holland, Amsterdam, 1960, 43-63.
9.227 Evan-Iwanowski, R.M., Cheng, H.S., and Loo, T.C., Experimental Investigations of Deformations
and Stability of Spherical Shells Subjected to Concentrated Load at the Apex, in: Proceedings
of 4th U.S. National Congress of Applied Mechanics, 1962, 563-575.
9.228 Weller, T., and Singer, J., Experimental Studies on Buckling of 7075-T6 Aluminum Alloy In-
tegrally Stringer-Stiffened Shells, TAE Report 135, Department of Aeronautical Engineering,
Technion, Haifa, Israel, October 1971.
9.229 Weller, T., and Singer, J., Further Experimental Studies on Buckling of Ring-Stiffened Conical
Shells under Axial Compression, TAE Report 70. Department of Aeronautical Engineering, Tech-
nion, Haifa, Israel, October 1968.
9.230 Foster, C. G., and Tennyson, R.C., Use of the Southwell Method to Predict Buckling Strength of
Stringer Stiffened Cylindrical Shells, Strain-The Journal of the British Societ_vfor Strain Mea-
surement, 19, May 1983, 63-67.
9.231 Royles, R., and Llambias, J.M., Buckling Behavior of an Underwater Storage Vessel, Experi-
mental Mechanics, 25, 1985, 421-428.
9.232 Valsangkar, A.J., Britto, A.M., and Gunn, M.J., Application of the Southwell Plot Method to the
Inspection and Testing of Buried Flexible Pipes, in: Proceedings, Institution of Civil Engineers,
Part 2, 71, 1981, 63-82.
9.233 Singer, J., The Effect of Axial Constraint on the Instability of Thin Cylindrical Shells under
External Pressure, Journal of Applied Mechanics, 27, (4), 1960, 737-739.
9.234 Sobel, L.H., Effects of Boundary Conditions on the Stability of Cylinders Subject to Lateral and
Axial Pressure, AIAA Journal, 2, (8), 1964, 1437-1440.
9.235 Thielemann, W., and Esslinger, M., Einfluss der Randbedingungen auf die Beullast von Kre-
iszylinderschalen, Der Stahlhau, 33, (12), 1964, 353-361.
9.236 Kirstein, A.F., and Wenk, E., Observations of Snap-Through Action in Thin Cylindrical Shells
under External Pressure, Proceedings of the Society for Experimental Stress Analvsis, 14, (l),
1956, 205-214.
9.237 Weingarten, V.I., and Seide, P., Elastic Stability of Thin-Walled Cylindrical and Conical Shells
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

under Combined External Pressure and Axial Compression, AIAA Journal, 3, 1965, 913-920.
9.238 Stracke, M., and Schmidt, H., Beulversuche an liingsnahtgeschweissten stablemen Kreiszylin-
derschalen unter Aussendruck im elastisch-plastischen Bereich, Forschungsbericht aus dem Fach-
bereich Bauwesen 28, University Gesamthochschule Essen, Dec. 1984.
9.239 Schmidt, H., and Stracke, M., Buckling Strength and Postbuckling Behavior of Short Cylindrical
Shells under External Pressure in the Elastic-Plastic Region, in: Proceedings of ECCS Collo-
quium on Stability of Plate and Shell Structures, Ghent University, April 6-8, 1987, P. Dubas
and D. Vandepitte, eds., 467-476.
9.240 Lundquist, E.E., Strength Tests of Thin-Walled Duralumin Cylinders in Pure Bending, NACA
TN 479, 1933.
9.241 Chwalla, E., Reine Biegung schlanker, diinnwandiger Rohre mit gerader Achse, Zeitschrift fiir
angewandete Mathematik und Mechanik, 13, (1), 1933, 48-53.
9.242 Fabian, 0., Collapse of Cylindrical, Elastic Tubes under Combined Bending, Pressure and Axial
Loads, International Journal of Solids and Structures, 13, 1977, 1257-1270.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 801

9.243 Ju, GT, and Kyriakides, S., Bifurcation Buckling versus Limit Load Instabilities of Elastic-
Plastic Tubes under Bending and External Pressure, Journal of Offshore Mechanics and Arctic
Engineering, Transactions ASME, 113, February 1991, 43-52.
9.244 Axelrad. E.L., Flexible Shells, in: Theoretical and Applied Mechanics, Proceedings of the 15th
IUTAM Congress, Toronto, 1980, F.P.J. Rimrott and B. Tabarrok, eds., North-Holland, Amster-
dam 1980, 45-56.
9.245 Mossman, R.W., and Robinson, R.G., Bending Tests of Metal Monocoque Fuselage Construction,
NACA TN 357, 1930.
9.246 Imperial, F.F., The Criterion of Elastic Instability of Thin Duralumin Tubes Subjected to Bending,
M.S. thesis, Department of Mechanical Engineering, University of California, 1932.
9.247 Peterson, J.P., Bending Tests of Ring-Stiffened Circular Cylinders, NACA TN 3735, July 1956.
9.248 Esslinger, M., and Geier, B., Buckling and Postbuckling Behavior of Conical Shells Subjected
to Axisymmetric Loading and of Cylinders Subjected to Bending, in: Theory of Shells, W.T.
Koiter and G.K. Mikhailov, eds., North-Holland, Amsterdam, 1980, 263-288.
9.249 Peterson, J.P., and Anderson, J.K., Bending Tests of Large-Diameter Ring-Stiffened Corrugated
Cylinders, NASA TN D-3336, March 1966.
9.250 Anderson, J.K., Bending Tests of Two Large-Diameter Corrugated Cylinders With Eccentric Ring
Sti!Ieners, NASA TN D-3702, Nov. 1966.
9.251 Anderson, J.K., and Peterson, J.P., Buckling Tests of Two Integrally Stiffened Cylinders Sub-
jected to Bending, NASA TN D-6271, June 1971.
9.252 Dow, M.B., and Peterson, J.P., Bending and Compression Tests of Pressurized Ring-Stiffened
Cylinders, NASA TN D-360, April 1960.
9.253 Kyriakides, S., and Shaw, P.K., Inelastic Buckling of Tubes under Cyclic Bending, ASME Journal
of Pressure Vessel Technology, 109, 1987, 169-178.
9.254 Shaw, P.K., and Kyriakides, S., Inelastic Analysis of Thin-Walled Tubes under Cyclic Bending,
International Journal of Solids and Structures, 21, (11), 1985, 1073-1100.
9.255 Corona, E., and Kyriakides, S., On the Collapse of Inelastic Tubes under Combined Bending
and Pressure, International Journal of Solids and Structures, 24, (5), 1988, 505-535.
9.256 Lundquist, E.E., Strength Tests on Thin-Walled Duralumin Cylinders in Torsion, NACA TN 427,
1932.
9.257 Sezawa, K., and Kubo, K., The Buckling of a Cylindrical Shell under Torsion, Tokyo Imperial
University Aeronautical Research Institute, Report No. 76, 6, (10), 1931, 251-314.
9.258 Moore, R.L., and Westcoat, C., Torsion Tests of Stiffened Circular Cylinders, NACA WR W-89,
1944 (formerly NACA ARR 4E3!).
9.259 Batdorf, S.B., Skin, M., and Schildcrout, M., Critical Stress of Thin-Walled Cylinders in Torsion,
NACA TN 1344, 1947.
9.260 Gerard, G., and Becker, H., Handbook of Structural Stability, Part III-Buckling of Curved Plates
and Shells, NACA TN 3783, August 1957.
9.261 Stang, A.H., Ramberg, W., and Back, G., Torsion Tests of Tubes, NACA Report 601, 1937.
9.262 Nash, W.A., An Experimental Analysis of the Buckling of Thin Initially Imperfect Cylindrical
Shells Subject to Torsion, Proceedings of the Society of Experimental Stress Analysis, 16, 1959,
55-68.
9.263 Kodama, S., Otomo, K., and Yamaki, N., Postbuckling Behavior of Pressurized Circular Cylin-
drical Shells under Torsion-I. Experiment, International Journal of Non-Linear Mechanics, 16,
(3/4), 1981, 337-353.
9.264 Weingmten, V.I., The Effect of Internal Pressure and Axial Tension on the Buckling of Cylin-
drical Shells under Torsion, in: Proceedings of the 4th U.S. National Congress of Applied Me-
chanics, 1962, 827-842.
9.265 Kodama, A .. and Yamaki, N., Postbuckling Behavior of Pressurized Circular Cylindrical Shells
under Torsion-H. Theory, International Journal of Non-Linear Mechanics, 16, (3/4), 1981,
355-370.
9.266 Ekstrom, R.E., Buckling of Cylindrical Shells under Combined Torsion and Hydrostatic Pressure,
Experimental Mechanics, 3, (8), 1963, 192-197.
9.267 Foster, C. G., Interaction of Buckling Modes in Thin-Walled Cylinders, Experimental Mechanics,
21, 1981, 124-128.
9.268 Galletly, G.D., and Pemsing, K., Buckling of Cylinders under Combined External Pressure and
Axial Compression, in: Collapse, the Buckling of Structures in Theory and Practice, J.M.T.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Thompson and G.W. Hunt, eds., Cambridge University Press, Cambridge, 1983, 505-527.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
802 Shell Buckling Experiments

9.269 Booton, M., and Tennyson, R.C., Buckling of Imperfect Anisotropic Circular Cylinders under
Combined Loading, AIAA Journal, 17, (3), 1979, 278-287.
9.270 Singer, J., On Experimental Technique for Interaction Curves of Buckling of Shells, Experimental
Mechanics, 4, (9), 1964, 279-280.
9.271 Berkovits, A., and Singer, J., Buckling of Unstiffened Conical Shells under Combined Torsion
and Axial Compression or Tension, Israel Journal of Technology, 3, (1), 1965, 15-24.
9.272 Berkovits, A., Singer, J., and Weller, T., Buckling of Unstiffened Conical Shells under Combined
Loading, Experimental Mechanics, 7, (11), 1967, 458-467.
9.273 Abramovich, H., Weller, T., and Singer, J., Effect of Sequence of Combined Loading on Buckling
of Stiffened Shells, Experimental Mechanics, 28, (1), 1988, 1-13.
9.274 Abramovich, H., Singer, J., and Weller, T., The Influence of Initial Imperfections on the Buckling
of Stiffened Cylindrical Shells under Combined Loading, in: Buckling of Shell Structures, on
Land, in the Sea and in the Air, J.F. Julien, ed., Elsevier Applied Science, London, 1991, 1-10.
9.275 Singer, J., and Eckstein, A., Recent Experimental Studies of Buckling of Conical Shells Under
Torsion and External Pressure, in: Proceedings of the Fifth Israel Conference on Aviation and
Astronautics, Jerusalem Academic Press, 1963, 135-146.
9.276 Seide, P., On the Buckling of Truncated Conical Shells Under Uniform Hydrostatic Pressure, in:
Proceedings of IUTAM Symposium on the Theory of Thin Elastic Shells, Delft, August 1959,
W.T. Koiter, ed., North-Holland, Amsterdam, 1960, 363-388.
9.277 Singer, J., Correlation of the Critical Pressure of Conical Shells with That of Equivalent Cylin-
drical Shells, AIAA Journal, 1, (11), Nov. 1963, 2675-2676.
9.278 Singer, J., Buckling of Orthotropic and Stiffened Conical Shells, in: Collected Papers on Insta-
bility of Shell Structures, 1962, NASA-TN-D-1510, December 1962, 463-479.
9.279 Singer J., On the Buckling of Unstiffened, Orthotropic and Stiffened Conical Shells, presented
at the 7th A.F.I.T.A.E. International Aeronautical Congress, Paris, June 1965, 1-22.
9.280 Singer, J., Buckling of Clamped Conical Shells under External Pressure, AIAA Journal, 4, (2),
Feb. 1966, 328-337.
9.281 Tani, J., Influence of Prebuckling Deformations on the Buckling of Truncated Conical Shells
under Axial Compression, Transactions, Japan Society for Aeronautical and Space Sciences, 16,
(34), 1973, 232-245.
9.282 Esslinger, M., and Ciprian, J., Buckling of Thin Conical Shells under Axial Loads with or without
Internal Pressure, in: Buckling of Shells, Proceedings of the State-of-the-Art Colloquium, Stuttgart
University, Germany, May 6-7, 1982, E. Ramm, ed., Springer-Verlag, Berlin, Heidelberg, New
York, 1982, 355-374.
9.283 Tokugawa. T., Experiments on the Elastic Stability of a Thin- Wall Cone under Uniform Normal
Pressure on All Sides, and an Approximate Method for Computing its Collapsible Pressure,
Congress of Applied Mechanics League, Z6sen Ky6kai (Shipbuilding Association), Miscella-
neous Publications, (125), 1932, 151-169 (in Japanese).
9.284 Magula, A.W., Structural Test-Conical Head Assembly, Test No. 815, North American Aviation
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Inc., Downey, Missile Development Division, Report MTL 531, 1954.


9.285 Jordan, W.D., Buckling of Thin Conical Shells under Uniform External Pressure, Bureau of
Engineering Research, College of Engineering, University of Alabama, Alabama, February 1955,
ASTIA AD-55384.
9.286 Westmoreland, R.T., Test of Model Conical Bulkhead-Test 1098, North American Aviation
Inc., Downey, Missile Development Division, Report MTL 652, May 1956.
9.287 Bowie, O.L., Parker, B.S., Radkowski, P.O., and Bluhm, J.I., A Study of the IRBM (Jupiter)
Bulkheads, Watertown Arsenal Laboratories Report No. 880-54, August 1956.
9.288 Shroeder. F.J., Kusterer, E.T., and Hirsch, R.A., An Experimental Determination of the Stability
of Conical Shells, Aircraft Armaments, Inc., Cockeyesville, Md., Report ER-1361, May 1958.
9.289 Lundquist, E.E., and Schuette, E.H., Strength Tests of Thin-Wall Truncated Cones of Circular
Section, NACA (ARR) WR L-442, December 1942.
9.290 Lofblad, R.P., Elastic Stability of Thin-Walled Cylinders and Cones with Internal Pressure under
Axial Compression, Aeroelastic and Structures Research Laboratory, Massachusetts Institute of
Technology, Cambridge, Mass., Technical Report 25-29, May 1959.
9.291 Lackman, L., and Penzien, J., Buckling of Circular Cones Under Axial Compression, Journal
of Applied Mechanics, 27, 1960, 458-460.
9.292 Schnell, W., and Schiffner, K., Experimentelle Untersuchungen des Stabilitiitsverhaltens von
diinnwandigen Kegelschalen unter Axiallast und Innendruck, DVL (now DLR), Report 243,
November 1962.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 803

9.293 Lutwak, S., Semiannual Report on Development of Design Criteria for Elastic Instability of Thin
Shell Structures, Space Technology Laboratories Report STL/TR-59-0000-00715, June 1959.
9.294 Seide, P., Weingarten, V.I., and Morgan, E.J., Final Report on the Development of Design Criteria
for Elastic Stability of Thin Shell Structures, Space Technology Laboratories Report STL/TR-
60-0000-19425, Dec. 1960.
9.295 Seide, P., Survey of Buckling Theory and Experiment for Circular Conical Shells of Constant
Thickness, in: Collerted Papers on Instability of Shell Structures-I962, NASA TN D-1510,
Dec. 1962.
9.296 Baruch, M., Harari, 0., and Singer, J., Influence of In-Plane Boundary Conditions on the Stability
of Conical Shells under Hydrostatic Pressure, Proceedings of the 9th Israel Annual Conference
on Aviation and Astronautics, Israel Journal of Technolog); 5, (1), February 1967, 12-24.
9.297 Singer, J., Baruch, M., and Reichenthal, J., Influence of In-Plane Boundary Conditions on the
Buckling of Clamped Conical Shells, Proceedings of the 13th Annual Conference on Aviation
and Astronautics, Israel Journal of Technology, 9, (1-2), March 1971, 127-139.
9.298 Singer, J., Eckstein, A., and Baruch, M., Buckling of Conical Shells Under External Pressure,
Torsion and Axial Compression, TAE Report 19, Department of Aeronautical Engineering, Tech-
nion, Haifa, Israel, Sept. 1962.
9.299 Baruch, M., and Singer, J., The Effect of Eccentricity of Stiffeners on the General Instability of
Stiffened Cylindrical Shells under Hydrostatic Pressure, Journal of Mechanical Engineering Sci-
ences (England). 5, (1), March 1963, 23-27.
9.300 Holt, M., A Procedure for Determining the Allowable Out-of-Roundness for Vessels under Ex-
ternal Pressure, Trans. ASME. 74, 1952, 1225-1230.
9.301 Vandepitte, D., Model Investigation of the Collapse of a Steel Water Tower, Preliminary Report
of the Second International Colloquium on Stability of Steel Structures, Liege, April 1977, 599-
607.
9.302 Vandepitte, D., Rathe, J., Verhegghe, B., Paridaens, R., and Verschaeve, C., Experimental Inves-
tigation of Buckling of Hydrostatically Loaded, Conical Shells and Practical Evaluation of the
Buckling Load, in: Buckling of Shells, Proceedings of a State-of-the-Art Colloquium, Stuttgart,
E. Ramm, ed., 1982, Springer-Verlag. Berlin, Heidelberg, New York, 375-399.
9.303 Paridaens, R., Vandepitte, D., Lagae, G., Rathe, J., and Van de Steen, A., Design Equations
Accounting for Elastic Buckling of Liquid-Filled Conical Shells, in: Stability of Plate and Shell
Structures, P. Dubas and D. Vandepitte, eds., Ghent University, 1987, 425-430.
9.304 Vandepitte, D., Van den Steen, A., Van Impe, R., Lagae, E., and Rathe, J., Elastic and Elastic-
Plastic Buckling of Liquid-Filled Conical Shells, in: Buckling of Structures-Theory and Exper-
iment, I. Elishakoff et a!., eds., Elsevier Science Publishers, Amsterdam, 1988, 433-449.
9.305 Esslinger, M., Geier, B., and Wendt, U., Berechnung der Spannungen im elasto-plastischen Ber-
eich, Der Stahlbau, 53, (1984 HI), 17-25.
9.306 von Karman, T.. and Kerr, A.D., Instability of Spherical Shells Subjected to External Pressure,
in: Topics in Applied Mechanics, Memorial Volume to Prof E. Schwerin, D. Abir, F. Ollendorf,
and M. Reiner, eds., Elsevier, Amsterdam, 1965 1-22.
9.307 Kollar, L., Buckling of Complete Spherical Shells and Spherical Caps, in: Buckling of Shells,
Proceedings of a State-of-the-Art Colloquium, Stuttgart, E. Ramm, ed., Springer-Verlag, Berlin,
Heidelberg, New York, 1982, 401-425.
9.308 Zoelly, R., Ober ein Knickproblem an der Kugelschale, thesis, Zurich, 1915.
9.309 Schwerin, E., Zur Stabilitiit der diinnwandigen Hohlkugel unter gleichmiissigem Aussendruck,
Z. angew. Math. Mech. (ZAMM), 2, 1922, 81-91.
9.310 van der Neut, A, De elastische Stabiliteit van den diinnwandigen bol, thesis, Delft, Netherlands,
1932.
9.311 Kloppel, K., and Jungbluth, 0., Beitrag zum Durchschlagproblem diinnwandiger Kugelschalen
(Versuche und Bemessungsformeln), Der Stahlbau, 22, (1953 H6), 121-130.
9.312 Krenzke, M.A., and Kiernan, T.J., Test of Stiffened and Unstiffened Machined Spherical Shells
uder External Hydrostatic Pressure, U.S. Department of the Navy, David Taylor Model Basin,
Washington, D.C., DTMB Report 1741, 1963.
9.313 Evan-Iwanowski, R.M., and Loo, T.C., Deformations and Stability of Spherical Shells Subjected
to the Action of the System of Line Loads and Concentrated Loads, Syracuse University Research
Institute (SURI) Report 834-5, 1962.
9.314 Litle, W.A., Reliability of Shell Buckling Predictions, Research Monograph No. 25, M.l.T. Press,
Cambridge, Mass., 1964.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
804 Shell Buckling Experiments

9.315 Budiansky, B., Buckling of Clamped Shallow Spherical Shells, in Proceedings of the Symposium
on the Theory of Thin Elastic Shells, Delft, W.T. Koiter. ed., North-Holland, Amsterdam, 1959,
64-94.
9.316 Huang, N.C., Unsymmetrical Buckling of Shallow Spherical Shells, AIAA Journal, 1, 1963, 945-
947.
9.317 Huang, N.C., Unsymmetrical Buckling of Thin Shallow Spherical Shells, Journal of Applied
Mechanics, 31, 1964, 447-457.
9.318 Parmeter, R.R., The Buckling of Clamped Shells under Uniform Pressure, GALCIT Report SM
63-53, California Institute of Technology, Pasadena, Calif.; also AFOSR 5362, November 1963.
9.319 Wang, L.R.L., Discrepancy of Experimental Buckling Pressures of Spherical Shells, AIAA Jour-
nal, 5, 1967, 357-359.
9.320 Wang, L.R.L., Boundary Disturbances and Pressure Rate on the Buckling of Spherical Capes,
AIAA Journal, 6, 1968, 2192-2193.
9.321 Sunakawa, M., and Ichida, K., Buckling of Spherical Shells Subjected to External Pressure, 1st
Report, Journal of the Japan Society of Aeronautical and Space Sciences, 21, 1973, 263-270
(in Japanese).
9.322 Yamamoto, Y., and Kokubo, K .. Effects of Geometrical Imperfections and Boundaries on the
Buckling Strength of a Spherical Shell, Computers and Structures, 19, (1-2), 1984, 285-290.
9.323 Krenzke, :\f.A., The Elastic Buckling Strength of Near-Perfect Deep Spherical Shells with Ideal
Boundaries, U.S. Department of the Navy, David Taylor Model Basin, Washington. D.C., DTMB
Report 1713, 1963.
9.324 Dadley, A.E., Tests of Machined High-Strength Steel Spherical Shells Subjected to External
Hydrostatic Pressure, U.S. Department of the Navy, David Taylor Model Basin, Washington,
D.C., DTMB Report 1854, 1964.
9.325 Krenzke, M.A., Hydrostatic Tests of Conical Reducers between Cylinders with and without
Stiffeners at the Cone-Cylinder Juncture, U.S. Department of the Navy, David Taylor Model
Basin, Washington, D.C., DTMB Report 1187, 1959.
9.326 Sabir, A.B., Large Deflection and Buckling Behaviour of a Spherical Shell with Inward Point
Load and Uniform External Pressure, Journal of Mechanical EnRineering Science, 6, (4), 1964,
394-404.
9.327 Lin, M.S., and Popov, E.P., Buckling of Spherical Sandwich Shells, Experimental Mechanics, 9,
(10), 1969, 433-440.
9.328 Pedersen, P.T., and Jensen. J.J., Buckling of Spherical Cargo Tanks for Liquified Natural Gas,
Royal Institution of Naval Architects, Supplementary Papers, 118, 1976, 193-205.
9.329 Odland, J., Theoretical and Experimental Buckling Loads of Imperfect Spherical Shell Segments,
Journal of Ship Research, 25, (3), 1981, 201-218.
9.330 Odland, J., Buckling Experiments with Spherical Shell Segments, Part III; Theory and Experi-
ments, Report No. 77-533, Det norske Veritas, 1977.
9.331 Yao, J.C., Buckling of a Truncated Hemisphere under Axial Tension, AIAA Journal, 1, (10),
1963, 2316-2319.
9.332 Raff, S., Dolensky, B., Krieg, R., and Wolf, E., Untersuchungen zur Beulsicherheit des
NET-Kryostaten-Uberblick i.iber die durchgefi.ihrten Arbeiten und erzielten Ergebnisse.
Kernforschungszentrum Karlsruhe, Primarbcricht 030801P08C, June 1989.
9.333 Dolensky, B., Raff, S., and Krieg, R., Buckling Experiments with a Spherical Steel Containment
Model Under Seismic Loading, in: Proceedings r~{ the Second International Conference on Con-
tainment Design and Operation, October 14-17, 1990, Toronto. Canada, 1, Session 5.
9.334 Wolf, E., Herstellung einer di.innwandigen Kugelschale fi.ir Schwingungsexperimente, Kernfor-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

schungszentrum Karlsruhe (KjK) Nachrichten, 17, (4), 1985, 181-190.


9.335 Wolf, E., Herstellung und Vermessung einer di.innwandigen Kugelschale fi.ir Schwingungsexperi-
mente, Kernforschungszentrum Karlsruhe, Primrbericht 120703P03B, January 1990.
9.336 Raff, S., Dolensky, B., and Krieg, R., Evaluation of Buckling Experiments with a Spherical Steel
Containment Model under Seismic Loading, presented at SMiRT 12, Stuttgart. 1993.
9.337 Ashwell, D.G., On the Large Deflection of a Spherical Shell with an Inward Point Load, in
Proceedings of the Symposium on the Theory of Thin Elastic Shells, Delft, W.T. Koiter, ed.,
North-Holland, Amsterdam, 1959, 43-63.
9.338 Evan-Iwanowski, R.M., Cheng, H.S., and Loo, T.C., Experimental Investigations of Deformations
and Stability of Spherical Shells Subjected to Concentrated Load at the Apex, in: Proceedings
of the Fourth U.S. National Congress of' Applied Mechanics, ASME, 1962, 563-575.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 805

9.339 Thurston, G.A., and Penning, F.A., Effect of Axisymmetric Imperfections on the Buckling of
Spherical Caps under Uniform Pressure, AIAA Journal, 4, (2), Feb. 1966, 319-327.
9.340 Machnig, 0., Ober Stabilitatsprobleme von torusfi:irmigen Schalen, Wiss. Zeitschrift der Hoch-
schule jiir Verkehrswesen, Dresden, 4, 1956, 179-204.
9.341 Machnig, 0., Uber die Stabilitat von torusfi:irmigen Schalen, Technische Mitteilungen, Krupp,
Germany, 21, (4), 1963, 105-112.
9.342 Sobel, L.H., and Fli.igge, W., Stability of Toroidal Shells under Uniform External Pressure, AIAA
Journal, 5, (3), 1967, 425-431.
9.343 Jordan, P.F., Vibration and Buckling of Pressurized Torus Shells, AIAA Paper 66-445, 1966.
9.344 Bushnell, D., Symmetric and Nonsymmetric Buckling of Finitely Deformed Eccentrically Stiff-
ened Shells of Revolution, AIAA Journal, 5, (8), 1967, 1455-1462.
9.345 Fedosov, Yu.A., Stability of a Toroidal Shell Subjected to External Pressure, Izvestiya VUZ,
Aviatsionnaya Tekhnika, 14, (3), 1971, 108-112 (English translation).
9.346 Jordan, P.F., Buckling of Toroidal Shells under Hydrostatic Pressure, AIAA Journal, 11, (10),
1973, 1439-1441.
9.347 Fedosov, Y.A., Refined Solution of the Toroidal Shell Stability Problem, in: Design of Strength,
Mashinostroenie, (17), 1976 (in Russian).
9.348 Fedosov, Y.A., Theoretical and Experimental Study of the Stability of Toroidal Shells with Ex-
tended Pressure, Izvestiya VUZ, Aviatsionnaya Tekhnika, 20, (4), 1977, 98-102 (English trans-
lation).
9.349 Panagiotopoulos, G.D., Stress and Stability Analysis of Toroidal Shells, International Journal of
Pressure Vessels and Piping, 20, 1985, 87-100.
9.350 Wang, A., and Zhang, W., Asymptotic Solution for Buckling of Toroidal Shells, International
Journal of Pressure Vessels and Piping, 45, 1991, 61-72.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

9.351 Almroth, B.O., Sobel, L.H., and Hunter, A.R., An Experimental Investigation of the Buckling
of Toroidal Shells, AIAA Journal, 7, (11), 1969, 2185-2186.
9.352 Fedosov, Y.A., Experimental Study of Toroidal Shell Stability, /zvestiya VUZ. Aviatsionnaya
Tekhnika, 12, (3), 1969, 154-157 (English translation).
9.353 Jordan, P.F., Analytical and Experimental Investigation of Pressurized Toroidal Shells, NASA
Contractor Report CR-261, July 1965.
9.354 Fishlowitz, E.G., Investigation of Elastic Stability of Circular Toroidal Shells under Uniform
External Pressure, Naval Ship Research and Development Center, Department of the Navy, Wash-
ington, D.C., NSRDC Rep. 3338, May 1970.
9.355 Nordell, W.J., and Crawford, J.E., Analysis of Behavior of Unstiffened Toroidal Shells, IASS
Paper 4-4, Pacific Symposium on Hydromechanically Loaded Shells, Oct. 1971, University of
Hawaii, Honolulu, 1973.
9.356 Blachut, J., Galletly, G.D., and Moreton, D.N., Buckling of Near-Perfect Steel Torispherical and
Hemispherical Shells Subjected to External Pressure, AIAA Journal, 28, (II), 1990, 1971-1975.
9.357 Jones, E.O., Jr., The Effects of External Pressure on Thin-Shell Pressure Vessel Heads, ASME
Journal of Engineering for Industry, 84, 1962, 205-218.
9.358 Bart, R., An Experimental Study of the Strength of Standard Flanged and Dished and Ellipsoidal
Heads Convex to Pressure, ASME Journal of Engineering for Industry, 86, 1964, 188-192.
9.359 Slember, R.J., and Washington, C.E., Interpretations of Experimental Data on Pressure Vessel
Heads Convex to Pressure, Welding Research Council Bulletin (119), January 1967.
9.360 Kendrick, S., The Strength of Domes under External Pressure, Naval Construction Research
Establishment, Dumfermline, Scotland, NCRE Report R.531, March 1967.
9.361 Kendrick, S., Externally Pressurized Vessels, Ch. 9 in The Stress Analysis of Pressure Vessels
and Pressure Vessel Components, S.S. Gill, ed., Pergamon Press, Oxford, 1970, 405-511.
9.362 Newland, C.N., Collapse of Domes under External Pressure, in: Vessels Under Buckling Con-
ditions, Institution of Mechanical Engineers, London, 1972, 43-52.
9.363 Blachut, J., and Galletly, G.D., Externally-Pressurized Hemispheres and Shallow Torispheres, in:
Proceedings of ECCS Colloquium on Stability of Plate and Shell Structures, Ghent, Belgium,
April 1987, P. Dubas, and D. Vandepitte, eds., 361-366.
9.364 Galletly, G.D., Kruzelecki, J., Moffat, D.G., and Warrington, B., Buckling of Shallow Tori-
spherical Domes Subjected to External Pressure-A Comparison of Experiment, Theory, and
Design Codes, Journal of Strain Analysis, 22, (3), 1987, 163-175.
9.365 Galletly, G.D., Buckling of Shallow Dished Ends Under External Pressure-A Caveat, in: Pro-
ceedings of Institution of Mechanical Engineers, Part C, 201, (C5), 1987, 373-378.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
806 Shell Buckling Experiments

9.366 Blachut, J., and Galletly, G.D., Externally-Pressurized Torispheres-Plastic Buckling and Col-
lapse, in: Buckling of Structures-Theory and Experiment, I. Elishakoff, J. Arbocz, C.D. Bab-
cock, and A. Libai, eds., Elsevier, Am>terdam, 1988, 29-45.
9.367 James, S., Hyperbaric Test Facility at the University of Liverpool, BSSM Annual Conference,
University of Lancaster, Sept. 1984; also University of Liverpool, Department of Mechanical
Engineering, Internal Report A/088/84, April 1984.
9.368 Kendrick, S.B., Shape Imperfections in Cylinders and Spheres. Their Importance in Design and
Methods of Measurement, Journal of Strain Analysis. 12, 1977, 117-122.
9.369 Blachut, J., and Galletly, G.D., Clamped Torispherical Shells under External Pressure-Some
New Results, Journal of Strain Analysis, 23, (I), 1988. 9-24.
9.370 Moffat, D.G., Blachut, J., James, S., and Galletly, G.D., Collapse of Externally-Pressurized Petal-
Welded Torispherical and Hemispherical End Closures, in: Pressure Vessel Technology-ICPVT-
7, Verband der Technischen Uberwachungs-Vereine e.V. (VdTUV), Essen, 1992, 119-137.
9.371 Blachut, J., and Galletly, G.D., Influence of Local Imperfections on the Collapse Strength of a
Torisphere, Proceedings of the Institution of Mechanical Engineers, 207, 1993, 197-207.
9.372 Galletly, G.D., Blachut, J., and Moreton, D.N., Internally-Pressurized Machined Domed Ends-
A Comparison of Plastic Buckling Predictions of Deformation and Flow Theories, Proceedings
of the Institution of Mechanical Engineers, 204, (C3), 1990, 169-186.
9.373 Moffat, D.G., Blachut, J., James, S., and Galletly, G.D .. Collapse of Externally Pressurized
Torispherical and Hemispherical Domes. University of Liverpool, Department of Mechanical
Engineering, Internal Report A/151/89, 1989.
9.374 Hilbitt, H.D., Karlson, B.I., and Sorem.en, E.P., ABAQUS-User's Manual, Version 4.8.5, HKS
Inc., Providence, R.I., USA, 1990.
9.375 Galletly, G.D., Stress Failure of Large Pressure Vessels-Recommendations Resulting from Stud-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

ies of the Collapse of a 68 ft high x 45 ft dia. Pressure Vessel, Tech. Report No. 45-57, Shell
Development Corp., Emeryville, California, March 1957.
9.376 Meesters, A.G., and Slaaf, C.M., An Investigation on the Stress Distribution, Elastic and Plastic
Deformations in a Large, Thin-Walled Dished Head of a Pressure Vessel by Means of a Hydro-
static Test, Staatsmijnen In Limburg, Report Chern. Constr. Department No. 88, 1949.
9.377 Fino, A., and Schneider, R.W., Wrinkling of a Large, Thin Code Head under Internal Pressure,
Welding Research Council Bulletin, No. 59, June 1961, 11-13.
9.378 Mescall, J., Stability of Thin Torispherical Shells under Uniform Internal Pressure, in: Collected
Papers on Instability of Shell Structures, NASA TN-Dl510, Dec. 1962, 671-692.
9.379 Galletly, G.D., The Buckling of Fabricated Torispherical Shells under Internal Pressure, in: Buck-
ling of Shells-Proceedings of a State-of-the-Art Colloquium, Stuttgart 1982, E. Ramm, ed.,
Springer-Verlag, Berlin, 1982, 429-466.
9.380 Galletly, G.D., The Background to Forthcoming Design Proposals for Two Shell Buckling Prob-
lems, in: Behaviour of Thin- Walled Structures, J. Rhodes and J. Spence, eds., Elsevier Applied
Science Publishers, London, 1984, 179-210.
9.381 Galletly, G.D., Torispherical Shells, in: Shell Structures Stability and Strength, R. Narayanan,
ed., Elsevier Applied Science Publishers, London and New York, 1985, 281-310.
9.382 Stennet, R., Gummed-up Valve Causes Vessel Collapse. Vigilance quarterly journal of National
Vulcan Engineering Insurance Group), 2, (4 ), 1970, 45-46; also Chartered Mechanical Engineer,
17, Oct. 1970, 404-405.
9.383 Kemper, M.J., Buckling of Thin Dished Ends under Internal Pressure, in: Conference on Vessels
Under Buckling Conditions. Institution of Mechanical Engineers, London, Paper Cl89/72, Dec.
1972, 23-32.
9.384 Galletly, G.D., Internal Pressure Buckling of Very Thin Torispherical Shells-A Comparison of
Experiment and Theory, Paper G2/3, Proceedings of 3rd International Conference on Structural
Mechanics in Reactor Technology (SMzRD, London, Sept. 1975.
9.385 Esztergar, E.P., Development of Design Rules for Dished Pressure Vessel Heads, Welding Re-
search Council Bulletin, No. 205, May 1976, 1-34.
9.386 Galletly, G.D., Some Experimental Results on the Elastic-Plastic Buckling of Thin Torispherical
and Ellipsoidal Shells Subjected to Internal Pressure, Preliminary Report 2nd International Col-
loquium on Stability of Steel Structures, Liege, 1977, 619-623.
9.387 Galletly, G.D., Plastic Buckling of Torispherical and Ellipsoidal Shells Subjected to Internal
Pressure, Proceedings, Institution of Mechanical Engineers, 195, (26), 1981, 329-345.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 807

9.388 Stanley, P., and Campbell, T.D., Very Thin Torispherical Pressure Vessel Ends under Internal
Pressure: Test Procedure and Typical Results, Journal of Strain Analysis, 16, (3), 1981, 171-
186.
9.389 Stanley, P., and Campbell, T.D., Very Thin Torispherical Pressure Vessel Ends under Internal
Pressure: Strains, Deformations and Buckling Behaviour, Journal of Strain Analysis, 16, (3),
1981, 187-203.
9.390 Roche, R.L., and Autrusson, B, Experimental Tests on Buckling of Torispherical Heads and
Methods of Plastic Bifurcation Analysis, ASME Journal of Pressure Vessel Technology, 108, May
1986, 138-145.
9.391 Raju, P.P., An Overview of Buckling and Rupture Tests of Torispherical Heads under Internal
Pressure, in: Pressure Vessel Components Design and Analysis, ASME PVP 98-2, 1985, 77-82.
9.392 Miller, C.D., and Grove, R.B., Pressure Testing of Large-Scale Torispherical Heads Subject to
Knuckle Buckling, International Journal of Pressure Vessels and Pipinr;, 22, 1986, 147-159.
9.393 Blackler, M.J., and Ansourian, P., Buckling Behaviour of Two Full-scale Tanks under Internal
and External Pressure, in: Proceedings International Colloquium on Stability of Plate and Shell
Structures, Ghent, April 1987, P. Dubas and D. Vandepitte, eds., 1987, 355-360.
9.394 Robinson, M., Kirk, A., and Gill. S.S., An Experimental Investigation into the Plastic Behavior
of Oblique Flush Nozzle in Spherical Pressure Vessels, International Journal of Mechanical
Sciences, 13, 1971, 41-61.
9.395 Roche, R.L., and Alix, M., Experimental Tests on Buckling of Ellipsoidal Vessel Heads Subjected
to Internal Pressure, in: Proceedings of 4th International Conference on Pressure Vessel Tech-
nology, Institution of Mechanical Engineers, London, 1980, 159-165.
9.396 Thurston, G.A., and Holston, A.A .. Buckling of Cylindrical Shell End Closures by Internal
Pressure, NASA CR-540, July 1966.
9.397 Sanders, J.L., and Liepins, A.A., Toroidal Membrane under Internal Pressure, AIAA Journal, 1,
(9), 1963, 2105-2110.
9.398 Mercier, J., Fremau, J., and Rocha, A., Large Deformations and Stresses of a Thin, Highly
Elastic, Toroidal Shell under Internal Pressure, International Journal of Solids and Structures,
6, 1970, 1233-1241.
9.399 Libai, A., Nonlinear Membrane Theory, in: Theoretical and Applied Mechanics 1992, Proceed-
ings of the 18th ICTAM, Haifa, Israel, August 1992, S.R. Bodner, J. Singer, A. Solan, and Z.
Hashin, eds., Elsevier Science Publishers, Amsterdam. 1993, 257-280.
9.400 Albus, J., and Ory, H., Comparison of Different Liquid Hydrogen Tank Integration Concepts for
the ELAC-1 Research Configuration, Zeitschriji fur Flugwissenschaften und Weltraumforschung,
17, 1993, 149-156.
9.401 Lundquist, E.E., Strength Tests of Thin-Walled Duralumin Cylinders in Combined Transverse
Shear and Bending, NACA TN 523, April 1935.
9.402 Yamaki, N., Naito, K., and Sato, E., Buckling of Circular Cylindrical Shells under Combined
Action of a Transverse Edge Load and Hydrostatic Pressure, in: Thin Walled Structures, J.
Rhodes and A.C. Walker, eds., Granada Publishing, 1980, 286-298.
9.403 Lu, S.Y., Buckling of a Cantilever Cylindrical Shell with a Transverse End Load, AIAA Journal,
3, 1965, 2350-2351.
9.404 Schroder. P., Uber die Stabilitat der querkraftbclasteten diinnwandigen Kreiszylinderschale,
ZAMM, 52, 1972, T145-T148.
9.405 Griguliuk, E.I., and Kabanov, V.V., Ustoiczivost Obolocek (Stability of Shells), Nauka, Moscow,
1978, 199-204 (in Russian).
9.406 Yamaki, N., and Kodama, S., Buckling of Circular Cylindrical Shells under Torsion-Report 2,
Reports lif the Institute of High Speed Mechanics, Tolwku University, 18, 1967, 121-142.
9.407 Galletly, G.D., and Blachut, J, Buckling of a Cantilevered Cylindrical Shell Subjected to a
Transverse Shearing Force at Its Tip, in: Proceedings of the 3rd International Colloquium on
Stability of Metal Structures, Paris, November 1983, Preliminary Report CTICM, 383-389.
9.408 Galletly, G.D., and Blachut, J., Plastic Buckling of Short Vertical Cylindrical Shells Subjected
to Horizontal Edge Shear Loads, Journal of' Pressure Vessel Technology, 107, 1985, 101-106.
9.409 Choi, H.-S., Tanami, T., and Hangai, Y., Failure Tests of Cantilevered Cylindrical Shells under
a Transverse Load, in: Shells, Membranes and Space Frames, Proceedings of IASS Symposium,
Osaka, 1986, K. Heki, ed., 1, Elsevier, Amsterdam, 1986, 265-272.
9.410 Murakami, T., Yoguchi, H., Hirayama, H., Sawada, Y., and Nakamura, H., Buckling of Short
Cylinders with Elliptical Head and Core Support Structure under Transverse Shearing Loads,
Transactions of SMiRT 10, Anaheim. California, 1989, 223-228.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
808 Shell Buckling Experiments

9.411 Hagiwara, Y., Akiyama, H., Kokubo, K., and Sawada, Y., Postbuckling Behavior during Earth-
quakes and Seismic Margin of FBR Main Vessels, International Journal of Pressure Vessels and
Piping, 45, 1991, 259-271.
9.412 von Karman, Land Tsien, H.S., The Buckling of Spherical Shells by External Pressure, Journal
of the Aeronautical Sciences, 7, 1939, 43-50.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
10
lniuallmperlecuons

10.1 Introduction
In the introduction to Chapter 3 it was pointed out explicitly that if one wants to achieve a
good correlation between test data and the corresponding theoretical buckling load predic-
tions, then one must account for the effects of the unavoidable initial imperfections. It was
further stated that depending on the particular application, initial imperfection could have
different meanings. Besides deviations from the perfect shape, the so-called geometric im-
perfections, load eccentricities in columns, residual stresses in welded assemblies (to be
discussed in Section 10.12) or delamination in layered composite structures (to be described
in Chapter 14) are all examples of initial imperfections. Finally, it was also stated that the
degree to which the presence of initial imperfections will affect the ultimate load-carrying
capacity of the structure depends on the particular combination of external load and the type
of structure under consideration.
In some cases the buckling load at a bifurcation point is not necessarily equal to the
maximum load the structure can support. In other cases the predicted bifurcation buckling
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

load of the structure can never be reached in experiments.


For instance, comparing the buckling and postbuckling behavior of imperfect columns
shown in Figure 2.2 with that of imperfect plates depicted in Figure 2.7, one sees that plates
supported along the unloaded edges can develop considerable postbuckling load carrying
capabilities. Looking now at the buckling and postbuckling behavior of axially compressed
isotropic shells displayed in Figure 3.2, one sees that here the role of initial geometric
imperfections is very dramatic. Indeed, in many practical applications only a fraction of the
predicted bifurcation buckling load of the perfect shell will be attained. As can be seen from
Figure 3.1, where for axially compressed isotropic shells the available experimental results
are compared with the theoretical predictions of the classical linearized stability theory [Eq.
(2.111 )], the empirical lower-bound curve, the so-called knock-down factor, can be as low
as 0.2 for the thin shells used in aerospace applications (Rih > 800, say).
In the past few decades it has been finn! y established that the largest portion of this knock-
down factor is due to initial imperfections in shell geometry, with smaller percentages due
to thickness variations, plastic behavior, initial stresses and poor definition of the boundary
conditions. Thus, as stated by Babcock [9.58] in 1972, "the most important problem con-
fronting the experimentalist is the influence of initial imperfections on the buckling loads
obtained experimentally." Though there has been considerable progress in the last two dec-
ades in both analysis and measurement techniques, Babcock's statement is still true today,
especially if "experimentalist" is replaced by "designer."
A vast amount of literature exists on the influence of initial imperfections on the buckling
of shells. Most of the work has been analytical (see for example [2.47], [3.16], [4.48], [1.14]
Copyright Wiley
Provided byBuckling Experiments:
IHS Markit under license withExperimental
WILEY Methods
in Buckling of Thin-Walled Structures:
Licensee=McDermott Inc - Shells, Built-UpUser=G,
India/8215328006, Structures, Composites
Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
and Additional Topics – Volume 2. J. Singer, J. Arbocz and T. Weller Copyright © 2002 John Wiley & Sons, Inc.
810 Initial Imperfections

and (10.1]) but in the la;,t decades much of it ha.' aho hccn experimental (~ee for example
(9.58), [ 1.17). [ 1.24) and (10.2)). However. the correlation among theory. imperfection mea-
surements and buckling loads is still incomplete. und considerable cffons are :.till needed to
prepare the basis for practi ca l design methods that will be able to reap the potential gai ns
in structura l efficiency.
The inlluence of initial imperfections has also been extensively studied for other structural
clements. Since. however. the e ffect is most pronounced for shells. the discus~ion focu~es
on them and not on other type~ of structures.

10.2 Early Incomplete Imperfection Surveys


The ftrst imperfection me:t~ui'Cments on cylindrical ~hell ~ were done in 1906 by Stewan
[9. 19]. who investigated the coll apsi ng pressure> of Bessemer steel lap-welded tubes. Since
he antic ipated that the out-of roundness of a tube would heavily infl uence its behavior. he
designed spec ial test equipment that would accurately and quickly indicate the deviation
from perfect roundness of the tube being tested. The con,truction of the ftnal design is shown
in Figure 10.1. in which ;ome minor details. such :1\ the chord for communicating the motion
of the tube under testing to the recording drum together "ith the nece,sary weights and
carrying pulleys. are omiucd in order to make the drawing show more clearly the main
features of the apparatu~. Figure 10.2 displays a ~chematic view of the te>t set-up di~playing
clearly its principle of action.
Notice that because o r the cord wrapping around the tube T and the pulley P anached to
the record drum D. the tube being measured and the record dnun were made to rotate
synchronously. To prevent slippage. the cord was loaded hy weights at both end~. On the
two records shown on the right the line XX is a reference line. It was drawn by rotating the
record drum by hand while a distance piece wa;, placed between the two indexing points.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Fij!urc 10.1 Autographic mcn\uring device cs(X'cially dc,igncd for obtai ning out-of·ruundnc>' mea-
suremems of tube' (lrnm (9.191)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Early Incomplete Imperfection Surveys 811

i
!
ti
iI
G
0
Figure 10.2 Sketch showing the principle of action of the autographic measuring dc,•ice (from [9.19])

The length of this distance piece was made equal to the nominal outside diameter of the
tube being measured. The resu lt. of course, was a straight horizontal line. The li ne YY was
then produced by rotating the tube between the indexi ng points as described above. The
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

distances between the two lines then showed, to a tenfold magnification, the variation of the
aclual diamelers from the nomina l diameters for any given cross-section. Figure I 0.3 shows.
to a reduced scale, a representalive example selec!ed from the about 6000 records collected
by S1ewar1.
After a very e laborale and meticu lous evaluation of his tesl results. Stewart concluded 1ha1

I. " [T]he e lemenl of greatest weakness in a commercial lap-welded tube is its deparlure
from roundness, even when this deparlure from roundness is comparatively small."
2. " [T]he slight out-of-roundness of the tubes 1ested was 1he chief factor in determi ning
the place of collapse.''
3. " [T]he usual departure from roundness has a more pronounced effect in determi ning the
manner of collapse."

About 30 years laler Jasper and Sullivan [9.36) essentially repeated Slewart's lests using
the improved-qualily commercial steel tubes then available. Besides measuring the out-of-
roundness of the tubes with an apparatus very similar to lhe one used by Stewart, they also
conducled exlensive lhickness measurements with the device shown in Figure 10.4.
The tube to be measured was suppor1ed by the rollers R al bolh ends. The measuring de-
vice M was then rolled along the track T to the poi nt a1 which a thickness measuremenl was
to be taken. where the upper a1m VA was on the outside and lhe lower arm LA was on the
inside of the lube. The measuring knob N was held in con1act with the inner wall by the coun-
terweight W. The wall thickness in thousandths of an inch was then read on 1he gage G.
The arms of the measuring device were adj ustable along the vertical post V. so !hat the
thickness could be measured for tubes of various diameters. Wilh the carriage C the mea-
suring device could be rolled back and forlh along one half to the lube. Reversing !he tube
enabled 1he thickness of the o1her half to be delermined. By revolving the tube on lhe rollers
R. the wall lhi ckness a1 various poi nts along the periphery could be measured.
Defining thickness varialions by the ratio (1,"• - t.,,.,)lOO!t,," !hey found thallhe varialion
of wall lhickness of the tubes tested by them at the A.O. Smith Corporation was 2.5 percenl,
whereas a re-analysis of Stewart 's dala [9.191 indicated a thickness variation of about 12
percent.
Due lo its inlerest in 1he design of submersibles, the U.S. Navy has been sponsoring shel l
research for many decades. As early as 1931 Saunders and Windenburg [9.33] reported on
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
812 lnitiallmperfections

~~'-··
~J -?~:
~~·
-----------(9~1
------=~------ l
~~·

~~

~~··
~7·~··
-------~!
-----------~ l,

~··
~!,

Figure 10.3 Results of out-of-roundness measurement (from [9.19])

the results of a research program completed under the auspices of the ASME Special Re-
search Committee on the Strength of Vessels under External Pressure. The aim of this in-
vestigation, carried out at the U.S. Experimental Model Basin (later renamed the David
Taylor Model Basin), was the development of a design formula for relatively short (L/ R -s
2.0) externally pressurized cylinders based upon the logical application of accepted theories.
Although it was known that for shorter rubes (L/ D < 8) the collapse pressure will increase,
Saunders and Windenburg were the first to trace the development of a collapse pattern
consisting of numerous lobes for externally pressurized cylindrical shells stiffened by closely
spaced internal rings.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Early Incomplete Imperfection SutVeys 813

, ..· T

'~--------------------------~'0
Figure 10 .4 Apparatus for measuring thicknc~< of a pipe (from [9.36])

To record the variation in circu lar fom1 of their test specimens they used the simp le device
shown in Figure 10.5. T his device consisted of a removable central bar or spi ndle upon
which a swinging arm of tubi ng was mounted. T his ann carried at its o uter end a tracing
wheel. with pointer and mu lt iplying mechanism. and a chart. The chart was graduated in
radial lines representing degrees o r divisions of a whole circle and in concentric circles
representing given absolute radial displacements of the tracing wheel. A chain running over
two pulleys, one mounted on the central spindle and one on the chan holder. maintained the
chan in a given direction in space as the arm was rotated.
The simple operation of rolling the tracing wheel around the inside of the cylinder caused
an automatic record, to a magnified scale. to be made of the exact shape of the section. By
repeating the operation after some external pressure has been applied, o ne could detect the
radial deformati ons caused by the external load . In Figure 10.6 the solid line represents the
init ial contour of the middle of the shell; the broken li ne shows the contour just prior to
collapse. Position of lobe format ion of the entire model was influenced by the position of
the lobe at point I, where the in itial local variation in radius was great. Final collapse
occurred at point I and the two adjacent lobes.
:"vvoving the ann along the spindle allowed similar records to be made at other sections.
Systematic imperfection measurements consisting of circumferential scans of cylindrical
and spherical shells under hydrostatic pressure were carried o ut in the fifties at the U.S.
Navy David Taylor Model Basi n. called now Naval Surface Warfare Center (sec for example
[9.5R]. [ 13.53]. (1 3.54] and I 10.3]). /1. schematic diagram of the test setup fo r cylindrical
shells is shown in Figure 10.7.
The probe. mounted on a carriage sliding alo ng the central dcflectometer shaft , was es-
sentially a circularity measuring device. Using a spring. the end of the probe was held in
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 10.5 D~vicc ror tracing comours of sections (from [9.331)

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
814 lnitial/mperfections

Figure 10.6 Diagram representing shape of thin cylindrical shell and mode of collapse by instability
(from [9.33])

contact with the inner wall of the test cylinder. Radial displacements at this measuring device
were transmitted electrically and amplified to actuate a pen in proportional motion. Recording
was then accomplished on a circular disc, which rotated in synchronism with the measuring
arm so that a continuous plot of circularity was obtained in polar coordinates. One such plot
was made at each stiffener and mid-bay position. Amplification with a 50-to-1 ratio was
available so that radial displacements as small as 0.001 in. could be detected.
A typical plot of the radial displacements is shown in Figure I 0.8. Notice that at the zero
pressure condition the circulatory chart shows the presence of a well-defined initial waviness,
or as the authors put it, "it reveals the existence of embryonic lobes." Under increasing
pressure the initial waviness was essentially preserved. That is, those regions initially dis-
placed inward moved inward with increasing pressure, whereas those displaced outward more
or less remained fixed or moved slightly outward. Under increasing pressure, displacements
occurred nonlinearly and without axial symmetry until suddenly a visible lobe was formed.

Probe PosiUan Probe Elev.


Indicator -~-t-1-r- Cable

Probe Sbri.JIIt Linl<


Retradnc -+++-'-1---l
Cable

37 in. Diameter
Probe Pressure Tank

Hold Down
Ring

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 10.7 Schematic diagram of the 37-in. diameter pressure tank showing the model and defiec-
Copyright Wiley
tometer in place (from [13.54])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Early Complete Imperfection Surveys 815
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 10.8 Progressive contours of the shell of model BR-5 midway between stations 9 and 11 (from
[13.54))

Upon reviewing the experimental results obtained, the authors could not decide what
precisely the effect of the observed initial imperfections was. They recommended carrying
out further carefully planned experiments in order to decide whether hydrostatically loaded
cylindrical shells behaved as do compressed bars, or as compressed cylinders, or as some
other perhaps mixed process.
In the early fifties the concept of maximum out-of-roundness was introduced as the cri-
terion for the imperfection of a shell (see for example [9.300]). The out-of-roundness ex-
tracted only the magnitude of the geometric imperfection and disregarded its shape (which
is actually the prime factor that determines the influence of the imperfection on the buckling
load and behavior) and could therefore not provide the desired correlation between prediction
and experiments (see for example [9.201]).
Axial imperfection scans were also made in the fifties and sixties at DTMB [10.4] and at
GALCIT [9.86] as well as measurements of spherical cap generators [9.339]. A number of
other imperfection measurements of the period are listed in [9.58]. Usually in all these
experiments the axial and circumferential scans did not have a common reference, and even
when they did this fact was not utilized in the data reduction.

10.3 Early Complete Imperfection Surveys


Only in the late sixties and early seventies were complete and automated imperfection sur-
veys introduced. The first ones were those developed simultaneously at GALCIT (Graduate
Aeronautical Laboratories at Caltech) and at Stanford University in 1968. The GALCIT
system (see [9.194], [9.221], [10.5] and [1.25]) used punched card recording and a perfect
reference cylinder found by a least-squares fit of the measured imperfection data. This perfect
reference cylinder was a modification of the mean cylinder concept introduced earlier by
Coppa [10.6]. The Stanford system [10.7] used analog recording with AD conversion and
digital data reduction and then presented the data in contour maps, with no further data
reduction.
The GALCIT imperfection measurement and testing rig is shown in Figure 10.9 with an
integrally stringer-stiffened shell in position. The specifications of the experimental program
called for a scanning device that could pick up and record imperfections of only fractions
of
Copyright Wiley
the test specimen's wall thickness of 0.1016Licensee=McDermott
Provided by IHS Markit under license with WILEY
mm (0.004Incin.). In addition, the scanning
- India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
816 lnitiallmperfections

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Figure 10.9 GALCIT laboratory-scale testing machine and imperfection scanning device with stringer-
stiffened shell (from [ 1.25])

device had to travel bo th in the axial and the circumferential di rections in order to record a
complete surface map of the shell being tested.
Such a scanning device was bui It around a noncontacting pickup that measured
the distance between the end of the pickup and the metal surface of the shell. The
pickup was installed in a movable support at the end of a long shaft that protruded inside
the shell being tested. The shaft was moved in the axial and circumferential directions by
small electric motors. The scanning sequence. cons isting of a circum ferential scan followed
by an axial advance, was controlled by strategically placed micro-swi tches. The controlled
end displacement-type testing machine, scanning mechanism and a stringer-stiffened shell
are shown in Figure I 0. 10. W hen carrying out the imperfection measurements the o utput of
the pickup was moni tored on a digital voltmeter, whose readings were recorded d uring the
circumferential scans at preset intervals o n punched cards. The o utput of the pickup was
also displayed o n an X-Y plotter. which afforded a constant surveillance of the test and an
instant warni ng if a breakdown in the equipment occurred.
The earlier studies with this scanning device on unstiffened electro-deposited copper shells
19. 194] used a reluctance-type pickup developed in the GALCIT electroni c shop. The initial
work on the stiffened shells showed that the eddy currents generated by this pickup pene-
trated int.o the surface being measured to a greater depth than previously thought. The stiff-
eners on the shell exterior significantly affected the displacement readings. Therefore, a
capacitance-type pickup was developed for use with the same data acquisition system. How-
ever, the capacitance pickup was less stable than the reluctance pickup. This was overcome
to some degree by controlling the temperature of the testing room and the electronic equip-

Figure 10.10 Cylindrical shell testing configuration (from [9. 194])


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Early Complete Imperfection Surveys 817

ment. In addition, a pickup calibration was performed immediately prior to and after each
buckling test.
Before the harmonic components of the measured imperfection surface could be computed,
it was necessary to define what was considered the perfect shell. This was done by finding
the best-fit cylinder to the measured data of the complete initial imperfection scan. Using
the method of least squares, the program first computed the sum S of the squares of the
normal distances from the measured points in space to the surface of the assumed best-fit
cylinder (see Figure 10.11). Minimizing S with respect to the unknown parameters X 1, Y 1,
s 1 = 7r/2 - a, s 2 = 71"!2 - f3 and R yielded five simultaneous algebraic equations in the
five unknowns. Numerical solution of these equations then determined the best-fit or perfect
cylinder. Notice that by this procedure the rigid body displacements and rotations of the
shell with respect to the scanning system were removed from the measured data. Next the
measured displacements were recomputed with respect to the newly found perfect cylinder.
A typical adjusted scan (initial imperfections referred to the best-fit cylinder) is shown in
Figure 10.12 for the copper electroplated seamless isotropic shell A-12, from the Caltech
test series described in [9.194].
At Caltech the imperfection, referred to the perfect cylinder, was expanded in the late
sixties in a double Fourier series, preparing the basis for the rational data reduction that is
becoming the standard today, as will be discussed later in more detail.
The concept of the GALCIT system was transferred to the Technion in Haifa, where a
basically similar imperfection scanning system was developed in the early seventies. The
Technion system consisted of a test rig, shown in Figure 10.13, in which the shells were
fixed and scanned with a pickup moving in a helical path and a recording system. The output
from the scanning pickup (the radial distance and the vertical and circumferential position)
was recorded digitally on a NOVA 2 minicomputer, with a parallel control display on an
X-Y recorder. This test rig was slightly larger than the earlier one at Caltech. It accommo-
dated shells of 240 mm (9.5 in.) diameter, was vertical instead of horizontal and scanned in
a smooth helical path instead of the intermittent circumferential-axial motion of the Caltech
rig. It too was designed for measurement of initial imperfections and their growth during
loading. The digital records of the scans from the NOVA minicomputer were in the form of
paper tapes, which were then read into the IBM 370/168 computer of the Technion com-
putation center for processing.
The measurement could be carried out on an unloaded shell, yielding the initial geomet-
rical imperfections, or on the shell under a specific axial load measured by a special load

X
x'

X,Y,Z Reference axis of traversing pick- up


--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`

X',Y:z• Reference axis of best fit cylinder


d; Normal distance from measured point
to best fit cylinder

Figure 10.11 Best-fit cylinder reference axes (from [9.194])


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
818 lnitiallmperfections

CUIICU .. tt lltt: HTIAL. AN G\.( (RA DIA NS )

Figure 10.12 lnilial imperfeclion measuremenls of shell A -1 2 (from 19. 194))

(a)

(b)

Figure 10.13 Technion labonuory-scale imperfeclion scanning rig (from [ 10.2]): (a) 1cs1 rig: (b) scan-
ning probe for sliflened shells
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Early Complete Imperfection Surveys 819

cell, yielding the growth of imperfection under load. In measurements on unloaded shells,
the shell to be measured was placed in the scanning rig with boundary conditions as close
as possible to those in the ensuing test. The exact procedure depended on the specific bound-
ary condition (see also [10.2]). For example, the shell could be put in the rig already properly
attached to one end of the end rings. The other end ring was then placed on the shell in the
rig and scanning was performed (see Figure 10.13 ), a simple contact probe, a linear variable
differential transformer (LVDT), being used for stiffened shells. The continuous measurement
of displacement along the helical path was digitized by sampling at regular intervals.
The digital records obtained along the helical path were adjusted with the aid of a special
program (INTER-PROG) to the circumferential-axial form (orthogonal grid) required for the
data reduction programs.
The records of the measured radial displacements at the chosen points (along the circum-
ference and at different heights) were the deviations of the inner surface of the shell from
the cylindrical surface generated by the probe during its scanning path. However, as in the
GALCIT system, the actual initial geometric imperfections were taken to be the deviations
from an imaginary cylindrical reference surface, defined as the best-fit cylinder to the mea-
sured data, computed by a least-square method. The recorded measured displacements were
therefore recalculated with respect to this best-fit reference surface, using the data reduction
program developed at Caltech. In this manner any rigid body displacements of the shell with
respect to the scanning system were removed.
A typical adjusted scan (the initial imperfections referred to the best-fit cylinder) is shown

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
in Figure 10.14b for an integrally stringer-stifiened cylindrical shell KR-1, from a Technion
test series studying the buckling of stiffened shells [10.2].

Circumferential angle (radians)


(a)

Circumterenti<~l angle (radiansl

(b)

Figure 10.14 Measured initial imperfection shapes of integrally stringer-stiffened cylindrical shells: (a)
stringer-stiffened shell AS-2 (from [1.75]): (b) stringer-stiffened shell KR-1 (from [10.2])

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
820 lnitiallmperfections

10.4 The Awakening of Imperfection Measurement


Awareness
The seventies, and in particular the late seventies, witnessed an important change in buckling
experiments on shells: the extent of geometric imperfection measurements greatly increased.
In the previous decade imperfection measurements were usually considered useless and both-
ersome, except by a few investigators who believed in their importance, like those at GAL-
CIT, Stanford and Technion mentioned above, or those at DTMB [13.54] and [10.3], or at
UTIAS in Toronto (for example [10.8] and 10.9]) and a few others mentioned in Babcock's
review [9.58]. At the end of the seventies, however, some type of geometric imperfection

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
measurement became an integral part of a properly carried out shell buckling test, be it on
a laboratory scale or on a large scale. A typical statement that illustrates this radical change
is the laconic sentence appearing in the short summary of two Det norske Veritas test reports
([10.10] and [13.71]). "The imperfections are measured and stored on tape."
It became widely accepted that significant advances towards more accurate predictions of
the buckling load of thin shells depended on the availability of extensive data of realistic
initial imperfections and their correlation with manufacturing techniques. For example, at
the IUTAM Symposium on Buckling of Structures held at Harvard University in 1974, this
was one of the main conclusions of the closing round-table discussion.
With the realization of the importance of the shape and amplitude of initial imperfections,
various measurement systems were developed in different laboratories, many of them listed
in Singer's reviews [1.17] and [1.24]), and some recent ones were presented at the Euromech
Colloquium 317 Buckling Strength oflmperfection Sensitive Shells, Liverpool, March 1994
[10.11].

10.5 Complete Imperfection Surveys on Large or Full-Scale


Cylindrical Shells
In order to apply the results of the imperfection measurements on laboratory scale and their
correlation with buckling predictions and tests to engineering practice, information on the
geometrical initial imperfection of full-scale shell structures is required. A program of com-
plete imperfection surveys of full-scale shells was therefore initiated by Arbocz and Babcock
in 1975, and a suitable test system was designed and built.
The first shell measured was a 3.05-m (10-ft) diameter aluminum alloy integrally stiffened
shell at NASA Langley Center [10.12]. The shell was later buckled by bending. The shell
and the initial imperfection survey instrumentation are shown in Figure I 0.15. As in the
laboratory scale systems described in Section 10.3 ([9.194], [1.25] and [10.2]), the technique
employed measured the deviation of the cylinder outer surface relative to an imaginary
cylinder reference surface. Here this was accomplished with a 3.05-m (10-ft)-long aluminum
guide rail supported on the outside diameter of the two steel end rings and a direct current
differential transformer (DCDT) instrument on a trolley car. The trolley car, which was spring
loaded to roll with continuous contact along the guide rail, was slowly driven along the
guide rail by an electric motor. The position of the car was electronically measured using a
potentiometer that rotated as the car moved along the track. Accuracy of the displacement
measurements was to within ± 0.05 mm ( ± 0.002 in.).
Surveys were made along the cylinder length for a distance of 2.06 m (81 in.). The guide
rail was moved stepwise around the cylinder circumference in sa increments, thus yielding
a total of 72 discrete scans for the complete cylinder. Some scans were performed more than
once to determine the repeatability of the measurements, which was found to be within the
tolerance limits of the measuring equipm~nt.
The data was recorded using two methods: ( 1) continuously with an X- Y recorder in which
the deviation from a neutral position (to the reference surface) was plotted as a function of
position along the cylinder length and (2) discretely digitized and recorded every 6 mm
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Complete Imperfection Surveys on Large or Full-Scale Cylindrical Shells 821

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Figure 10.15 Initial imperfectiOn survey instrumentation on NASA shell LA-I (from )10.12])

(0.236 in.) along the cylinder length. thus yielding 343 data points for a typical ~can. Di screte
data acquisition was triggered by a phototube circuit that was opened and closed by a notched
d i~k . which rotated as the trolley car traversed along the gu ide rail. The digitized data was
then w.ed for data reduction. in a manner simi lar to that employed o n the laboratory scale.
The adjusted initial imperfection scan of this large NASA aluminium alloy cy lindrical
shel l is presented in Figure 10.16.
In the late seventies the Cahcch group carried out a ~cries of imperfection surveys at
different aerospace companies in the United States ((10.13] and [10. 14)). They adjusted to
the realities of working on a manufacturing floor and on a tight schedule and u~cd some
rotation hardware present at the manufacturing site.
They u~ed the two-piece portable rail system shm\ n in Figure 10.17a to carry out the
imperfection scans of different Oight vehicles. In the same figure o ne of the te~t shells is
shown mounted in a rotation device. The rail consisted of Mandard 1-beams that were con-
nected with tapered r ins at the center. It was supported hy a center stand that was movable
in the vertical (~-)direct ion and by two end stands that had x-y-: positioning capabi lity. The
displacement transducer was mounted on a cart that ~lid along the rail. The can was spring
loaded so that its precision bearing rollers were pre\~cd against the reference axi~ (the front

•J

••

¥ .
i
,..,,.

t.

Fi~:ure 10.16 Measured initial ,hapc of the 3.05-m ( 10-fl) diameter integrally st iffened alumi num alloy
NASA shell LA· I (from [10. 12))
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
822 Initial Imperfections

{a)

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
{b)

Figure 10.17 Measuremenl or impcrlc ctions on a full -scale aerospace shell: {a) two-piece ponable
rail system: {b) instrumenl cart and position tape on reference beam {from [10.13]
and [10. 14))

edge of the top tlange). The rai l could be aligned optical ly using a theodolite and an optical
target moun ted on the cart. Deviations from straightness were determined using this instru-
ment. The initial error cou ld be determined to about ±0.0 1 mm over the 7.6-m length of
the beam. The exact shape of the beam was recorded and later removed from the measured
data during the data reduction process. The shape of the shell generator was thus measured
with respect to this rail by means of an LVDT of sufficient range to determine not only the
expected initial imperfections. hut also any misalignment of the reference rail wi th respect
to the best-fit perfect shell axis system. The posi tion of the carl with the LVDT along the
rail was determined by a black-and-white tape attached to the rail. The tape was read by a
light-emitting photo-diode pair, which generated a square wave to trigger the data acquisition
system. The cart and the rai l are shown in Figure 10.17b. The data (LVDT and positi on
signal) were carried to the data acq uisi tion system by a nex ible cable that followed the cart
on a slide wire. The cart was moved by a steel cable driven by a variable speed motor and
limited by switches moun ted on the rail.
The portable data acquisition system was specially designed for this type of field testing.
It consisted of an analog-to-digital converter, an HP 9825A programmable calculator LO
control the data acquisition sequence and perform on-site data reduction and a digital X- Y
plotter to di splay the raw data as well as the reduced data.
A three-dimensional plot of the measured initial imperfections referred to the best-fit
cylinder from a large (radius 12 12.1 mm. length 6454.1 mm, wall thickness 1.549 mm)
integrally stiffened shell made out of three pieces is shown in Figure 10.18. The li nes are
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Complete Imperfection Surveys on Large or Full-Scale Cylindrical Shells 823

Figure 10.18 Measured initial shape of aerospace shell X-I (from [10. 14))

drawn in the circumferential direction. but the data were collected from axial scans. The
three welded seams are clearly visible in the imperfection plot. The disturbance halfway
along the length resulted from a ring frame at this location, which tended to minimi ze the
weld imperfection. The waviness was caused by the pockets of the isogrid stiffeners of the
shell wall. A better view of th is is shown in Figure 10.19, where the regular pattern of the
pockets is clearly seen. The large radial displacements at the shell edges were due to the
end domes of the shell bei ng force-fitted inside the cylindrical section.
In all three-dimensional plots positive imperfections are pointed outward. An exception to
this ntle is Figures 10.18 and 10. 19, where positive imperfections are pointed inward. This
change in orientation was used because of the form of the measured initial imperfection near
the two edges. A closer look at the axial cross-plot shown in Figure I 0. 19 reveals that the
initial imperfections consist of three main components, namely a large step function-like
component due to the uniform rad ial displac-emenl produced by the force-fi tted end domes,
a large half-wave sine component in the axial direction plus a small short wave imperfection
component. wh ich accQunts for the waviness caused by the pockets of the isogrid sliffencrs.
Since for stability calculations the uniform radial expansion is of no consequence. the edge
zones have heen eliminated from the data. Thus. the Fourier decomposi tion of the measured
data was done with the ini tial imperfections shown in Figure 10.20. Here positive imperfec-
tions are pointed outward as usuaL
The measured initial imperfections referenced to the best-fit cylinder from another large
(radius 1527.4 mm, length 6047.7 mm. mean wall thickness 2.6629 mm) stringer-stiffened

••
•\
tl) f ,.,. \.\..

.. : •"v,"'·
~~~-----------------

...
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 10.19 Typical meridional shape of aerospace shell X-I (from [10. 14])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
824 Initial Imperfections

Figure 10.20 Measured initial shape of aerospace shell X- 1 with edge zones removed (from [10. 14))

shell made ou t of four pieces arc shown in Figure I 0.21. The four longitudi nal welded seams
are once again clearly visible. Also, since this shell had onl y longitudina l stiffeners. the short
wave waviness caused by the imersccting transverse stiffeners observed on the previous shell
is absent. On the other hand, as can be seen in Figure 10.2 1, this shell had more waviness
in the circumferential direction.
With similar scans on aerospace shells in Europe in the early eighties, the Delft University
of Technology group used a one-piece portable rai l system. shown in Figure 10.22, for
performing imperfection scans on an Ariane interstage II/III shell [ I 0.15]. In thi s setup the
shell was positioned upright on a two-piece turntable. with the reference beam placed parallel
to it on an adjustable tripod.
Three LVDT pick ups were installed on the reference beam. The fixed ones on the top and
on the bottom were beari ng agai nst the machi ned end rings. The third one was installed on
a carriage, wh ich was moved along the beam by an electric drive to record the shape of the
corresponding shell generator. Next the shell was rotated to a new position, followed by
another axial scan. The process was continued until the whole surface was surveyed and
recorded. The exact shape of the reference beam was measured optically and was removed
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY
Figure 10.21 Measured initial shape Licensee=McDermott
of aerospace shell X-2 (from [10. 14'))
Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Imperfection Surveys on Large Shells of Revolution 825

Figure 10.22 TV-Delft initial imperfection survey system employed on Ariane intcrstagc II /I II stringer
stiffe ned shell ( from f 10. 1S])

from the measured data during the data reduclion step. The axial posi1ion of the carriage
was recorded by an electro-oplieal device that scanned a strip with equally spaced cuwuts.
The resulting square-shaped signal was used 10 digilize the data in intervals of 10 mm .
Spec ial care was laken 10 delect and record possible random rigid body displacements of
the shell with respect to the fixed position of the reference beam while rotal ing the turntable
10 a new circumfererllial posilion. This was accomplished by monitori ng the planar displace-
ments of a cal ibrated circular ring placed in 1he center of the turntable. T hese displacements
were I hen removed from the measured data during data rcduc1ion (see f I 0.1 5] fo r a detailed
description of this step). The measuring and data acquisi1i on sys1em was in general similar
10 lhat employed by the Callech group for their aerospace shell measurements (fo r more
details see [ 10. 14] and [1 0.1 5]).
The Ariane interstage 11 / lll shell shown in Figure 10.22 (radius 1300.0 mm, length 2730.0
mm, wall thickness 1.2 mm) was huih up out of eight identical panels. The joi nts between
adjacent panels were offset lap splices and o n the o utside 120 equally spaced hat-shaped
stringers were riveted to the shell surface. The whole structure was held circular by two
precision-machined end rings on lhe outside and five equally spaced bracket-shaped rings
o n the inside. A three-dimensional view of the measured initi al imperfections referred to the
so-called best-fit cylinder is shown in Figure I 0.23. In alllhe tluee-dimensiona l views shown.
the measured initial radial imperfections have been normal ized by the coJTesponding wall
thicknesses.
As pointed Out in ( 1.1 7 ). f 1.24), f 10. 13) and f 10.14). addi tional large cylindrical shells
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

were scanned in the late seventies and early eighties by the Caltech and TU-Delft gmups,
as well as at Del norske Veritas [I 0.1 0] and [ 13.71]. Imperial College [ 13.69] Glasgow
Univcrsily f 13.7], Georgia Tech [ 13.87] and more recentl y al the Universily of Toron1o
[ I0. I 6], the Uni versity of Sydney [ I0.1 7] and the Pol itecnico di Milano [ I0.1 8].

10.6 Imperfection Surveys on Large Shells of Revolution


Imperfection surveys were also carried out during the same period on o ther shells of revo-
Copyright Wiley
lution. One example were the large spherical shell
Provided by IHS Markit under license with WILEY
segments, representing spherical cargo
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Initial Imperfections

Circumferential angle (radians)


Fil,'UI'C 10.23 Measured initial imperfection shape of Ariane imerstage ll / fll shell AR 23-1 (from
[10. 14])

tanks for LNG transport. tested at Oct norske Veritas (see [9.1 82). [9.183) and [9.329)).
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 10.23 shows a diagram of the system. T he machined end rings of the spherical
segment (6) and (7) in Figure I 0.24 served as rai ls for a carriage (8). whi ch was equipped
with an angle counter (9) and displacement transducers ( 10). that could be moved vertically.
In the scan the transducers were fixed in certain vertical positions and the carriage was
moved manually circumferentially. The circumferential position. recorded by the angle
counter, was stored on tape together with the data from the displacement transducers. The
radial deviation (the imperfection) was measured within ± 0. 1 mm accuracy ( ± 5% of the
thickness 1) and related to !he mean radius. assumed 10 be the correspondi ng radius of an
ideal sphere. B01h initial imperfections and growth of imperfections were measured.
Results of the inilia l imperfcclion mcasuremcnls from model No. 10 LC J are shown in
Figure 10.25. Notice thai the locations of the three welded joi nts in the meridional direction
can clearly be identi fied. For the same model, the development of the rad ial denections at
lhc equator under increasi ng load is displayed in Figure I 0.26.
Another example is the experiments on large fabricaled torispherical domes under pressure
carried out in !he eighties at the Chicago Bridge and Iron Co. by Miller and his co-workers
([9.391) and [9.392)). Two 4.877-m ( 192-in.) diameter (D) welded steel domes. wi th knuckle
radius (r) 0.829 m (32.64 in.), spherical radius (R,) 4.389 m (172.8 in.) and wil h (D//r) =
980 and 711, respecti vely, were fabricated using mclhods. workmanship, inspection tech-

Copyright Wiley
Figure 10.24 Diagram of initial geometry mapping system employed by Det
Provided by IHS Markit under license with WILEY norske Veritas on large spherical
Licensee=McDermott shell segments
Inc - India/8215328006, (from [9. 182])
User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
828 Initial Imperfections

Figure 10.27 Chicago Bridge and Iron Co. test set -up for torispherical head models-detail of trans-
ducers fo r measuremenl of initial shape and its deformation under pressure (couneS)' of
Dr. C. D. Miller)

ducers measured the shape of the shell relative to the machined surface of the arm. Any
change in the radi al dimensions cou ld therefore be presented as a contour of the actual shape
or as a change in shape from some previous measurement. There were 15 LVDT's mounted
on the arm, as shown in the figure. to measure the initial shape and monitor its change during
the test. Shape measurements were taken at 3° increments around the entire model at pre-
determined pressure values. each scan yielding 1680 read ings.

10.7 Recent Laboratory Scale Imperfection Measurement


Systems
As pointed out in Section I 0.3. the displacement probe employed in the scanni ng system is
an element of cardi nal imponance. Linear variable differenti al transformers (LVDT's) or
other special "soft " contacting probes have been used in most scans on small stiffened shells
and on the systems for large shells discussed above. However. for very thin shells. panicu-
larly isotropic ones, noncontacting probes arc preferable to prevent distortion of the imper-
fection measurement by the small probe force. T his led to the noncontact reluctance type
pickup used in the early Caltech system. For stiffened shells a capacitance pickup was then
developed, as di scussed in Section I 0.3, which, however, was less stable and required con-
trolling the temperature of the testing room and equipment. Temperature variations also
caused difficulties to a noncontact capacitance probe developed at DFVLR Braunschweig
f(lr their Mylar cyli nders [1 0.19]. as well as 10 a capacitance probe tried at the Technion.
Hence at the Technion LVDT probes were used on stiffened shells (sec Figure. 10.13b).
and a special noncontact mapping servo-system was developed jointly with instruments and
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Control Ltd., Haifa [ 10.20). The system. based on a closed loop transducer, tracks metall ic
surfaces by means of a miniature inductive proximity sensor attached to the moving part of
a linear motor. The displacement of the sensor is measured by a displacement transducer,
leading to very accurate mapping (up to 0.1 % of full scale). T he transducer is placed inside
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Recent Laboratory Scale Imperfection Measuremen t Systems 829

a housing. An ex ternally controlled servo-motor moves the sensor radially to or away from
the shell inner surface. To prevent damage during the sudden buckling of the shell. the
transducer is held inside its housing by a spring that pecmi ts a recoi l of the whole transducer.
The transducer can move also axial ly along a frame composed of three circular steel columns.
The axial mo tion is controlled by a second servo-mowr. and the axial location of the probe
is measured by another LVDT. The whole frame can rotate 180° about the axi s of the
cylinder. permitting any desired ci rcumferential posi tioning of the transducer. This rotation
is controlled by a torque mo tor positioned inside the lower end plate.
The circumferential and axial locatio ns of the probe are continuously shown by two digital
indicators and also stored in digi tal form for further analysis. To control the whole scanning
process. a map of the shell imperfections is plolled o n-line.
This closed loop noncontact probe was used in a unified mu ltipurpose scanning and mea-
surement system for imperfections and vibrations at the Techni on. As poi nted o ut in (1.24],
[ 10.2] and [10.21] , the deficiency of some of the imperfection measurement systems in use
was that imperfectio ns were measured before the shell was fixed in its fi nal bo undary con-
ditions. This was parti cularly emphasized in relation to vibration correlatio n studies [ 13.4 I].
The new system [I 0.20] was designed to overcome this d rawback.
The noncontact probe was used to measure the vibrations and imperfections of cyl indrical
shells. Measurements included natural frequencies. modes of vibration and mapping of im-
perfections. All measurements and mapping were carried o ut by the same probe inside the
closed shell. which was fixed in its final boundary condi tions, and an electronic control
permitted autOmatic execution of the different modes of o peration. Figure 10.28 shows the
complete system bei ng prepared for operation on the spot-welded shel l AACX-11. of a joint
RWTH Aachen-Technion project ( 10.22 ]. and the complete system, in operation o n a spot-
welded shel l AAC- 1 of the same project, is shown in Figure 15.23 of Chapter IS. This
system cou ld accommodate spec imens of diameter 240- 500 mm and length up to 470 mm.
In the imperfection mode the digita l data was recorded at that time o n a NOVA 2 minicom-
puter connected to the experimental system, and analog plots of the imperfection shapes of
the shell were also obtained for real-time control. The digital resu lts were stored o n a com-

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

f igure 10.28 Technion tesJ setup for vibration cocTeial ion 1es1 on sp<H· wcldcd shell AACX-11 with
mu ltipurpose scanning and measuremenl sys1em for vibr.uions and impcrfcclions (from
Copyright Wiley
r1o.22D
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
830 Initial Imperfections

puter disk, which enabled funher numerical analysi~. The scan could be taken at different
load levels. ln the vibration mode. the test procedure wa!> essentially similar to that employed
on the smaller integrally sti iTened shell s (see for example (13.41 ]), though operation of the
new system di ffers considembly. The vibration plo ts obtained were very clear.
Another in situ test setup fo r initi al imperfection measureme nt has also been developed
at the Technion to fi t lhe test rig used for combined loading experiments on sti ffened cyli n-
drical shells. Its purpose was to avoid the problems of change in boundary condi t ion~ arising
from the measurement of initial imperfections in a separate measuring rig and then trans-
ferring the shell to the final test setup. The system, ~hown in Figure 10.29. pem1its mea-
surement of the initial imperfections and their growth while the shell is ready for vibration
and buckl ing tests (the (VCf) vibration correlation technique approach-sec for example
[9.2741). It consists of an LVDT probe that is rotated manually circum ferentiully inside the
shell at preselected heig hts (usually 8 to 10 positions, dependi ng on the leng th of the shell).
The outpu t of the LVDT probe is sto red in a PC. Duri ng measurement of the impe rfections,
an on-l ine plot is generated. Once the shell has been completely scanned. the data is stored
on a tape or disc for f unher analysis.
Another example of a recent laboratory imperfection scanning setup is the STONTVOKS
system (9.167( developed at Delft University of Technology. This test setup was conceived
to ~uppon a research program dealing with the statiMical analysis of the influence of the
measured initial imperfections on lhe buckling behavior of nominally identical shells. The
test apparatus (see Fig ure 10.30) was used to test a large number of nominally ident ical

.._~-=...t-L g ~h
IRIII~::ij~l - 0~1L- -
~, ,l---
~::

[]L. ..·-· ·'il~:l~t~.::-:


'

~GJ
..,.., _ •r• --·-, ,

~Gl C:JlG··-·--1
(\.. ... ~
8
!-----}: :+:: ::·::!....... !-:~;_:_::_____•__i__.

1. &.GAD au. U, .UI AI. IOTINTIOHetU


2. tlliloct CJl atlT 14 , Cl llCI.M,PIJfTIM. P01'VfT1CKlU
), SCQ.If JIICU U . 05C I LL.OS(OPI
4, N JOCI CJJIOJIT •• · 1u.na
S. SlUUI r.A.QU 11. H4fLIFIU
•. $1\AU:l 11. m - wur~ta
1. AHPUrna lt. HICI OPitOHI
I,
t,
PUSSUII VUSEL
Q)IITaot. YALVE
10.
11.
~"""'
~IL L
10, OSCILLATOR 21. [NO PLAn
11 . 'UQUeHCY coun t• U . LYOT
12. 1 • n liCOI.OEil 24 . P'Oifl!l .SU, LY

Figure 10.29 Technion test sclllp for combined loading of >tiffcncd cylindrical shell> wilh ;, si111
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
imperfection scanning system and vibration mc:l'urcment system (from [9.2741)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Evaluation of Imperfection Data 831

Figure 10.30 The TU-Delfl STONIVOKS testing machine (from [9. 167))

specimens (say, beer cans) under easi ly reproducible end conditi ons. The test equipment was
bu ilt around a stable rotating platform and has been fabricated with a high precision. It uses
an LVDT contact probe of very small contact force (about 5 g) and the measuring procedure
and data acquisition are fully automated.

10.8 Evaluation of Imperfection Data


The procedure of evaluating the measured initi al imperfection data has changed considerably
over the past 50 years. Initially one wou ld be concerned only with determin ing the maximum
value of the rad ial variation from the mean circular form. This so-called maximum out-of-
roundness was then used as a measure for the initial imperfection of the shell. Attempts to
establish a correlation between. for instance, the load-carryi ng capabilities of the hydrostat-
ically loaded cylinders and the effect of the measured initial imperfections based on this out-
of-roundness criterion were unsatisfactory (sec r13.53], r13.54] andl10.3)).
Nowadays it is accepted that the shape of the initial imperfections is the prime factor that
determi nes the buckling load and buckling behavior of thin-walled shells; sti ll , practically
all current shell design manuals (see [I 0.23)-[ I 0.27]) recommend the replacement of the
actual imperfections by some empirical equivalent imperfection measure. Although this ap-
proach seems to work satisfactorily (after all. struct.ural collapses seldom occur), at least for
weight-sensitive appl ications. another approach is called for.
To reap the benefits of Koiter's imperfection sensitivi ty theory [3. 15) and [3.16), one must
devise a rational design procedure whereby the engineer can take full advantage of the
theoretical knowledge, experimental results and computational tools at his disposal. It is
apparent that he must also have information about the expected initial imperfection distri-
butions that wil l be present once his struct ure has been completed.
One widely accepted way for presenting the measu red initial imperfection data for further
use is to calculate the coefficients of the appropriate Fourier series representations of the
adjusted initial imperfection data referred to the best-fit cylinder. One can use either the half-
--`,`,`````,`,````,,``,``

wave sine
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
832 lnitiallmperfections

w(x, y) = h ~1 ~sin br~ ( Ck, cos f~ + Dk, sin£~) (10.1)

or the half-wave cosine

w(x, y) = h LL cos /err~ (Ak, cos £_RY + Bk, sin C_RY) (10.2)
k~r~o L
Fourier representations. Chryssanthopoulos [10.28] chose to work with a modiJied form of
Eq. (10.1):

(10.3)

where

(10.4)

whereas Arbocz and Babcock [10.29] used still another expression:

- - 'Y
w(x, y) - h ~ A,o cos l1T
. X
L h '\"""~'Y
+ t~,.c;.:
.
Sill
k
1T
X
L (c kr cos
o
t
y + Dk, Sill. ot Ry)
R (10.5)

for their correlation studies.


In the following, the imperfection measurements on the 10-ft diameter shell structure of
[ 10.12] will be used to illustrate the diJferent ways the results of complete imperfection
surveys can be displayed. At first one can plot the radial deviations from the perfect shell
(from the best-fit cylinder) as the three-dimensional plot shown in Figure 10.16. As can be
seen from this picture, the longitudinal welded seams, have produced a very characteristic
initial imperfection consisting of a half-wave sine in the axial direction and nine full waves
in the circumferential direction. The imperfection amplitudes are greatest at the location of
the welded seams, and the maximum peak-to-peak imperfection is equal to about one wall-
thickness (0.10 in.).
Figure 10.31 shows the variation of the calculated half-wave cosine Fourier coefficients
as a function of the circumferential wave number efor the selected axial half-wave numbers
indicated on the figure. Similar plots for the half-wave sine Fourier coefficients are shown
in Figure 10.32. By computing the respective mean imperfection amplitudes g(=g~c,) the
phase shift in the circumferential direction is eliminated. Once again the characteristic im-
perfection produced by the three welded seams in the axial direction used in assembling the
shell is clearly evident since the plots are dominated by those Fourier coefficients that are
multiples of 3. A comparison of the plots shown in Figures 10.31 and 10.32 further indicates

CIRCUMFERENTIAL WAVE NUMBER R.


--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

0 5 10 15 20 25 30

1
~05
"'
04
0 k: 0
0 k"' 1
"'
<I
~ 03

"'
6 02
>=
u
w
lL 01
a:
w
~
0 00225 0225 ...!Rti\1
0674
BETA"J2CTRJ

Figure 10.31 Circumferential variation of the half-wave cosine Fourier representation-shell LA-I
(from [10.12])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Evaluation of Imperfection Data 833

CIRCUMFERENTIAL NUMBER i
10 15 20 25 30

0 00225 0225 0 674

BETA'd*"i*l

Figure 10.32 Circumferential variation of the half-wave sine Fourier representation-shell LA-1 (from
[10.12])

that for this shell the half-wave sine axial representation is superior to the half-wave cosine
representation since it can reproduce the main features of the measured initial imperfection
survey with fewer Fourier components.
This last statement is further reinforced by the results shown in the following two figures.
Figure 10.33 displays the variation of the calculated half-wave cosine Fourier components
as a function of the axial half-wave number k for e = 0 (axisymmetric components) and for
f = 9 (nine full waves in the circumferential direction). Similar plots for the half-wave sine
Fourier representation are shown in Figure 10.34. Comparing the two figures, as indicated
by the rate of decay of the amplitudes with increasing axial half-wave numbers k, for the
axisymmetric components (£ = 0) the half-wave cosine, and for the asymmetric components
(f = 9) the half-wave sine representation is more appropriate.
It has been shown in [10.30] and [10.12] that by using the complex Fourier representation
the axial dependence of the measured initial imperfections, for a given value of the circum-
ferential wave number ~can be represented as a discrete power spectral density S(w), where,
using the notation of [10.12],

(10.6)

and

AXIAL HALF· WAVE NUMBER k


s 10 1!1 20

0 t •0

~020 0 '•9

.....,• O.IS
ill
......~ 0.10

...
II:
Q,
l! o.os

0 0.0!516 0.!516 1.032

ALFA • k \fl[(f)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Figure 10.33 Axial variation of the half-wave cosine Fourier representation-shell LA-1 (from [1 0.12])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
834 lnitiallmperfections

AXIAL HALF- WAVE NUMBER k


0 5 10 1!5 20
~r-r-.,-r-r-r--r-f--r--.-..,.-.--T-,.--r-""'r-T-:;,
0.6 r-x-

o l•O
0.5
0 t. 9

~··
... • 0.3

g
.... 0.2

~II! 0.1
!
~A ~n-9- c -D-o A-a
0 0.0516 0.~16 1.032

ALFA. k Wtl
Figure 10.34 Axial variation of the half-wave sine Fourier representation-shell LA-I (from [10.12])

(10.7)

Figure 10.35 shows on a logarithmic scale the nondimensional discrete axisymmetric (f =


0) power spectral density (PSD) functions plotted as a function of the nondimensional spatial
frequency w. Notice that only a few low-frequency components of the half-wave cosine
Fourier representation are within one decade of each other, whereas the half-wave sine Fou-
rier representation has many higher-frequency components with significant amplitudes. This
confirms the statement made previously, that for the axisymmetric imperfections the half-
wave cosine representations is more suitable; that is, it represents adequately the significant
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

features of the measured imperfections with fewer Fourier coefficients.


The nondimensional discrete axial power spectral density for the asymmetric components
with nine circumferential full waves is shown in Figure 10.36. In this case the half-wave

o HALF-WAVE FOURIER COSINE


0 HALF-WAVE FOURIER SINE
2
6 • 0.0046-AOOT MEAN SQUARE

!
<I)
10" 3

::
~ 10" 4
~

SPATIAL FREQUENCY, (&j • •t, el


Figure 10.35 Discrete axial power spectrum for n = O-shell LA-I (from [10.12])

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Evaluation of Imperfection Data 835

10

a HALF·WAVE FOURIER COSINE


0 HALF~WAVE F()I,JRIER SINE
2
6 • 0 18S -ROOT MEAN SQUARE

SPATIAL FREQUENCY,"' • k/
'ct

Figure 10.36 Discrete axial power spectrum for n = 9-shell LA-1 (from [10.12])

sine Fourier representation is the more suitable one because when it is used all but the
lowest-frequency components arc insignificant. The zig-zag nature of the PSD functions
shown is due to the fact that in the axial direction the measured initial imperfections are
essentially symmetric with respect to the middle plane of the shell. The root-mean-square
values of the measured imperfections for (' = 0 and (' = 9 are given on the respective figures.
Although the above format of presenting the results of imperfection surveys is not uni-
versally accepted, there is an increasing number of investigators who report the results of
their imperfection measurements in a similar format. See for example publications by the
Caltech group [9.194], [9.221], [1.25], [10.5] and [10.13], by research workers at the Tech-
nion [9.274], [10.2], [10.21], [10.31] and [10.32], at TU-Delft [10.14] and [10.15] at the
University of Surrey [13.8] and [13.109], at the University of Glasgow [13.7], at the Imperial
College [13.69], [10.28], [10.33]-[10.35], at the University of Trondheim [10.36]), at the
INSA in Lyon [9.199] and [10.37] and at the Politecnico di Milano [10.38], just to name a
few.
In investigating the effect of initial imperfections on the stability behavior of thin-shell
structures, one should always remember that Koiter's work [2.36], [2.89] and [3.15] has
shown conclusively that the knock-down factor depends not only on the magnitude but also
on the shape of initial imperfections. Thus, it is not sufficient to spot check the shell surface
for the maximum imperfection amplitude by carrying out selected circumferential and/ or
axial scans. One must always provide for sufficient cross-reference data, so that later the
individual scans can be pieced together to a complete imperfection map of the measured
structure via numerical techniques on a digital computer. Only after this step has been suc-
cessfully completed can one proceed to eliminate the rigid body displacements by finding
the best-fit cylinder to the measured data of the complete initial imperfection scan, followed
by a harmonic analysis to determine the coefficients of the chosen Fourier representation.
Hereby it must be remembered that the number of Fourier coefficients to be calculated may
not exceed the total number of measured discrete data points.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
836 Initial Imperfections

10.9 Characteristic Initial Imperfection Distributions


Essentially two new approaches are conceivable whereby the measured initial imperfections
can be included directly into the design process. For a prototype the imperfections can be
measured experimentally, and then they can be incorporated into the theoretical analysis to
predict the buckling load accurately. Although there have been some attempts in this direction
(see for example [13.41] and [10.39]), this approach is not suited for the prediction of the
buckling load of shells manufactured in normal production runs. For these shells one must
try to establish the characteristic initial imperfection distribution that a particular fabrication
process is likely to produce, and then combine this information with a statistical analysis of
both the initial imperfections and the corresponding critical loads, a kind of statistical im-
perfection sensitivity analysis [10.40]-[10.42].
The use of probabilistic buckling prediction methods depends strongly on the availability
of the so-called characteristic initial impe1fection distributions. Thus, the critical question is,
Can one associate characteristic initial imperfection distributions with a specified manufac-
turing process? That the answer to this question is an unconditional "Yes" will be demon-
strated by a few examples.

10.9. 1 Laboratory Scale Shells


Figure 10.14a shows the measured initial imperfections of the integrally stringer stiffened
shell AS-2, which has been tested at Caltech [1.25]. Figure 10.14b shows the measured
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

initial imperfections of a similar shell KR-1 tested at Technion [ 10.2].


For the sake of comparison, Figures 10.3 7 and 10.38 display the variation of the calculated
half-wave sine Fourier coefficients as a function of the circumferential wave number £ for
selected axial half-wave numbers k for these shells. As one can see, in both cases the
amplitudes of the Fourier coefficients decay with increasing wave numbers both in the axial
and in the circumferential directions. The Donnell-Imbert [10.43] analytical imperfection
model

(10.8)

where the coefficients X, r and s are determined by least squares fitting the measured initial
imperfection data displayed in Figures 10.37 and 10.38, represents the variation of the har-
monic components with axial (k) and circumferential (~ wave numbers satisfactorily.
Since both shells were machined out of seamless thick-walled aluminum alloy tubing in
different parts of the world, the imperfection model given by Eq. (10.8) represents the char-
acteristic imperfection distribution for this fabrication process.

Circu•f•~rtntill w1v1 nu•ber f


0 _ 4 ,_..~........_~.,..;10;:.~~..:.:1S:,.___..;2;.:0_......;2=..S_ _.:;;10'-

Cl 1\:1
• k= 2
• k=3

Figure 10.37 Circumferential variation of the half-wave sine Fourier representation-shell AS-2 (from
[1.25])

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Characteristic Initial Imperfection Distributions 837

Circumferonti•l wove number~

08 ~~~~~~~o~~~s~~2~o~~2~s~~3o~

~06
....
a
0
k='
II= 2
" • ~=3

Figure 10.38 Circumferential variation of the half-wave sine Fourier representation-shell KR-1 (from
[10.2])

10.9.2 Full-Scale Shells-Riveted Seams


We tum now to large-scale or full-scale tests. Figure 10.39 shows the three-dimensional plot
of the measured initial imperfections of a large-scale shell (radius 945.8 mm, length 2743.2
mm, wall thickness 0.635 mm) tested at Georgia Tech. [13.87]. This shell was assembled
from six identical longitudinal panels and reinforced by 312 closely spaced Z-shaped string-
ers on the inside. The edges of the panels were joined by offset lap splices with stringers
riveted along each joint line. The shell was held circular by means of heavy rolled bracket-
shape external frames located 3.175 mm from each shell end. In addition, seven Z-shaped
equally spaced rings were riveted to the outer skin. As can be seen from Figure 10.40,
amplitudes of the Fourier harmonics with a single half-wave in the axial direction have two
distinct maxima, one at C = 2 (out-of-roundness) and another at e = 6 (number of panels
the shell was assembled from). The Fourier coefficients with more than a single half-wave
in the axial direction are in comparison much smaller.
Figure 10.23 shows the measured initial imperfections of the ARIANE interstage II/III
shell AR 23-1 [10.14]. This shell was assembled out of eight identical longitudinal panels.
Adjacent panels were joined by offset lap splices and one of the 120 equally spaced hat-
shaped stringers was riveted along the joint line on the outside. As mentioned earlier, the
shell was held circular by two precision-machined end rings on the outside and five equally
spaced bracket-shaped rings on the inside. As can be seen from Figure 10.41 the amplitudes
of the Fourier harmonics with a single half-wave in the axial direction this time have a

Figure 10.39 Measured initial shape of Horton's shell H0-1 (from [13.87])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
838 Initial Imperfections

Circumferential wave number I


10 15 20 25 30
""'"
g 2.0
" 1\ =1
~ 0 1\ = 2
-t..
c.
61\ =3

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
.§ 1.0

Figure 10.40 Circumferential variation of the half-wave sine Fourier representation-shell H0-1 (from
[10.14])

distinct maximum at e = 8 (number of panels the shell was assembled from). There is also
a sizable f = 2 (out-of-roundness) component. All other Fourier coefficients are in compar-
ison much smaller.
Thus, it appears from the results presented in Figures 10.40 and 10.41 that for full-scale
aerospace shells assembled out of a fixed number of curved panels, the initial imperfections
will be dominated by two components only, if the joints are riveted. Using the half-wave
sine axial representation, both component~. will have a single half-wave in the axial direction
and, respectively, two and NP full-waves in the circumferential direction, where NP is the
number of full-length panels out of which the shell was assembled. By the use of accurately
machined rigid end rings the f = 2 (out-of-roundness) component can be significantly re-
duced in size.
The variation of the measured Fourier coefficients with axial half-wave (k) and circum-
ferential full-wave (C) numbers can be approximated by expressions of the following type:
- 2
gk,- Vck, + Du- k-
2- I
{
(n, -
x,
-
~2 + 2~,f2 + (n2- C)2 +
x2
-
2~2£2
} (10.9)

where the coefficients X,, X2 , r, n,, n2 , ~~ and ~2 are determined by least squares fitting the
measured initial imperfection data displayed in Figures. 10.40 and 10.41. Thus Eq. (10.9)
represents the characteristic imperfection distribution of full-scale aerospace shells assembled
out of a fixed number of full-length panels by riveted joints.

10.9.3 Full-Scale She/Is-Welded Seams


For full-scale shells that have been assembled out of curved panels and dome like end pieces
using welded seams, only a limited number of imperfection surveys are available. Looking

Circumferential 11ave number I


10 15 20 25 30

""'c" o K =1
0 K=2
0

u c. K=3
-t.. 0.5
c.
E

..
c

"' 0.1
·~
w 0 *"l.-___Ji~~'ti£~~~~-~dHi~

Figure 10.41 Circumferential variation of the half-wave sine Fourier representation-shell AR 23-1
Copyright Wiley
(from [10.14])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Characteristic Initial Imperfection Distributions 839

at the three-dimensional plots of the measured initial imperfections shown in Figures 10.16,
10.20 and 10.21 makes it evident that these imperfection distributions belonging to the
Langley shell LA-1 [10.12] and the aerospace shells X-1 and X-2 [10.14] are rather com-
plicated. This impression is reinforced when one considers the variation of the corresponding
half-wave sine Fourier coefficients displayed in Figures 10.32, 10.42 and 10.43.
The dominant imperfections harmonics have a single half-wave in the axial direction. The
number of full waves in the circumferential direction was obviously influenced by the number
of full-length longitudinal panels used to assemble the shell. Further, it appears that the
welding procedure used also had a strong influence on the resulting number of circumfer-
ential full waves.
Both the Langley shell LA -1 and the first aerospace shell X -1 were assembled out of three
full-length curved panels. Both have negligibly small out-of-roundness (the e = 2 Fourier
coefficients arc negligibly small), though for different reasons. For the LA-1 shell the cir-
cumferential reference surface was provided by the two very accurately machined (3000.0
± 0.15 mm) rigid end rings, whereas the out-of-roundness of the aerospace shell X-1 was
practically eliminated by force-fitting the two circular end domes into the cylindrical body
after welding. However, for the Langley shell LA-1 the largest Fourier coefficient has nine
full waves in the circumferential direction (three times the number of welded seams), whereas
for the first aerospace shell X-1 the largest Fourier coefficient has three full waves in the
circumferential direction (equal to the number of welded seams).
We turn now to the second aerospace shell X-2. This shell also has a large Fourier co-
efficient, with the same number of full waves in the circumferential direction as the number
of full-length longitudinal curved panels out of which it was assembled, namely four. How-
ever, the initial imperfection distribution of this shell also has large harmonics, with e = 2
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(out-of-roundness) and e = 6 full waves in the circumferential direction, besides the har-
monics that are integer multiples of 4, the number of welded seams.
The comparatively large out-of-roundness component is easy to explain. With this shell
the end domes were butt welded to the cylindrical parts. That the harmonics that are integer
multiples of 4 (the number of welded seams) have significant amplitudes agrees with the
results of the Langley shell LA-1. However, up to now no explanation has been found for
the significant size of the e = 6 harmonic.
Thus, for the shells assembled out of a fixed number of curved panels by welded longi-
tudinal seams an important question remains to be answered, namely, when is the number
of circumferential full waves of the dominant Fourier coefficient equal to the number of
welded longitudinal seams and when will it be equal to the number of welded longitudinal
seams times an integer?
The above examples demonstrate unequivocally that characteristic initial imperfection dis-
tributions can indeed be associated with the different fabrication processes. Thus, it is clear
that further advances toward more accurate buckling load predictions of thin shells depend
on the availability of extensive information about realistic initial imperfections and their

Circumferential wave number I


10 15 20 25 30

o K=1
o K= 2
t>. K= 3

Figure 10.42 Circumferential variation of the half-wave sine Fourier representation-shell X-1 (from
Copyright Wiley [10.14])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
840 Initial Imperfections

Circumferential wave number f


10 15 20 25 30
...,"
" 0.5
.S! " K =1
u
o K =2
-t.. A K =3
c.
E

Figure 10.43 Circumferential variation of the half-wave sine Fourier representation-shell X-2 (from
[10.14])

correlation with the different manufacturing techniques. Hence the need for the establishment
of an International Imperfection Data Bank, which will be discussed in the next section.

10.10 Imperfection Data Banks


At present the International Imperfection Data Bank consists of two branches, at the Faculties
of Aerospace Engineering of Delft University of Technology and at the Technion in Haifa.
The purpose of creating the International Imperfection Data Bank is twofold:
1. All the imperfection data obtained at different laboratories by different investigators are
presented in identical format. This makes comparison and critical evaluation possible,
resulting in characteristic imperfection distributions for the different manufacturing pro-
cesses used.
2. For those who want to carry out correlation studies using the advanced nonlinear shell
analysis codes on today's powerful mpercomputers, the much-needed realistic imper-
fection distributions are made available.
For each shell the data are stored using the following uniform format of data presentation
in five separate files:
1. COMMENTS: This file is the "identity card" of the shell. It includes the shell geometry,
boundary conditions, buckling loads (theoretical and experimental values), method of
fabrication, material properties and other relevant information needed for the interpre-
tation of the imperfection data by other users.
2. ADJUSTED MEASURED INITIAL SHAPE w(i, j): This file is obtained using a data
reduction code that recalculated the recorded measured displacements with respect to
the best-fit reference surface, thus removing any rigid body displacements of the shell
with respect to the scanning system.
3. Ak,, Bk, COEFFICIENTS of the half-wave cosine Fourier representation (Eq. 10.2).
4. Ck,, Dk, COEFFICIENTS of the half-wave sine Fourier representation (Eq. 10.1 ).
5. ACTUAL MEASURED (raw) DATA from the scanning device v(i, j), which was used
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

as the input to the data reduction process.


Besides contributions by Caltech [9.194], [9.221], [10.5], [1.25], Technion [9.274], [10.2],
[10.21], [10.31], [10.32] and the TU-Delft [10.44]-[10.47], the International Imperfection
Data Bank contains results of initial imperfection surveys carried out at the University of
Glasgow [ 13. 7], Det norske Veritas [ 10. 10] and others.
It appears that till the mid-eighties there was extensive data collection activity in many
groups and organizations. One might say that considerable investment in the International
Imperfection Data Bank was accomplished. But then the momentum waned, perhaps because
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Probabilistic Design Methods 841

the industry did not want to have its production processes evaluated openly (as suggested
by Arbocz in [ 10.48]), or perhaps because the industry and many researchers believed that
faster and larger computers would provide the answers. But unfortunately, this is definitively
not the case. If the initial imperfections have to be guessed, then one cannot expect that the
computations will provide accurate and reliable buckling loads. To achieve reliable predic-
tions, the design engineer must be able to input both the shape and the sizes of the expected
initial imperfections closely enough to simulate numerically the real behavior of the structure
in question under loading. For accurate imperfection estimates, the engineering community
must have access to initial imperfection data banks containing characteristic initial imper-
fections classified according to fabrications processes.
As emphasized by Singer and Abramovich in 1995 [I 0.11], "Thus today we are at the
crossroads. If the initial imperfection shapes are known, or can be reliably estimated, and
the boundary conditions (including the load eccentricity) are well defined [see Chapter 11
for further details], adequate analytical tools and computer codes are available for calculation
of the buckling load of the structure. But if we do not know the initial imperfections and
the boundary conditions, we cannot improve our predictions of the buckling loads, no matter
how sophisticated our codes are and how large and fast our computers become!"
However, as pointed out by Babcock in 1983 [1.18], "[U]nless better use can be made of
the imperfection data immediately by the experimenter, the motivation for taking the data
will rapidly die out. This is particularly true of tests carried out to meet an immediate industry
need." Luckily the emerging field of probabilistic design methods [ 10.49]-[ 10.51] provides
the means for accomplishing this.

10.11 Probabilistic Design Methods


Since initial imperfections are obviously random in nature, some kind of stochastic stability
analysis is called for. The buckling of imperfection sensitive structures with small random
imperfections has been studied by several investigators, such as Bolotin [10.52], Fraser and
Budiansky [10.53] and Amazigo [10.54], just to name a few. In the absence of experimental
evidence about the type of imperfections that occur in practice and in order to reduce the
mathematical complexity of the problem, all the above-named investigators have worked
with some form of idealized imperfection distribution.

10.11.1 Closed-Form Solution


Thus, Roorda and Hansen [I 0.55] m 1972 assumed an axisymmetric imperfection of the
form
- - X
w(x) = h/; 1 cos i"1TL (I 0.10)

where

l.- L {2;:
n-;{Rh and c = V3(1 - v 2 ) (10.11)

They used one of Koiter's formulas [2.89], Eq. (5.4):

(10.12)

as the nonlinear transfer function between the imperfection I; 1 and the normalized buckling
load A ( = O"! O",, where O",, = Eh/ cR). Assuming that I; 1 is a random variable X 1, then A must
also be a random variable A. Hence
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
842 Initial Imperfections

(10.13)

or
2(1 - A)"
(10.14)
3cA
The reliability R(A) is defined as the probability that the (random) buckling load A will
be greater than some specified value A. From the transfer function this is equivalent to the
probability that the absolute value of the (random) imperfection X 1 will be less than the
value given by Eq. (10.14). Hence
W} - <X <-

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
2(1 -
R(A) = Prob (A 2: A)= Prob { IX1 1 s; cA = Prob (-g* 1 {;"*) (10.15)
3
where

-*
g -- 2(1 -
~-.:_
w (10.16)
3cA
If, following Roorda and Hansen [10.55], one assumes that the random variable X 1 is nor-
mally distributed, then its probability density is given by

fx(x)=
1 { 1 (x-
.';;:;-.=exp --2 2
mJ} (10.17)
u, v 2~rr u,
where
m, = mean value of X 1 = E(X 1)
u, = standard deviation of X 1 = ~(X 1 )
and one can write

R(A) = Prob (A > A) = Prob CIX 1 :s: 1 g*) = f~, fx(x) dx (10.18)
- -
R(A) = Prob ( -{;"* :s: x 1 :s {;"*)

(10.19)

where

erf (x) = error function of x = -


1
\12; ()
- J' e -r' 12 dt (10.20)

As has be~n s~own by Elishakoff [ 10.56] in the particular case when the interval of inte-
gration (- {;"*, {;"*) is symmetric about the mean value of X 1, one can write
where s > 0 (10.21)
and recalling that erf (-x) = -erf (x), Eq. (10.19) takes the form

R(A) = Prob CIX 1 - mxl < s) = 2 erf (;,) (10.22)

In particular, if s = ju,, where j is an integer, then

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Probabilistic Design Methods 843

R(A) = Prob CIX - mxl < jcrJ = 2 erf (j) (10.23)


and erf (j) can be read from standard tables (see [10.57], for instance).
Using an isotropic shell with the dimensions L = 141.0 mm, R = 101.6 mm, h = 0.2634
mm and Jetting v = 0.3 yields (,. = 16.0. Assuming that the shape of the initial imperfection
is given by Eq. (10.10) and that its amplitude is a normally distributed random variable with
a mean of m, = 0.1 and a standard deviation of cr, = 0.05, one obtains for the reliability
R(A) the solid curve shown in Figure I 0.44.
As can be seen from this figure, the availability of the reliability function R(A) for a certain
fabrication process makes it possible to determine the allowable load, defined as the nor-
malized load level A, for which the desired reliability is achieved, for the whole ensemble
of similar shells produced by the same fabrication process. Notice that the normalized al-
lowable load level A, plays a similar role to the so-called knock-down factor y used in the
deterministic buckling load calculations to identify the lower-bound design curve shown in
Figure 3.1. Thus, to make practical application of this approach possible, the user must be
able to calculate the appropriate reliability function R(A) once the characteristic initial im-
perfection distribution of a specific fabrication process has been identified (see also Figure
10.44).

10.11.2 Monte Carlo Method


It was not obvious how the methods used by the early investigators could be extended to
the general imperfections observed in practice. Thus, it was not until 1979 that a method
was proposed by Elishakoff [10.58] that made it possible to introduce the results of the
experimentally measured initial imperfections routinely into the analysis.
Basically, Elishakoff suggested utilizing the Monte Carlo method to obtain the reliability
function R(A) for a specified shell structure produced by a fabrication process with a known
characteristic initial imperfection distribution. The relative ease with which one can apply
this procedure, once a sufficiently large sample of initial imperfection measurements is avail-
able, will be demonstrated for the case of axially compressed cylindrical shells with random
axisymmetric initial imperfections.
The measured initial imperfections of a sample of M shells are represented by
N
- 'V X
w(x) = h ~ Ai'" 1 cos i7r- (m = 1, 2, ... , M) (10.24)
1~1 L
where now the Fourier coefficients Ai"' 1 are the random variables. One proceeds by first
calculating, by taking ensemble averages, the estimated mean of the Fourier coefficients

R(}..l:Prob(A>\ I

rz
Requi.rad
Reliability
R(}..}a J
-r, t,. (xldx
Monte Corto sim!Jation
0.5

0 :-------:~--~~-::'::---~-==---~-}..
0 l.a 0.5 10
\... Allowable buckling load

Figure 10.44 Comparison of analytical reliability function with results of the Monte Carlo method-
m, = 0.1, a, = 0.05 (from [10.59])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
844 Initial Imperfections

- 1 M
A'")""-
M m~=l
1
A 2: (10.25)

and then the estimated variance-covariance matrix


1 ~ - -
ule)
11
= - - - L..J [A'"') - A re'][A
1
rmJ -A lei]
1
(10.26)
M - 1 m= 1 1
.1

Since uit is a positive-definite symmetric: matrix, it can be decomposed into a product of


lower and upper triangular matrices by Cholesky decomposition:
u•e)
•}
= ccr (10.27)
where C is a lower triangular matrix. Next the vector A of the simulated initial imperfections
is obtained as
(10.28)

where A 1" 1 = estimated mean vector


r = random vector
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

The r is a vector consisting of normally distributed random numbers with zero mean and
unit variance computed by standard random number generators. Using, for example, a ran-
dom vector r of length 1089 X N, one then obtains 1089 simulated imperfection vectors A
of length N, that is, 1089 different simulated shells, with the A's containing the Fourier
coefficients of the corresponding initial imperfections. With these simulated imperfect shells
(which are statistically equivalent to the initial experimentally measured sample) one can
proceed to carry out repeated buckling load calculations generating the histogram of buckling
loads shown in Figure 10.45. Since the reliability function R(A) has been defined as the
probability that the random buckling lmd A will exceed the prescribed value, one then
proceeds to calculate R(A) from the histogram of the buckling loads by the frequency inter-
pretation (i.e. fraction of an ensemble of shells of which the buckling loads exceed the
specified load). The accuracy of the Monte Carlo method can be seen in Figure 10.44 from
the close coincidence of the dots, showing the results obtained via the Monte Carlo method,
with the solid line representing the analytical solution of [10.59].
The accuracy of the Monte Carlo method can further be improved by increasing the
number of simulations. To estimate the maximum difference between the exact and the
simulated reliability curve one can use the Kolmogorov-Smirnov test for goodness of fit
[10.60]. It has been shown by Elishakoff [10.56] that the maximum absolute difference
between the exact and empirical reliability functions is proportional to 1/~, where M, is
the number of simulations used.
It is evident from the initial imperfection surveys presented earlier (see Figures 10.12,
10.14, 10.16, 10.18, 10.21, 10.23 and 10..39) that in a realistic stochastic stability analysis
one must include both axisymmetric and asymmetric imperfections. Using the so-called
multi-mode analysis [10.61], Elishakoff and Arbocz have demonstrated the feasibility of
using the Monte Carlo method to derive functions for very general imperfections. As can be

"'
"~
". 500
~
.0

0
~
.0
E
~
z
0 >.
0 0.2

Figure 10.45 Histogram of the normalized buckling loads A (from [10.59])


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Probabilistic Design Methods 845

seen from Figure 10.46 (here reproduced from [10.62]) the inclusion of asymmetric com-
ponents results in a lower allowable load level A, (see curve 2) for a specified reliability of
0.98, say, than for the case of axisymmetric imperfections only (see curve 1). Thus, the
reliability estimates depend strongly on the number of terms taken into account in describing
the initial imperfection.
The importance of using the correct deterministic model in a probabilistic analysis to
describe the mechanics of the problem being analyzed cannot be stressed enough. Thus, if
in the solution presented by Roorda and Hansen [10.55] one would replace Eq. (10.12) by
one of the other formulas presented by Koiter in [2.89], then in the region of high reliability
the shape of the corresponding reliability functions R(A) would change noticeably. As one
can seen in Figure 10.47 for a specified reliability of 0.98, say, one would obtain different
values for A,, ranging from 0.17 to 0.35.

10.11.3 Response Surface Method


Clearly, if for each application one could quantify and understand the problem uncertainties
and their influence on the design variables, one would obtain a better engineered, better
designed and ultimately a safer product. The reliability-based design approach basically
provides the means to achieve this goal. For instance, the collapse problem of axially com-
pressed stiffened cylinders can best be formulated in terms of a response (or limit state)
function
g(X) = AJX) - A (10.29)
where A is a suitable normalized load parameter ( = PIP", say), A, is the random collapse
load of the structure and the vector X represents the random variables of the problem. The
components of the random vector X; may be Fourier coefficients of the initial imperfections
and other parameters quantifying the uncertainties in the specified boundary conditions, the
constitutive equation used to describe the nonlinear material behavior, thickness distribution,
residual stresses, etc. Notice that the evaluation of the response function thus involves the
solution of a complicated nonlinear structural analysis problem, represented by a detailed
and possibly large finite element model. However, with today's computational resources this
in itself should not pose any unsurmountable difficulties.
Notice that the response function g(X) = 0 separates the variable space into a safe region
where g(X) > 0 and a "failure region" where J?(X) s 0. The reliability R(A) or the probability
of failure Pr(A) can then be calculated as
R(A) = 1 - P 1 (A) (10.30)
where

Pr(A) = Prob{g(X) s 0} = r··f


g( \:)SO
fx(x) dx (10.31)

and f x(X) is the joint probability density function of the random variables involved.
The credibility of this approach depends on two factors, the accuracy of the mechanical
model used to calculate the limit state function and the accuracy of the probabilistic tech-
niques employed to evaluate the multidimensional integral.
Whereas with today's advanced nonlinear finite element codes like STAGS [2.53] and
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

ABAQUS [2.102], the limit state function (if so desired) can be determined with great ac-
curacy, the evaluation of the multidimensional integral Eq. (10.31), where the domain of
integration depends on the properties of the limit state function, is still the subject of detailed
investigations. Since an exact numerical evaluation of this multidimensional integral is con-
sidered impractical, major developments are concentrated on approximate (first and second
order) and accurate (efficient Monte Carlo simulation) reliability methods [10.63].
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
846 Initial Imperfections

R(k) =Prob(A>k)

1.0---

Required
Reliability
curve 1

0.5

1.0

Figure 10.46 Reliability functions for simulated group of 500 B-shells (from [10.62])

For a successful application of reliability-based design methods the existence of a test-


originated database is very helpful. It can guide users in choosing the appropriate probability
density distributions for the input variables. If test data are not available, then the user must
select probability density distributions based on his judgment. In such cases the availability
of design software for calculating the parameter distribution sensitivities can be of great help
in making the proper choice.
Last but not least, before large-scale acceptance of reliability-based probabilistic design
methods by the engineering community is to occur, true reliability must be demonstrated
and not simply estimated from engineering analysis. Thus, it is not surprising that the first
successful integration of probabilistic de~.ign methods with existing design processes has
been accomplished in applications like the design of specific components of aircraft engines,
where failure and failure rate databases are available. It is also interesting to note that in the
cases reported by Fox [10.64], the probabilistic approach was built around the existing design
methodology. It is part of an integrated process, rather than requiring the designer to make
specific stress analysis runs and then enter the data and run the probabilistic code in an
iterative loop. The probabilistic output is printed in addition to the normal deterministic
output, and it contains information about the model accuracy, so that the designer can decide
whether the results from the probabilistic analysis are acceptable.
In all those applications where failure and failure rate databases are as yet not available,
the probabilistic methods can best be utilized to great advantage as a design tool to help
identify the sensitivities of problem parameters.
To facilitate the use of probabilistic methods in the design process, one can best employ
a hierarchical approach as proposed in [3.18]. In a hierarchical approach, initially simple
mechanical and stochastic methods of known accuracy are used to carry out the necessary

Rt:l.l =Prob IA>:I.l

Required
Reliability

0.5
- - - Eq. {5.51
Eq. {5. 7)

Eq. 15.4)
l of Ref • 2.89

0 L---~!..!...~:........--'-~~--==-_._....;.:1.
0 l.8 0.5 1.0
"- J~llowable buckling load
--`,`,`````,`,````,,``,``,,`-

Figure 10.47 Reliability functions using different deterministic models


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Residual Stresses 847

parameter studies. This phase includes the evaluation of the sensitivity derivatives needed to
decide which arc the important random variables that must be included in the refined reli-
ability calculations. Once the principal dimensions and the layout of the structure have been
fixed, the structural reliability of the final design should be calculated by carrying out a
detailed analysis employing a refined finite-element model to describe the mechanical be-
havior and an accurate reliability method to complete the probabilistic calculations.

10.12 Residual Stresses


Large metallic structures are usually assembled by welded scams. The nonuniform cooling
associated with this fabrication process leaves the finished structural members not only with
initial distortions of the nominal shape but also with a set of self-equilibrating stresses, called
residual stresses, locked into them. As pointed out by J.B. Dwight and K.E. Moxham in
1969 [8.46], the continuous longitudinal or circumferential welds are invariably stressed up
to yield in tension, with the areas carrying this locked in tension extending typically two to
four thicknesses from the weld on each side.
The rest of the section must then be in a state of residual compression in order to preserve
longitudinal equilibrium.
In Figure 10.48, reproduced from [8.46], a typical residual stress pattern in a welded box
member is displayed. Notice that the residual stress patterns along the four webs with edge
welds can be idealized into the residual stress pattern shown in Figure 10.49. The yield stress
tension blocks of width YJh at the edges are then balanced by the average stress a, arising
in the compression zone. From longitudinal equilibrium along the weld one then obtains the
expression

cr,. = a- (10.32)
b/h - 2YJ .\
Studying the relation between the size of the weld and the resulting residual stresses,
Moxham found that for a plate of given thickness h, with a given thermal history, the width
of the tension block YJh is largely independent of the overall width of the plate b, provided
b I h > 25. Thus, the problem was reduced to relating the value of YJh to the weld size. It
was shown by Dwight and Moxham in the Appendix to [8.46] that when two or more plates

c."'
3
l
2
'it-
TON/ tN 2:
~ 6
I TENSION
\. :r-r-
\
()
0
) ;::
'tl
D

"'
II>
II>
5
z
,----..:

~ ~ 0
0 0 0
1
N/mm

\
Figure 10.48 Typical residual stress pattern in a welded box member (from [8.46] by permission of
the Institution of Structural Engineers) (b/ h = 80, u,. = 402 N /mm')
Copyright Wiley
--`,`,`````,`,```

Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
848 lnitiallmperfections

.--~ r

cry z
0
v;
~
....
crr
I
CCIMPRE SSION

,_ \- 1 -1 r--
b

Figure 10.49 Idealized residual stress pattern in a web with edge welds (from [8.46] by permission of
the Institution of Structural Engineers) (definition of Yf)

meet at a weld, the value of YJh is the same for each plate regardless of differences m
thicknesses and that their value is given approximately by
CA
YJJhl = YJ2h2 = ... = "'h. (10.33)
(J\'L.._; I

where L.h, = sum of the plate thicknesses (mm)


A = cross-sectional area of added metal (mm2 )
C = empirical coefficient, with dimensions of stress ( ~6000 N I mm 2 )
Suppose one joins two webs of equal thickness by a fillet weld of leg w and takes A = 0.6
w 2 to allow for a little convexity of the weld. Then by combining Eqs. (10.32) and(l0.33)
one obtains

(10.34)
lr,blh
1
0.6C(w!hf-
This is a relatively simple formula where the empirical constant C must be determined by
experiments. Based on available experimental results, Dwight, and Moxham recommend the
use of C = 6000 N/mm 2 •
The simple analysis of Dwight and Moxham presented here only gives a guide for the
residual stress levels likely to occur. Its accuracy depends on a number of simplifying as-
sumptions and is, strictly speaking, only applicable to a continuous, single-pass fusion weld
made by any of the common processes. The method should not be used for multi-pass welds,
for which it is known that the residual stresses will be less severe. For a more detailed study
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

of the origin and evaluation of residual stresses the interested reader should consult [10.65].
The experimental determination of residual stresses usually involves the measurement of
the strains induced by welding and the use of theory of elasticity to calculate the corre-
sponding self-equilibrating stress fields. To obtain the strains, one has to measure the distance
between certain specified reference points before and after the welding has taken place. Two
different experimental techniques have been in use.
Residual stresses can be determined by the direct measurement method, whereby strain
changes due to welding are calculated from measurements made on the surface of the struc-
ture before and after welding. As pointed out by R.J. Denston and J.D. White [8.68]
The method is simple and works well provided that
(i) measurements are made in the regions that remain elastic, and
(ii) the plate is initially stress free or the initial stress pattern is known.
This last restriction is necessary since one measures changes in strain and not absolute values
of strain. In addition, one must remember that stresses cannot be obtained in regions that
have yielded.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Residual Stresses 849

An alternative way for determining residual stresses is the sectioning method, in which
the welded plate is cut up into short narrow pieces and the strain change in each strip is
measured as the initial stresses are relieved. Notice that the strain relaxation is elastic and
thus residual stresses can be calculated using elasticity theory, even though the measurements
may have been taken in regions that have yielded.
A comparison between results obtained by the two methods has been reported by Denston
and White [8.68] and is shown in Figure 10.50. The authors point out that in order to
guarantee that the specimen was stress-free before the weld was applied, the specimen used
for this experiment was sawn from an as-rolled plate so as not to cause residual stresses and
then was stress-relieved before welding to remove the initial stresses. With this precautionary
measure Denston and White have accounted for the fact that by the direct measurement
method only the changes of strain due to welding are determined, whereas by the sectioning
method the total strain in the specimen, including the initial stresses, is measured. Thus, it
is not surprising that Figure 10.50 shows an excellent agreement between the results obtained
by the two different methods away from the weld.
Originally surface strains were measured by Demec gages like the one shown in Figure
10.51. For use on steel specimens, the Demec gage can sit either in holes in glued-on
discs-the same technique as used on concrete-or in holes drilled directly in the specimen.
Although the dial gage had to be read for each setting separately and the results recorded
manually, thus making the determination of residual stresses a tedious and lengthy operation,
several interesting investigations have been carried out. Thus, for instance, in 1977 Sum-
mervile, Swan and Clarke [10.66] reported on the results of an investigation where residual
stresses and initial distortions in welded structures were measured during the manufacture
of warship hull sections. Their paper was intended to provide information for use in future
design calculations of the buckling strength of ship hulls.
In another study, carried out at Lehigh's Fritz Engineering Laboratory, Ostapenko and
Gunzelman [9.172] investigated the local buckling behavior of tubular steel columns fabri-
cated from high-strength steel by cold-rolling and welding. To determine the residual stresses
in one of the tubes tested (specimen P5), target holes were drilled 0.254 m apart in the
longitudinal direction on the inside and outside surfaces as shown in Figure 10.52. After the
specimen was rolled, the distance between these holes was measured before and after weld-
ing. The calculated longitudinal residual stresses are displayed in Figure 10.53.

x Direct measurement

a Sectioning I
I.
I ~Weld Seam
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley Figure 10.50 Comparison of direct measurement and sectioning methods (from [8.68])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Initial Imperfections

Figure 10.51 Dcrncc gage (from [8.68])

More recently, Denston and White at the Cambridge University Structural Engineering
Laboratory have developed a new extensometer, which employs a strain-gaged cantilever as
the measuring element. This system, shown in Figure 10.54, is compact and accurate and
gives an electrical output signal that is suitable for automatic data recording. When used for
strain measurements. the 1.5-nun diameter steel balls visible on the extensometer are located
onto 1-mm diameter punched indentations on the specimen surface. Notice that these gage
marks are easy to prepare. Further details of the development and the practical use of this
new extensometer system can be found in [8.68].
Finally, in a short note Pineault and Brauss [ 10.67] presented a fully automated, PC-
controlled stress mapping system that employs X-ray diffraction methods for residual stress
analysis. Figure I 0.55 displays a typical result obtained by this system.
The effect of residual stresses on the critical stress and the ultimate load-carrying capacity
of columns of various cross-sectional shapes has been extensively investigated by the dif-
ferent Task Groups of the ECCS (European Convention for Constructional Steelwork). The
resu lts of a systematic theoretical investigation based on experimental results showed a wide
scatter of column strength depending on the cross-sectional shape and the manufacturing
process used (10.68 ]. This resulted in the selection of three representat ive curves displayed
in Figure 10.56. to which the strength of the most commonly used cross-sectional shapes
can be related. In Figure I 0.56 ([ I0.69), p. 58) the following nondimensional parameters are
used

~::,~ -+---11
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Gage
Lines

Figure 10.52 Target holes on test specimen for measuring residual stresses (from [9.172]))
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Residual Stresses 851

o Outside Surfctt
6 ln5ide SUrfoce

I I
I I
I I
I
v -LII\t Opposltt Line Opposite-......J
I
I Weld Seam Weld Seam I
-90
I
- 120 MPoH7.4 0kSI) I
I
Dtnloped Plott Width
·I
Figure 10.53 Welding res idual stresses in specimen ( from [9. I72])

-A=- A ~··
-
L
(10.35)
.,.,£ {J

with
u,., = limit stress
u, = yield stress
p = mi nor radius of gy ration of the cross-section
An excellenl description of the ways curves a. band c can be derived is given in H.G. Allen
and P.S. Bulson [2.3], pp. 128- 15 1
The effect of the residual stresses on the critical stress and the ultimate load-can)'ing
capacity of ax iall y compressed plates was covered in Section 8.1.5. For shell-like structures,

I
Figure 10.54

Copyright Wiley
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

[8.68])

Provided by IHS Markit under license with WILEY


C l\4

'
New extensometer from Cambtidge University Structural Engineeri ng LaboratOry (from

Licensee=McDermott Inc - India/8215328006, User=G, Boopathi


No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
852 Initial Imperfections

s h····· (\&:•' )

....
44
28
12
-4
-28
-3G
-:52
-GO
-84

X scale 5.0 mmjdiv. #X spots 10 T spots 100 Elapsed Time 0 hr 46 min


Y scale 5.0 mm/div. #Y spots 10 P'hi ang. 0.0 Avr. Stress -6.60 ksi Stdv. 37.3

Figure 10.55 Residual stresses in a through-hardened steel plate for hand tool (from [10.67])

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
the effect of residual stresses on the critical stress and the ultimate load can be accounted
for by using one of the finite-difference or finite-element codes, such as STAGS [2.53],
NASTRAN [2.98], ADINA [2.99], MARC [2.100], ANSYS [2.101] and ABAQUS [2.102],
just to name a few.

10.13 Imperfection MeasurE~ments and Data Banks in


Columns and Plates
It was shown in Chapters 2 and 3 that plates have stable postbuckling behavior. As can be
seen from Figure 2.7, both initially fiat and slightly imperfect plates subjected to in-plane
compression will carry additional load after buckling if the unloaded edges are supported.
It is clear, however, that larger initial imperfections will result in lower plastic collapse loads.
Thus although the buckling problem of imperfect plates was solved around 1950 [2.13]
and [2.14], it was not until a series of box girder bridge collapses in the late sixties and
early seventies [10.70]-[10.72] that large-scale research into the effect of initial imperfections
on the stability of box girder diaphragms under various external loads was initiated. To obtain
experimental data for improved design recommendations and provide accurate input data for
the new generation of general-purpose finite-element codes that can handle both geometric

welded
D box-shape
rolled 1-srope
I M>>(2
rolled I -shape
Fl H h/b<!2
w f---..=------~ b T
_
-r-shape
flame-cui fig pi
H
Q8 H v::feaOO,,I;,';';
Q6 a( ~ 9,;;;uape
rolledi-sl>ape w
H ~e:.!t-:0~;;
I !-shape
amealed
0, I welded cover- pi
·• 0 box- shape
annealf:od
I ~~M><u
{ I--l T-shape
Q2 H~'il c welded I -shape
rolled flange pl.
r l U-shape
Dt0--~Q2~~Q~,--Q~5~-Q-8~~1.0~~U~~~-,~~~~~~~.B~X

Figure 10.56 ECCS column-strength curves plotted in nondimensional form (from [10.69])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Imperfection Measurements and Data Banks in Columns and Plates 853

and material nonlinearities, laboratory-scale and (nearly) full-scale test programs were carried
out.
The laboratory-scale tests were designed to achieve conditions conforming as closely as
possible to the requirements of the theory in order to make correlation between theoretical
predictions and experimental results possible. Tillman and Williams [10.73] employed an
imperfection measuring device consisting of noncontacting inductance type transducer
probes, which were arranged in a holding frame placed around the test specimen. The five
probes could be moved simultaneously in the longitudinal direction by sliding the supporting
carriage along accurately machined steel rods, which also served as the reference for the
deflection measurements. Typical initial imperfections are shown in Figure 10.57. The au-
thors reported good agreement between the buckling loads and modes predicted by a com-
puter program based on a finite strip approach and the test results.
In another experimental program, Fok and Murray [8.67] used a newly developed leaf
spring-based transducer, actuated by a piston working against a very light helical spring, to
carry out initial imperfection measurements on a fine rectangular grid. It was found that the
small contact force of each piston on the plate of about 0.002 N had a negligible effect upon
the measured initial imperfections. Figure 8.55a displays a three-dimensional view of initial
imperfections in a typical specimen with blown-up vertical scale. The theoretical predictions,
based on numerical solution of the nonlinear Marguerre equations, including the measured
initial imperfections, agreed closely with the experimental results.
When the investigations of the premature failure of box girder bridges were started, an
accurate determination of the initial imperfections was considered essential [10.74]. In ad-

Contour level key (mm}


n -o.ao
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

21 -0 70
3}- o.60
4 I -O.SO
5} -0 40
6} -0.30
7} -0 20
8} -0.10
91 000
10} 0 10
11} 0 .)0
12} 0.30
131 0 40
14} o.so
151 o.60
16} 0 70
171 0 80
18} 0 90

7/"--_.
OuNard 1mperfechons
7 are pos1hve

Figure 10.57 Initial geometric imperfections on one face of a box section (from [10.73])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
854 Initial Imperfections

dition to laboratory-scale specimens, there was also an interest in determining the initial
imperfections of the actual full-size structures. Thus, a technique was needed that could be
used in both the laboratory and the real-world sites. In addition, measurements would have
to be carried out when the structure in question was virtually inaccessible. Megson and
Hallak [10.75] describe such a system based on electronic theodolites, which became avail-
able around 1983.
The basic components of the system, shown in Figure 10.58, arc two electronic theodolites,
a scaling bar of known length and a microcomputer. The horizontal length £of the baseline
joining the two theodolites and the height difference h between the vertical positions of the
same theodolites must be determined very accurately. First the theodolites are used to mea-
sure the vertical and the horizontal angles between the ends of the baseline and of targets
placed at each end of the scaling bar of known length C (see Figures 10.59 and 10.60). With
this information, one can calculate the length e of the baseline and the height difference h
between the positions of the two theodolites as outlined in [10.75].
The points of the structures where the deflections are to be measured are identified by
paper targets glued to a set of grid points . These targets are then sighted by both theodolites
simultaneously. With the aid of a computer one can then derive from the measured angles
and the previously determined values of Cand h the global coordinates (X, Y, Z) of the point.
Once the measurements have been completed, one can determine the initial imperfections
of the structure by transforming the global coordinates (X,, Y,, Z;) of the measured points
into a local (reference) coordinate system. This local coordinate system defines the surface
of the perfect structure. Thus, for instance, any three of the corner points of a flat rectangular
surface of a structure can form such a local coordinate system (see Figure I 0.61 ).
Figure 10.62, reproduced from [10.75], shows a three-dimensional plot of the measured
initial imperfections of an isolated diaphragm. In Figure 10.63 the measured initial imper-
fections of the same diaphragm are displayed after it has been installed in a part of a box
girder. For both figures the reference frame passes through three corner points.
It is obvious that if another three reference points on the surface of the diaphragm are
chosen the shape of the initial imperfections will be different. It appears that it is more
accurate to establish an imaginary perfect diaphragm obtained by finding the best-fit plane
for all the measured points by the method of least squares (see [9.194]).
With more recent equipment the use of the attached paper targets can be avoided by using
a laser eyepiece fitted to one of the theodolites. This then provides visible light-spot targets,
which can then be intersected by the other theodolite. The accuracy of these systems is ± 0.1
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

mm if the structure is placed at a distance of 10 m.

1 - electronic: theodolite
2 - supplyanddatatransm•ss•oncables
3 - system trolley
4 - microcomputer
5 - printer
6 - scaling bat with known length tor the
determtnattOn of the basehne data
7 - the structure

u--
Figure 10.58 Automated spatial coordinate measuring equipment (from [ 10.75])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Imperfection Measurements and Data Banks in Columns and Plates 855

Is

Theodolite 1

Figure 10.59 Plan view of the spatial coordinate system (from [10.75])

Finally, besides the International Imperfection Data Bank for shell structures, described in
Section 10.1 0, the establishment of other Data Banks for columns and plates has also been
reported in the literature (see for instance [6.125], [6.136], [6.140], [6.141], [10.76]-[10.78]).
For example, the NDSS (Numerical Data Bank for Steel Structures) was developed in the
course of extensive experimental investigations of the buckling and failure behavior of steel
columns, beams and plates at the Civil Engineering Department of Nagoya University in
Japan under the leadership of Professor Y. Fukumoto [10.76]:
Quoting from the introduction of [10.76]:
Many countries are now in the process of introducing new editions of their design specifications
based on modem limit state principles. This has increased the need for a large volume of good
quality test data since such information is essential for the satisfactory calibration of these methods,
so as to insure appropriate level of structural reliability.
The requirement for good quality test data is also expected to increase with the continuing
advancement of powerful, computer-based analytical methods since these require very careful
checking against well documented test results if they are subsequently to be used with confidence.

It appears that the reasons for the establishment of the Nagoya University Data Bank are
identical to those quoted earlier in Section 10.10 for creating the International Imperfection
Data Bank for shells. It seems, however, that this Data Bank also shares the same problems
as the Shell Data Bank discussed earlier, namely the unwillingness of the user community
to cooperate and to share information.

N'
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

1'h•::odo 11 te

Figure 10.60 Elevation of the spatial coordinate system (from [10.75])


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
856 Initial Imperfections

Globalt)( 1 ,Y .l 1 )

.. ,,, ' Local 0.0.01

Global (X 1,Y 1 .Z 1 1
local t • 1.0. OJ
' I
--..
Global lX 1 . Y3 . Z 3 1
local (• 1 .0.z 3 )

Figure 10.61 Layout of spatial coordinate system with global and local coordinate systems (from
[10.75])

Among the other data banks for columns, one may note for instance a data bank on steel-
beam-to-column connections, based on collected test data, developed at the Purdue Univer-
sity Computer Center in the eighties (see [6.136] or [6.141]). This program (SCDB) tabulated
and plotted data for potential design use.

10.14 Concluding Remarks


As pointed out earlier in this chapter, as a means to improve the failure load prediction of
buckling critical structures, the measuring of initial imperfections has been widely accepted
and is being carried out rather routinely. However, due to the manifold scheduling conflicts
on the workfioor, one is often tempted to use shortcuts to save on time and expenses with
such imperfection measurements. Arguing that knowing the shape and size of local imper-
fections is sufficient to make accurate buckling load predictions, one often attempts to carry
out only a limited number of individual wall profile scans at different circumferential loca-
tions without establishing the errors in circularity at sufficient axial stations in order to make
a two-dimensional harmonic analysis feasible.

117

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley Figure 10.62 Measured initial imperfections on isolated diaphragm (from [10.75])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 857

w
0.010

0 OJ

Figure 10.63 Measured initial imperfections on installed diaphragm (from [10.75])

This omission may limit the accuracy of the subsequent numerical buckling load calcu-
lations severely. Only when both local and global initial imperfections are known can one
proceed to make reliable buckling load predictions. Thus, future imperfections surveys should
always include the necessary cross-reference scans in order to make the assembly of complete
imperfection surface maps possible.
Another important feature that is often forgotten is the fact that during the data reduction
process one must initially establish the so-called best-fit shape of the structure by removing
the rigid body components (the f = 1 harmonics) from the measured data (see [10.5] and
[10.28]). Only this important step is completed and the measured initial imperfections are
recalculated with respect to the newly found reference surface can one proceed with the
harmonic analysis to obtain the Fourier harmonics representing the real initial imperfection
components.
There are applied loads, such as external pressure, that make existing asymmetric initial
imperfection shapes grow fast, resulting in a collapse pattern that is similar to the initial
shape. However, there are also cases, such as axially compressed cylindrical shells with many
(nearly) simultaneous buckling modes, where nonlinear interaction plays a major role in
determining the final collapse mode, often resulting in mode shapes that can hardly be
identified in the initial stress-free state. Thus, the prebuckling growth patterns of imperfect
shells under increasing external load arc often quite different from the dominant modes of
the initial imperfection pattern.
In contemplating the measuring of residual stresses, one should always remember that
thermal processes such as welding produce both thermal stresses and thermal deformations
concurrently. Thus, if thermal stresses are recorded, one must also carry out imperfection
scans in order to measure the thermal deformations if an accurate determination of the critical
load is to be obtained.
For an up-to-date review of the state-of-the-art of dealing with the effect of initial imper-
fections on the load carrying capacity of thin-walled structures, the interested reader should
consult [10.77] and [10.78]. [10.77] contains most of the papers presented at the Euromech
Colloquium 317, held at the University of Liverpool on March 21-23, 1994, and [10.78]
contains all the papers presented at the five Stein Memorial Sessions of the 38th AIAA/
ASME/ ASCE/ AHS/ ASC Structures, Structural Dynamics and Materials Conference in Kis-
simmee, Florida, covering a wide range of buckling problems.

References
10.1 Arbocz, J., Shell Stability Analysis: Theory and Practice, in: Collapse: The Buckling of Structures
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

in Theory and Practice, J.M.T. Thompson and G.W. Hunt, eds., Cambridge University Press,
Copyright Wiley Cambridge, 1983, 43-74.
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
858 Initial Imperfections

10.2 Singer, J., Abramovich, H. and Yaffe, R., Initial Imperfection Measurements of Stiffened Shells
and Buckling Predictions, Proceedings 21st Israel Annual Conference on Aviation and Astro-
nautics, Israel Journal of Technology, 17, 1979, 324-338.
10.3 Galletly, G.D. and Bart, R., The Effect of Initial Out-of-Roundness on the Strength of Thin-
Walled Cylinders Subject to External Hydrostatic Pressure, Journal of Applied Mechanics, Trans-
actions ASME, 78, 1956, 351-358.
10.4 Lunchick, M.E. and Short, R.D., Behavior of Cylinders with Initial Shell Deflections, Journal
of Applied Mechanics, 24, 1957, 559-564.
10.5 Arbocz, J., The Effect of General Imperfections on the Buckling of Cylindrical Shells, Ph.D.
thesis, California Institute of Technology, Pasadena, Calif., 1968.
10.6 Coppa, A.P., Measurements of Initial Geometric Imperfections of Cylindrical Shells, AIAA Jour-
nal, 4, (1), Jan. 1966, 172-177.
10.7 Shover, D.R., On the Problem of Geometric Imperfections on Thin Circular Cylindrical Shells,
Ph.D. thesis, Stanford University, Stanford, Calif., 1968.
10.8 Tennyson, R.C., Muggeridge, D.B. and Caswell, R.D., New Design Criteria for Predicting Buck-
ling of Cylindrical Shells under Axial Compression, Journal of Spacecraft and Rockets, 8, 1971,
1062-1067.
10.9 Tennyson, R.C., Interaction of Cylindrical Shell Buckling Experiments with Theory, in: Theory
of Shells, W.T. Koiter, and G.K. Mikhailov, eds., North-Holland, 1980, 65-116.
10.10 Grove, T. and Didriksen, T., Buckling Experiments on 4 Large Axial Stiffened and I Ring
Stiffened Cylindrical Shell, Det norske Veritas, Report No. 76-432, Hovik-Oslo, 1976.
10.11 Singer, J. and Abramovich, H., The Development of Shell Imperfection Measurement Tech-
niques, in: Special Issue on Buckling-·Strength of Imperfection Sensitive Shells, Galletly, G.D.
(Guest Editor), Thin-Walled Structures, 23, (1-4), 1995, 379-398.
10.12 Arbocz, J. and Williams, J.G., Imperfection Survey on a 10-ft Diameter Shell Structure, AIAA
Journal, 15, (7), July 1977, 949-956.
10.13 Babcock, C.D. and Arbocz, J., Imperfection Measurements of Large Scale Shells, GALCIT
Report SM 78-7, California Institute of Technology, Pasadena, Calif., 1978.
10.14 Arbocz, J., Imperfection Data Bank, a Means to Obtain Realistic Buckling Loads, in Buckling
of Shells, Proceedings of a State-of-the-.1\rt Colloquium, E. Ramm, ed., Stuttgart, May 6-7, 1982,
Springer-Verlag, Berlin, Heidelberg, New York, 1982, 535-567.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

10.15 Sebek, R.W.L., Imperfection Surveys and Data Reduction of ARIANE Interstages l!II and II/
III, Ir. thesis, Delft University of Technology, Department of Aerospace Engineering, 1981.
10.16 Tennyson, R.C. and Hansen, J.S., Optimum Design for Buckling of Laminated Cylinders, in:
Collapse: The Buckling of Structures in Theory and Practice, J.M.T. Thompson and G.W. Hunt,
eds., Cambridge University Press, Cambridge, 1983, 409-429.
10.17 Clarke, M.J. and Rotter, J.M., A Technique for the Measurement of Imperfections in Prototype
Silos and Tanks, Research Report No. R565, School of Civil and Mining Engineering, University
of Sydney, March 1988.
10.18 Giavotto, V., Poggi, C., Chryssanthopoulos, M. and Dowling, P., Buckling Behavior of Composite
Shells under Combined Loading, in: Buckling of Shell Structures, On Land, In the Sea and In
the Air, Proceedings International Conference, J.F. Jullien, ed., INSA Lyon, September 17-19
1991, Elsevier Applied Science, London, New York, 1991, 53-60.
10.19 Garkisch, H.D. and Schwarz, H.W., Kapazitive Wegmessung bei Beu1- und Schwingungsunter-
suchungen an di.innwandigen Kreiszylinderschalen, DFVLR Report DLR-FB 64-42, December
1964.
10.20 Rosen, A., Singer, J., Grunwald, A., Nachmani, S. and Singer, F., Unified Noncontact Measure-
ment of Vibrations and Imperfections of Cylindrical Shells, in: Proceedings of the 7th Interna-
tional co11{erence on Experimental Stress Analysis, Haifa, Israel, August 23-27, 1982, 524-538.
10.21 Abramovich, H., Yaffe, R. and Singer, J., Evaluation of Stiffened Shell Characteristics from
Imperfection Measurements, Journal of Strain Analysis, 22, (1), 1987, 17-23.
10.22 Weller, T., Abramovich, H. and Singer, J., Application of Non-Destructive Vibration Correlation
Techniques to the Buckling of Spot-Welded and Riveted Stringer-Stiffened Cylindrical Shells,
Zeitschriftfur Flugwissenschaften und Weltraumforschung, 10, (3), May/June 1986, 183-189.
10.23 Buckling of Thin-walled Cylinders, NASA Space Vehicle Design Criteria (Structures), NASA
8007, revised (1968).
10.24 Rules for the Design, Construction and Inspection of Offshore Structures, Det norske Veritas,
Oslo, 1977.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 859

10.25 DAST-Richlinie 013-Beulsicherheitsnachweise fi.ir Schalen. Deutscher Ausschuss fi.ir Stahlbau,


Cologne, 1980.
10.26 Buckling of Shells-European Recommendations, 4th ed., European Convention for Construc-
tional Steelwork, Brussels, 1988.
10.27 Deutsches Institut fi.ir Normung: DIN 18800 Part 4-Steel Structures, Stability, Buckling of
Shells, Beuth Verlag, Cologne, 1990.
10.28 Chryssanthopoulos, M.K., Giavotto, V. and Poggi, C., Statistical Imperfection Models for Buck-
ling Analysis of Composite Shells, in: Buckling of Shell Structures, on Land, in the Sea and in
the Air, Proceedings International Conference, J.F. Jullien, ed., TNSA Lyon, September 17-19,
Elsevier Applied Science, London, New York, 1991, 43-52.
10.29 Arbocz, J. and Babcock, C.D. Jr., Utilization of STAGS to Determine Knockdown Factors from
Measured Initial Imperfections, Report LR-275, Department of Aerospace Engineering, Delft
University of Technology, November 1978.
10.30 Tennyson, R.C., Muggeridge, D.B. and Caswell, R.D., Buckling of Circular Cylindrical Shells
Having Axisymmetric Imperfection Distributions, AIAA Journal, 9, (5), May 1971, 924-930.
10.31 Abramovich, H., Singer, J. and Jaffe, R., Imperfection Characteristics of Stiffened Shells-Group
I, TAE Report 406, Faculty of Aerospace Engineering, Technion, Israel Institute of Technology,
Haifa, Israel, 1981.
10.32 Yaffe, R., Singer, J. and Abramovich, H., Further Initial Imperfection Measurements oflntegrally
Stringer-Stiffened Cylindrical Shells-Series 2, TAE Report 404, Faculty of Aerospace Engi-
neering, Technion, Israel Institute of Technology, Haifa, Israel, 1981.
10.33 Fahy, W., Collapse of Longitudinally Stiffened Cylinders Subject to Axial and Pressure Loading,
Ph.D. thesis, Imperial College, University of London, 1984.
10.34 Agelides, N., Harding, J.E. and Dowling, P.J., Buckling Tests on Stringer-stiffened Cylinder
Models Subjected to Load Combination, Det norske Veritas Report 82-098, Imperial College,
1982.
10.35 Agelides, N., Collapse of Stringer-Stiffened Cylinders, Ph.D. thesis, Imperial College, University
of London, 1984.
10.36 Odland, J., On the Strength of Welded Ring Stiffened Cylindrical Shells Primarily Subjected to
Axial Compression, Report UR-81-15, Marine Technology Center, University of Trondheim,
June 1981.
10.37 Waeckel, N. and Jullien, J.F., Experimental Study on the Instability of Cylindrical Shells with
Initial Geometric Imperfections, Pressure Vessels and Piping Conference and Exhibit, June
17-21, 1984, San Antonio, Tex., in: ASME Special Publications PVP, 89, 1984, 69-77.
10.38 Giavotto, V., Poggi, C. and Chryssanthopoulos, M., Buckling of Imperfect Composite Shells
under Axial Compression and Torsion, in: Proceedings of Annual Forum of the American Hel-
icopter Society, Atlanta, Georgia, March 1991.
10.39 Arbocz, J., Collapse Load Calculations for Axially Compressed Imperfect Stringer Stiffened
Shells, in: Proceedings of AIAAIASMEIASCE/AHS 25th Structures, Structural Dynamics and
Materials Conference, May 14-18, 1984, Palm Springs, Calif., 130-139.
10.40 Elishakoff, I., Manen, S. van, Vermeulen, P.G. and Arbocz, J., First-Order Second-Moment Anal-
ysis of the Buckling of Shells with Random Imperfections, AIAA Journal, 25, (8), August 1987,
1113-1117.
10.41 Arbocz, J. and Ho1, J.M.A.M., Collapse of Axially Compressed Cylindrical Shells with Random
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Imperfections, AIAA Journal, 29, (12), December 1991, 2247-2256.


10.42 Chryssanthopoulos, M. and Poggi, C., Stochastic Imperfection Modelling in Shell Buckling Stud-
ies, Thin-Walled Structures, 23, (1-4), 1995, 179-200.
10.43 Imber!, J., The Effect of Imperfections on the Buckling of Cylindrical Shells, Aeronautical En-
gineer thesis, California Institute of Technology, Pasadena, Calif., 1971.
10.44 Arbocz, J. and Abramovich, H., The Initial Imperfection Data Bank at the Delft University of
Technology-Part I, Report LR-290, Delft University of Technology, Department of Aerospace
Engineering, 1979.
10.45 Dancy, R. and Jacobs, D., The Initial Imperfection Data Bank at the Delft University of Tech-
nology-Part II, Report LR-559, Delft University of Technology, Department of Aerospace En-
gineering, June 1988.
10.46 Klompe, A.W.H. and den Reyer, P.C., The Initial Imperfection Data Bank at the Delft University
of Technology-Part III, Report LR-568, Delft University of Technology, Department of Aero-
space Engineering, March 1989.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
860 /nitiallmperfections

10.47 Klompe, A.W.H. and den Reyer, P.C., The Initial Imperfection Data Bank at the Delft University
of Technology-Part IV, Report LR-569, Delft University of Technology, Department of Aero-
space Engineering, May 1989.
10.48 Arbocz, J., Towards an Improved Design Procedure for Buckling Critical Structures, in: Buckling
of Shell Structures, on Land, in the Sea and in the Air, Proceedings of International Conference,
J.F. Jullien, ed., INSA-Lyon, September 17-19, 1991, Elsevier Applied Science, London, New
York, 1991, 270-276.
10.49 Ang, A.H-S. and Cornell, C.A., Reliability Bases of Structural Safety and Design, Journal of
the Structural Divison, ASCE, 100, (ST9), Sept. 1974, 1755-1769.
10.50 Ravindra, M.K. and Galambos, T.V., Load and Resistance Factor Design for Steel, Journal of
the Structural Division, ASCE, 104, (ST9), Sept. 1978, 1337-1353.
10.51 Ryan, R.S. and Townsand, J.S., Application of Probabilistic Analysis/Design Methods in Space
Programs: The Approaches, the Status, and the Needs, AIAA Paper 93-1381, 34th AIAA/ ASME/
ASCE/ AHS I ASC Structures, Structural Dynamics and Materials Conference. April 19-22,
1993, La Jolla, Calif.
10.52 Bolotin, V.V., Statistical Methods in the Nonlinear Theory of Shells, NASA TT F-85, 1962.
(Translation of a paper presented at a seminar in the Institute of Mechanics of the Academy of
Sciences of the U.S.S.R., 1957.)
10.53 Fraser, W.B. and Budiansky, B., The Buckling of a Column with Random Initial Deflections,
Journal of Applied Mechanics, 36, 1969, 233-240.
10.54 Amazigo, J.C., Buckling under Axial Compression of Long Cylindrical Shells with Random
Axisymmetric Imperfections, Quarterly of Applied Mathematics, 26, 1969, 537-566.
10.55 Roorda, J. and Hansen, J.S., Random Buckling Behavior in Axially Loaded Cylindrical Shells
with Axisymmetric Imperfections, Journal of Spacecraft, 9, (I), 1972, 88-91.
10.56 Elishakoff, 1., Probabilistic Methods in the Theory of Structures, John Wiley & Sons, New York,
1983.
10.57 Dwight, H.B., Tables of Integrals and Other Mathematical Data, Macmillan, New York, 1969.
10.58 Elishakoff, I., Buckling of a Stochas1:ically Imperfect Finite Column on a Nonlinear Elastic
Foundation-A Reliability Study, Joumal of Applied Mechanics, 46, 1979, 5411-5416.
10.59 Elishakoff, I. and Arbocz, J., Reliability of Axially Compressed Cylindrical Shells with Random
Axisymmetric Imperfections, International Journal of Solids & Structures, 18, (7), 1982,
563-585.
10.60 Massey, F.J. Jr., The Kolmogorov-Smirnov Test for Goodness of Fit, Journal of the American
Statistical Association, 46, (253 ), 1951, 68-78.
10.61 Arbocz, J. and Babcock, C.D., Prediction of Buckling Loads Based on Experimentally Measured
Initial Imperfections, in: Buckling of Shells, Proceedings of IUTAM Symposium Buckling of
Structures, B. Budiansky, ed., Harvard University, Cambridge, MA, June 17-21, 1974, Springer
Verlag, Berlin, Heidelberg, New York, 1976, 291-311.
10.62 Elishakoff, I. and Arbocz, J., Reliability of Axially Compressed Cylindrical Shells with General
Nonsymmetric Imperfections, Journal ofApplied Mechanics, 52, March 1985, 122-128.
10.63 Schueller, G.!. and Bayer V., Computational Procedures in Structural Reliability, in: Proceedings
of ISUMA '93-The Second International Symposium on Uncertainty Modeling and Analysis,
April 25-28, 1993, University of Maryland, College Park, Md., IEEE, 552-559.
10.64 Fox, E.P., Methods of Integrating Probabilistic Design within an Organization's Design System
Using Box-Behnken Matrices, in: Proceedings of 34th AIAAIASME!ASCEIAHS!ASC Struc-
tures, Structural Dynamics and Materials Conference, April 19-22, 1993, La Jolla, Calif.. AIAA,
714-723.
10.65 Wohlfahrt, H., Residual Stresses Due to Welding: Their Origin, Calculation and Evaluation, in:
Residual Stresses, E. Macherauch, and V. Hauk, eds., DGM Informationsgesellschaft Verlag,
Oberursel, Germany, 1986.
10.66 Somerville, W.L., Swan, J.W. and Clarke, J.D., Measurement of Residual Stresses and Distortions
in Stiffened Plates, Journal of Strain Analysis, 12, (2), 1977, 107-116.
10.67 Pineault, J. and Brauss, M., Stress Mapping: A New Way of Tackling the Characterization of
Residual Stresses, Experimental Techniques, March/ April 1995, 17-19.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`

10.68 Beer, H. and Schulz, G., The Theoretical Bases of the New Column Curves of the European
Convention of Constructional Steelwork, Construction Metallique, (3), Sept. 1970 (in French).
10.69 Introductory Report of the Second Imernational Colloquium on Stability, Liege, April 13-15,
1977, European Convention for Constructional Steelwork (ECCS).
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 861

10.70 Cicin, P., Betrachtungen i.iber die Bruchursachen der neuer Wiener Donaubri.icke, Tiejhau, 12,
1970, 665-67 4.
I 0. 71 Shirley-Smith, H., Report on Collapse of Milford Haven Bridge-Fatal Accident 2nd June, 1970
(Expert's report).
10.72 McCormick, M.M. and Banks, E.E., Selected General Information-Project 52 West Gate Col-
lapse Investigation, Melbourne Research Laboratories, November 1970.
10.73 Tillman, S.C. and Williams, A.F., Buckling under Compression of Simple and Multicell Plate
Columns, Thin-Walled Structures, 8, 1989, 147-161.
10.74 Leonhardt, F. and Hommel, D., The Necessity of Quantifying Imperfections of All Structural
Members for Stability of Box Girders, in: Steel Box Girder Bridxes, P. Cartledge, ed., Institution
of Civil Engineers, London, 1973, 11-19.
10.75 Megson, T.H.G. and Hallak, G., Measurement of the Geometric Initial Imperfections in Dia-
phragms, Thin- Walled Structures, 14, 1992, 381-394.
10.76 Fukumoto, Y., Numerical Data Bank for the Ultimate Strength of Steel Structures, Der Stahlbau,
51, (!),January 1982, 21-27.
10.77 Aoki, T. and Fukumoto, Y., An Experimental Study of Local and Overall Buckling of Thin
Walled H-Columns, in: Stability of Plate and Shell Structures, Proceedings of an ECCS Inter-
national Colloquium, Ghent, Belgium, April 6-8, 1987, P. Dubas and D. Vandepitte, eds., Ghent

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
University, 1987, 129-134.
10.78 Sheer, J. and Fuchs, G., A Critical Evaluation of 705 Buckling Tests on Plates, in: Stability of
Plate and Shell Structures, Proceedings of an ECCS International Colloquium, Ghent, Belgium,
April 6-8 1987, P. Dubas and D. Vandepitte, eds., Ghent University, 1987, 119-128.
10.79 Galletly, G.D. (Guest Editor), Special Issue on Buckling-Strength of Imperfection Sensitive
Shells, Thin- Walled Structures, 23, (1-4), 1995.
10.80 Proceedinxs of 38th AIAA/ASMEIASCE/AHS/ASC Structures, Structural Dynamics and Ma-
terials Conference, Kissimmee, Florida, April 7-10, 1997, 1904-1963, 2104-2223, 2400-2445,
2446-2531, 2639-2712.

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
11
Boundary Conditions and
Loading Conditions

It was shown in Chapter 2 that the analytical formulation of the stability problems of con-
tinuous structures usually involves the solution of boundary value problems. Moreover, it
was indicated that the boundary value problem becomes an eigenvalue problem whenever
the governing differential equation and the given boundary conditions are homogeneous and
nontrivial solutions exist only for certain values of a parameter, say A,, the critical buckling
load. In this chapter the effect of different boundary and loading conditions on the critical
buckling load of a number of important structural elements will be considered.

11.1 Column Buckling


From the different examples considered in Chapter 2 it is evident that the elastic buckling
load of an axially compressed column can always be written as

(ll.l)

where the value of the end-fixidity coefficient c depends on the boundary conditions speci-
fied. The quantity L/Vc is called the effective length. To obtain the value of c for a variety
of different boundary conditions, one can usc to great advantage the characteristic equation
of an axially compressed bar with elastically restraint ends given by Eq. (2.24). It is relatively
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

easy to show (see for example [2.2]) that for

One end free, one end clamped c = I /4


Both ends simply supported c= l
Both ends clamped c = 4

In all these cases it is assumed that the axial force is applied without any eccentricity. To
take load eccentricity into account for the simply supported column shown in Figure 11.1,
one must solve the following problem:

E/w,"" + Pw,,.,. = 0 for 0 :<o:: x :<o:: L (11.2a)

w=O at x = O,L (11.2b)


M, = -Elw,,,, -Pe
Its solution is
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
Buckling Experiments: Experimental Methods in Buckling of Thin-Walled
No reproduction or networking permitted without license from IHS
Structures: Shells, Built-Up Structures, Composites
Not for Resale, 02/13/2019 01:32:58 MST
and Additional Topics – Volume 2. J. Singer, J. Arbocz and T. Weller Copyright © 2002 John Wiley & Sons, Inc.
864 Boundary Conditions and L.oading Conditions

P_j_
__·_J= i•X
Figure 11.1 Eccentrically loaded column

w = -.-e- {sin kL(l -- cos kx) - sin kx(l - cos kL)} (11.3)
Sill kL

where

k= If (11.4)

Calculating the deflection at x = L/2, one obtains, after some regrouping,

(11.5)

where

P. = El (11.6)
. 71"2
u
is the critical buckling load of the simply supported, centrally loaded column. As can be
seen in Figure 11.2, the effect of load eccentricity e is similar to the effect of initial imper-
fections shown in Figure 2.2. Notice that the deflections of a column with small load ec-
centricity grow very fast close to and at P = P,. Due to these rapidly increasing bending
deformations, the stresses on the concave side soon exceed the yield stress and in practical
applications collapse of slender columns occurs at P slightly below but close to P,.
The effect of load eccentricity is covered extensively in Chapter 4. It has been shown by
von Karman [4.4] that by using adjustable end fixtures one can minimize the degrading
effect of load eccentricity (see Figure 4.1 ). Experimental verification of the effect of rigid
end supports and the restraining of end rotation is discussed in Chapter 6. It has been shown

Effect of Load Eccentricity on Colun1n Buckling


..... Displacerr1ent at x=L/2 ~~axial load
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

~ e-0.001

-o-G-e e-0.01

v
0
0

,,00
Displacement ot x=L/2

Figure 11.2 Nonlinear behavior of eccentrically loaded columns


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plate Buckling 865

that to reproduce experimentally simply supported or fully clamped boundary conditions is


not simple at all. It is further concluded that additional experiments about the buckling and
postbuckling behavior of elastically connected frames are still needed.
Elaborate design recommendations for axially compressed columns, taking into account
the effects of load eccentricity, initial imperfections, residual stresses and the different bound-
ary constraints can be found in the publications of the European Convention for Construc-
tional Steelwork (ECCS) such as the Guide to Design Criteria for Metal Compression Mem-
bers [ 11.1].

11.2 Plate Buckling


The buckling behavior of plates depends not only on the type of applied loading but also
on the edge supports used. To illustrate these phenomena, let us consider the axially com-
pressed long rectangular plate shown in Figure 11.3. By using different boundary conditions
along the unloaded edges, one obtains three completely different buckling patterns. Both
unloaded edges free yields the single wave buckling mode shown in Figure 11.3a, known
as the column behavior. One unloaded edge simply supported and the other free produces
the twisted wave buckling mode displayed in Figure 11.3b, known as the flange behavior.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Finally, with both unloaded edges simply supported, one obtains the multiple-buckle pattern
depicted in Figure 11.3c.
In the following the stability problem of an axially compressed flat plate will be solved
in order to illustrate the way in which different boundary conditions influence the buckling
behavior. The linearized stability equation of the axially compressed long rectangular plate
shown in Figure 11.3 is
U1;74w + N0 w,, = 0 (3.28a)
Assuming that the loaded edges are simply supported then
. X
w = w(y) sm mn-- (11.7)
a
satisfies these boundary conditions identically. Substitution into Eq. (3.28a) and regrouping
yields

Figure 11.3 Transition from column to plate buckling as supports are added along the unloaded edges
(from [2.13])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
866 Boundary Conditions and l.oading Conditions

( )2w,,_, + [(---;- )4 -
N0
_ m1T _
w,_,y_,T - 2 ---;-
m1T
D (---;-
m1T
)2] w_ = 0 (11.8)

a constant coefficient ordinary different:lal equation, which always admits an exponential


solution. The corresponding characteristic equation is

(11.9)

The roots of Eq. (11.9) are

(11.10)

It was shown in [2.2] that the corresponding solution can be written as

w
(i
= ( C 1 cos h by + C 2 sm
. h(i
by + C, [3 + C
cosby .
sm
[3).sm m1r;X
by (11.11)
4

where

(11.12a)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(ll.l2b)

Here €, = aim is the axial buckle half wave length and k, is the plate buckling coefficient
from Eq. (2.7). The coefficients C 1, C2 , C,, C4 are to be determined from the specified
boundary conditions along the unloaded edges.

11.2. 1 Column Behavior


To obtain the column-like behavior, both unloaded edges must be free. In order to take
advantage of the symmetry of the buckling pattern in the y-direction, it is advantageous to
place the origin of the coordinate axes halfway along the loaded edge. Thus, at y = ± b/2
the bending moment
(11.13)
and the reduced shear
M, _ , + (M,.,. + M,,J., = -D[w,,.,.,. + (2 - v)w,,".] (11.14)
must both vanish. Introducing the expression for w from Eq. (11.11) into these homogeneous
boundary conditions and taking advantage of the symmetry of the buckling mode with re-
spect to the x-axis yields a determinantal equation which can be put into the following form
(see [2.13]):

--
{3p 2 tan l~· - 2i
+ iiq 2 tanh = 0 (11.15)

where

v(;~r=ii2 - v (m:br (11.16a)

q = [32 + v C;'~Y = [32 + v (m:br (11.16b)

Equation (11.15) forms an explicit expression for the nondimensional buckling load coeffi-
cient k, in terms of the wavelength parameter m, the Poisson's ratio v and the aspect ratio
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plate Buckling 867

alb. Calculations show that for all values of alb the minimum buckling load occurs for
m = 1. The buckling coefficient k, is shown as a function of v and bl a in Figure 11.4. To
understand the results of this figure properly, one must first rewrite the standard buckling
formula for plates as follows:

a, = k, 12(;~
"E
v 2)
(h)b 2

= k,
(
~
)
2
1r2£
12(1 - v 2 )
(h)~ 2
_ 1r2E
= k, 1 - v 2
I
a 2A (2.7)

where I = bh 3 !12 and A = bh. Next one assumes that I = Ap2 , where pis the radius of
gyration, yielding the formula used to plot the results of Figure 11.4:
k 7T2£
a, = I _' v2 (LI r? (11.17)

where L = a. Notice that in the range where bl a = b/ L > > 1, k, approaches 1 and Eq.
(11.17) reduces for p = hlvTi to Eq. (2.10), the wide column formula. In the other limit
when b/a = b/L << 1, k, approaches I - v 2 , and Eq. (11.17) reduces to Eq. (2.5), the
Euler buckling formula for simply supported columns. These solutions for wide columns
were first published by Houbolt and Stowell [ 11.2].

11.2.2 Flange Behavior


To obtain the flange-like behavior, and using the coordinate axes as shown in Figure 11.3b,
the unloaded edge at y = 0 is assumed to be simply supported and that at y = b to be free.
Thus, at y = 0
w=O (11.18a)
M, = -D(w," + vw,,J = 0 (11.18b)
whereas at y = b
M,. = -D(w,n + vw,") = 0 (11.19a)

M,.,,. + (M,,. + M,,J,, = -D[w,,n + (2- v)w,x,J = 0 (11.19b)


Introducing the expression for w from Eq. (11.11) into these homogeneous boundary con- --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

ditions yields a determinantal equation that can be put into the following form (see [2.2]):
(11.20)
where ii, /3 are defined by Eqs. ( 11.12) and p, q by Eqs. (11.16), respectively. Equation
(11.20) is the characteristic equation for determining the nondimensional buckling load co-
efficient k, in terms of the wavelength parameter m, the Poisson's ratio v and the plate aspect

1.00
II II /
F="'
~~~.k5 / v
.96 -
~9
.9 2
1-v2

,.= _v
- Exoct solution
Energy solution
.8 8 v 0

.2 .4 2
II4 IIIII10 2040
II 100
b
a
Figure 11.4 Compressive buckling coefficients ( of plate columns [see Eq. (11.17)] as functions of v
and hi a (from [11.2])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
868 Boundary Conditions and Loading Conditions

ratio a/ b. Calculations show that for all values of a! b the minimum load occurs form = 1.
The buckling coefficient k, is shown as a function of v and a/ b in Figure 11.5. Notice that
for long plates (a/ b > 3, say) k, approaches a value between 0.395 and 0.456, the exact
value depending on the Poisson's ratio v. These solutions were published initially by Lund-
quist and Stowell [11.3] in 1942.

11.2.3 Plate Behavior


Finally, to obtain the multiple-buckle pattern typical of plate buckling, the unloaded edges
must not be free. For simply supported edges one must satisfy the following boundary
conditions at y = O,b
w=O (11.21a)
M,. = -D(w •.,.r + vw,,,J = 0 (11.2lb)
Notice that since w = 0 along the edges y = 0 andy = b, therefore also w," = 0 along the
same edges. Thus, Eq. (11.2lb) reduces to
(11.21c)
Introducing the expression for w from Eq. (11.11) into these homogeneous boundary con-
ditions yields the following determinantal equation:

[(~-)2 + (il)2]
[;; sinh a sin - = 0 {3 (11.22)

For a =F 0 and 13 =F 0 this implies


>in 13 =0 (11.23)
yielding the following eigenvalue:

13 = 7T {f.
{€, 'Y
j_ !!_e, + Vk., = n7T n = 1, 2, ... (11.24)

For the lowest eigenvalue, n = 1 and one obtains the following expression for the buckling
coefficient

k, = ( ~ +;
Notice that this equation is identical to Eq. (2.8) with n = 1. To obtain its minimum value
r :b :br
= ( + (11.25)

one plots k, as a function of a/ b for different values of m as shown in Figure 2.4. The

u,---,----~---,---,---,

a/b
--`,`,`````,`,````,,``,``,,`-

Figure 11.5 Compressive buckling coefficients k, of plate flanges as functions of v and a I h (from
[2.13])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plate Buckling 869

minimum value of k, is given by the lower envelope of the curves, indicated in Figure 2.4
by a solid line. Notice that for a long plate, say a/ b > 3, k, = 4.0. Figure 2.5 presents the
values of k, as a function of the aspect ratio a/ b for various combinations of simply sup-
ported, clamped and free edges. The results shown in Figure 2.5 were first published by Hill
[11.4].

11.2.4 Elastically Supported Unloaded Edges


In many practical applications the boundary conditions lie somewhere between simple sup-
ports and fully clamped. To obtain a solution for the critical compressive stress of flat
rectangular plates supported along all edges and elastically restrained against rotation along
the unloaded edges, it is advantageous to place the origin of the coordinate axes halfway
along the loaded edge. Considering Figure 11.6 and remembering that the elastic restoring
moment always acts opposite to the applied rotation, then at y = -b/2
(11.26)
and at y = b/2
(11.27)
where a 1, a 2 are the torsional stiffncsses per unit length of the restraining medium at the
unloaded edges. Recalling from the constitutive equation that for isotropic plates
M, = -D(w,"" + vw,,J (11.28)
where if w = 0 at y = ±b/2, then also w," = 0 along the same edges. Thus, one must
enforce the following boundary conditions along the unloaded edges, at y = -b/2
w=O (11.29a)
£ 111/,y- W, 11 =0 (11.29b)
whereas at y = b/2
w=O (11.30a)
E 2tV, 1. + tV,'"" = 0 (11.30b)
and where 10 1 = a 1 /D, 10 2 = aJD and D = Eh 1 /12(1 - v 2 ). Assuming identical rotational
restraints 10 = 10 1 = 10 2 and introducing the expression for w from Eq. (11.11) into the above
homogeneous boundary conditions yields, after some regrouping, the following determinantal
equation:

{ (Ci + [3 2 ) + 10 ( ii tanh~ + [3 tan~)} { (ii 2


+ [3 2
) + 10 (a coth ~ - [3 cot~)} = 0
(11.31)
where the values for a and [3 are given by Eqs. (11.12). Equation (11.31) is the characteristic
equation for determining the nondimensional buckling load coefficient k, in terms of the
wavelength parameter m, the Poisson's ratio v and the plate aspect ratio alb. By setting the
first factor in Eq. (11.31) equal to zero,

(a 2 + f3- 2 ) + 10
( a tanh 2
ii + {3- tan 2[3) = 0 (11.32)

z.w

w.y(-b/2) -w.y(b/2)

eM,
Copyright Wiley
y
Figure 11.6
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

End moments along the unloaded edges of a plate


Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
870 Boundary Conditions and Loading Conditions

one obtains the symmetric buckling modes. This equation was first published by Dunn in
1940 [11.5]. Similarly, the anti-symmetric buckling modes are obtained by setting the second
factor in Eq. (11.31) equal to zero:

(a + {3 + e
2 2
) (a coth ~- j3 cot ~) = 0 (11.33)

Calculations indicate that of these two possible buckling modes the symmetrical one given
by Eq. (11.32) will occur at the lower cntical stress. Thus, in [11.6] Lundquist and Stowell
used Eq. ( 11.32) to calculate the values of the critical compressive buckling coefficient k,
as a function of plate aspect ratio a I b for various amounts of edge rotational restraint. Their
results are displayed in Figure 11.7.
Before one can use Figure 11.7 to determine k,, it is necessary to evaluate the torsional
restraint coefficient e. In deriving the results presented, Lundquist and Stowell have assumed
that rotation of the elastic restraining medium at one point does not affect the rotation at
another point of the medium. The practical case where the elastic torsional restraint is pro-
vided by a stiffener, for which rotation at one point affects rotation at another point, was
treated by the same authors in [11.7]. They pointed out that when the stiffener rotational
rigidity has been found, e can be computed by forming the ratio of this rigidity to the
rotational rigidity of the plate. Assuming that the torsional restraint of a flat plate is provided
by a stiffener that does not suffer cross-sectional distortion when moments are applied to
some part of the cross-section, they derived the following formula for the torsional restraint

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
coefficient:

e = -
n2b ( GJ - a1 + -.
7r2
Ef ) (11.34)
e;D o e,
~r--r---------------,

14-

E
H\\'--'+1'~~-- 1-c ---+-~-''1---+~+----+-=--1 "'

a
b
Figure 11.7 Compressive buckling coefficients k, of plates as functions of a/ b for various amounts of
edge rotational restraint (from [ 11.6])

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plate Buckling 871

where b = width of plate being restrained


E, = axial buckle half wave length(= aim)
D = flexural rigidity per unit length of plate = [Eh 3 1l2(l - v 2 )]
GJ = torsional rigidity of stiffener
o- = uniformly distributed compressive stress in stiffener
/ 0 = polar moment of inertia of stiffener sectional area about axis of rotation

r = Wagner torsion-bending constant of stiffener sectional area about axis of rota-


tion at or near edge of plate

In conclusion, according to Gerard and Becker [11.8], it is not necessary to determine this
stiffness to a high degree of accuracy, since in many cases the influence of e on k, embraces
a large range of stiffness ratios, as can be seen in Figure 11.8 for infinitely long plates.

11.2.5 Effect of Boundary Conditions for Shear Loading


The effect of different in-plane boundary conditions on the buckling and postbuckling be-
havior of long orthotropic plates loaded in combined shear and compression was investigated
by Stein in [11.9]. He converted the governing von Karman-type nonlinear, large deflection
partial differential equations [ 11.1 0] into a set of first -order nonlinear, ordinary differential
equations by assuming the following truncated Fourier series for the displacements of the
problem:

u = -u,n
- (X
~ - 21) + U 0 (y) + u,(y) sin T
27TX
+ uJy) cosT
27TX
(ll.35a)

2m: 2-rrx
v = u0 (y) + v,(y) sin T + v,(y) cos T (11.35b)

. 7TX 7TX
W = w,(y) Sill f + wJy) COS f (11.35c)
X .\

where f, is the half-wavelength in the x-direction. Application of the principle of virtual

FLANGE ~

Df
I~

Ll

10

.1, .8 ~--------+-----_fl~Lr~-------~--------~6 .1,


FLANGE PlATE

.6

" *c•o.400
AT €"0

.2

0
.I LO

Figure 11.8 Compressive-buckling coefficients for infinitely long flanges and plates as functions of
edge rotational restraint (from [11.8]) --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
872 Boundary Conditions and L.oading Conditions

work yields, not only the set of governing nonlinear, first-order ordinary differential equations
but also the admissible set of boundary conditions.
Stein chose to investigate the cases where the edges are held straight and are either simply
supported or clamped. To apply the specified shearing displacement (see Figure 11.9), u,h
the edge at y = 0 is displaced relative to the edge at y = h by a rigid frame pinned at the
corners of the plate. Here one must remember that by this kind of frame one is also applying
a compressive displacement across the width equal to
, -, I/'
h-(lr--u-) 1 I -,
-=--u-
,h 2 h _,h

Notice that instead of the rigid frame condition, one could also assume that the edges are
constrained such that they do not move together or apart.
These boundary conditions can be expressed as follows:
1. Simply supported or clamped at y = O,h
w, = w, = 0 and M,, = M,, = 0 or {3, = {3, = 0 (11.36)
2. Straight edges at y = O.h
(11.37)
3. Applied shearing displacement

at y = 0 (11.38)

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
at y = h

4. Transverse compressive displacements due to a rigid frame:

at y = 0 at y = h Vo = (11.39a)

or, edges are constrained so that they do not move together or apart:
at y = O,h u0 = 0 (11.39b)
The set of first-order nonlinear, ordinary differential equations subjected to the above-
listed boundary conditions can be put into the form of the following two-point boundary
value problem
d
- Y = F(y, Y) (11.40)
dy
g(Y(O), Y(h)) = 0 (11.41)
where Y is the 20-dimensional vector of dependent variables andy is the independent vari-
able defined in the interval, (O,h). The boundary conditions of the problem are specified by
the vector function g. Stein used an algorithm based on Newton's method developed by
Lentini and Pereyra [ 11.11 J to solve this problem. The algorithm uses finite differences with

Figure 11.9 Buckled plate under combined shear and compression


(from M. Stein, "Analytical Results for Post-buckling
Behaviour of Plates in Compression and Shear," in:
Aspects of the Analysis of Plate Structures, D.J. Dawe,
R.W. Horsington, A.G. Kamrekar, and G.H. Little,
eds., copyright 1985, by permission of Oxford Uni-
versity Press)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plate Buckling 873

deferred corrections. Adaptive mesh spaces are automatically generated so that mild bound-
ary layers are detected and resolved. When the solution is carried out, the wavelength f, of
interest is the one that corresponds to minimum energy, and the solution of interest is on
the equilibrium path that gives non-zero deflections. Results were obtained for isotropic and
±45°-laminated composite plates with a balanced and symmetric layup.
Stein presented his results [11.9] and [11.10] as characteristic load-displacement curves,
where the long edges are either simply supported or clamped. In addition, the long edges
are held straight with either the average transverse stress intensity N'"' equal to zero or the
displacement normal to the long edges v equal to zero, where
1 ('' (21,
N,a, = 2 £.b Jo Jo N, dx dy (11.42)

As an alternative, Stein investigated also the case where the long edges displace toward each
other as required by rigid frame loading device pinned at the corners.
In Figure 11.10 for isotropic plates the average shear stress intensity coefficient is plotted
as a function of the applied shear displacement coefficient. Depending on the in-plane bound-
ary conditions the curves branch at the values of the coefficient that correspond to buckling
loads for simply supported and clamped plates. The slopes of these curves indicate that in
the postbuckling range plates with clamped edges are stiffer than plates with simply sup-
ported edges. Also plates with zero displacement edge condition (v = 0) are stiffer than the
ones supported by a rigid frame. Notice further that these plates arc much stiffer than plates
with average transverse stress N,a, equal to zero.
Curves for the postbuckling behavior of long, ±45°-laminated composite plates loaded in
shear are plotted in Figure 11.11. The slopes of the curves presented show the same trends
as the ones displayed for the isotropic plates in Figure 11.10, except that they indicate that
the stiffness of the ± 45o-laminate is less than the isotropic plates for all cases.
The ±45°-laminate results apply to graphite-epoxy filamentary material with properties
given by the dimensionless quantities

(11.43a)

AIIA22 ~ Af2 -- 2AI066


~0.431 (11.43b)
2A66 \1A I IA22

(11.43c)

Stein introduced these dimensionless quantities in [11.12]. For additional results the inter-
ested reader should consult [11.9] and [11.10].

11.2.6 Experimental Verification


As shown earlier (sec Figure 2.5 and [2.13]), the buckling load of a rectangular plate depends
strongly on its out-of-plane boundary conditions, especially on those along the unloaded
edges. When the plate continues to deform in the postbuckling regime, membrane stresses
come into action, whereby the postbuckling stress distributions depend strongly on the in-
plane boundary conditions used (see Figure 2.6 and [2.14]). Thus, as stated in Section 8.1.3,
the precise definition and control of both out-of-plane and in-plane boundary conditions has
been of primary concern in all plate buckling and postbuckling experiments.
--`,`,`````,`,````,,``,``,,`-`-`,,`,

As it turned out, to reproduce simply supported edge conditions especially along the
unloaded edges, and to achieve a uniform load distribution until buckling is initiated, in a
plate buckling experiment is by no means a trivial task.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
874 Boundary Conditions and L.oading Conditions

30 In-plane
condition

;:;,h/2

8~ ~~1
jjsh/2

--Clamped
- Simply supported

20 40 60
Applied shear displacement coeffic1ent
Ush Ass bz

Jou D~2 nl

Figure 11.10 Normalized load-deflection curves for isotropic plates loaded in shear (from M. Stein,
"Analytical Results for Post-buckling Behaviour of Plates in Compression and Shear,"
in: Aspects of the Analysis of Plate Structures, D.J. Dawe, R.W. Horsington, A. G. Kam-
rekar, and G.H. Little, eds., copyright 1985, by permission of Oxford University Press)

Considering first axially compressed plates, L. Schuman and G. Back [8.6] in 1930 tried
to model simply supported unloaded edges by adjustable bars with 4SO V-grooves. As can
be seen in Figure 8.15, simple support was not precisely achieved and the edge support
perpendicular to the plate was actually weaker than presumed.
C.F. Kollbrunner [8.16] in 1946 used accurately machined circular cylindrical steel riders
that were fixed with set screws to the unloaded plate edges and could rotate about the
effective edge of the plate between two rigid steel guide beams to simulate simply supported
unloaded edges (see Figures 8.22-8.26).
N.J. Hoff, B.A. Boley and J.M. Coan [8.53] in 1948 employed split-needle bearings to
simulate simple support along the loaded edges (see Figure 8.45).

30
In-planEt
conditic1n

j " __ ""____________ _
t._--------
I.~·--
/.
_:C---
-- N
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

--Clamped
- Simply supported

20 40 60
Applied shear displacement coefficient
u," Ass b
··~2
.~o, o22"

Figure 11.11 Normalized load-deflection curves for a ±45"-laminated composite plate loaded in shear
(from M. Stein, "Analytical Results for Post-buckling Behaviour of Plates in Compres-
sion and Shear," in: Aspects of the Analysis of Plate Structures, D.J. Dawe, R.W. Hor-
sington, A.G. Kamrekar, and G.H. Little, eds., copyright 1985, by permission of Oxford
Copyright Wiley
University
Provided by IHS Markit under license with WILEY Press) Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Buckling of Circular Cylindrical Shells 875

Walker [8.41] in 1967 used rounded knife edges to simulate simple supports and restrain-
ing bars to simulate clamped support along the unloaded edges (see Figure 8.16). However,
friction caused unwanted restrictions to in-plane movements.
Schlack [8.54] in 1968 employed a series of independently acting needle bearing blocks
(see Figure 8.19) for all four edges of his square plate.
Finally, Cambridge University investigators [8.43]-[8.45] used a system of discrete "fin-
ger" supports, stiff against out-of-plane displacement of the plate while flexible to in-plane
displacements, both in the direction of the load and perpendicular to it (see Figures 8.30-
8.32).
In general, as can be seen in Figure 8.6, the agreement between the theoretically /numer-
ically predicted and the experimentally measured load-deflection curves can be called good
if initial imperfections are accounted for.
We turn now to rectangular plates subjected to shear stresses along their edges. Their
buckling and postbuckling deflection shapes are composed of a combination of several wave
forms with inclined nodal lines (see Figure 8.75). This makes the analysis of the buckling
and postbuckling behavior much more difficult and laborious. However, investigators at Tech-
nion [8.125] have demonstrated that by using the shadow-moire technique to observe the
progressive behavior of the buckled shear web and the advanced nonlinear computer code
STAGS [8.13] to calculate the postbuckling deflection pattern, it is possible to obtain very
good agreement between experimental results (see Figure 8.89a) and the theoretical/numer-
ical results (see Figure 8.89b).

11.2. 7 Effect of Boundary Conditions on Postbuckling Behavior


As pointed out in Subsection 8.1.3 of Chapter 8, when a plate continues to deform in the
postbuckling region, membrane stresses come into action to resist deformation, in addition
to the bending stresses that resisted initial buckling. The in-plane boundary conditions there-
fore join the out-of-plane ones in strongly influencing the postbuckling behavior. Significant
changes in the boundary conditions may also appear with increase in postbuckling defor-
mation, partly due to changes in the buckled shape (secondary buckling). The definition and
control of the in-plane boundary conditions, in addition to the out-of-plane ones, is therefore
of primary importance in all plate buckling experiments that continue into the deep post-
buckling region.
Indeed, in many of the plate buckling experiments discussed in Chapter 8 the in-plane
boundary conditions were carefully considered (see for example [8.18], [8.26], [8.34]-[8.36],
[8.43]-[8.45], [8.57] and [8.61]). However, additional precise experiments, focusing on the
influence of in-plane boundary conditions on the postbuckling behavior of plates (preferably
also correlated with numerical studies), are needed to provide quantitative data on this effect.
In future buckling tests on plates, it would therefore be advisable also to include mea-
surements of in-plane deformations and stresses near the boundaries.

11.3 Buckling of Circular Cylindrical Shells


Returning to Figure 3.1 and the large scatter of experimental results for axially compressed
cylindrical shells, it has been pointed out in Chapter 3 that besides initial imperfections, the
different types of boundary conditions do have a significant influence on the value of the
critical buckling load.
The effect of boundary conditions on the critical buckling load of axially compressed
cylindrical shells has been investigated extensively in the early sixties (see Refs. [11.14]-
[ 11.16]). These investigations have shown clearly that there are fundamental differences
between the stability problems of fiat plates under edge loading and the stability problems
involving shells of revolution. Besides the effect of curvature, which for shells of revolution
couples the out-of-plane equilibrium and the compatibility equations, there is also the effect
of prebuckling rotations.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
For a fiat plate under edge loading, the out-of-plane displacement w 0 and its derivatives
Copyright Wiley
Providedare
by IHSzero andlicense
Markit under thewithlinearized
WILEY stability equation has the particularly
Licensee=McDermott simple
Inc - India/8215328006, form
User=G, given by
Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
876 Boundary Conditions and Loading Conditions

Eq. (3.28). For a circular cylindrical shell under axial compression, the prebuckling defor-
mation caused by the edge constraints is not rotation free and thus the coupled linearized
stability equations (Eqs. (2.100a) and (2.100b)) contain the prebuckling rotation terms
lVo.rx' Wo,xv' and Wo.n·
Notice that by assuming a membrane prebuckling state (w 0 = constant, N,0 =
- A(Eh 2 I cR), Nxm = N"11 = 0) the variable coefficients linearized stability equations reduce
to Eqs. (2.l03a) and (2.103b), a set of constant coefficients equations, which for simply
supported boundary conditions can be solved easily by separation of variables, yielding the
eigenvalues and eigenfunctions given by Eqs. (2.105)-(2.107).
A rigorous solution of the equations governing the axisymmetric prebuckling state can be
found in Section 2.l.13a. It is shown that for simply supported boundary conditions axisym-
metric collapse will occur when the normalized axial load parameter A = N,(cRI Eh 2 ) = I.
Further, it is pointed out that bifurcation into an asymmetric mode with an integer number
of waves in the circumferential direction may occur before the axisymmetric collapse load
is reached.

11.3. 1 Effect of Boundary CondiUons Using Membrane Prebuckling


In trying to explain the large discrepancy between the experimental results shown in Figure
3.1 and the theoretical predictions based on the use of membrane prebuckling and the lin-
earized stability equations, Hoff and his co-workers published many pioneering papers in
the sixties (see [2.69], [11.17], and [11.18]). Introducing the nondimensional quantities

x = x* (2£; ¢- y* /2E. A= a«J; ad Eh c = Y3(1 - v 2)


R y-;;:; - R y-;;:;' a<~ cR' (11.44)

u = u* {2E. v = v* {2E. w = w*
y-;;:;' y-;;::'
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

R R R
where the starred quantities are the ones with physical dimensions, they succeeded in writing
the uncoupled linearized stability equations of Donnell ([2.2], p. 156) in the following par-
ticularly simple form:
(ll.45a)
(l1.45b)

(ll.45c)
Here subscripts following a comma indicate partial differentiation with respect to the inde-
pendent variables following the comma and the radial displacement w is defined positive
inward. The constant coefficient equation (ll.45c) admits an exponential solution of the
form
w == e"'' sin n¢ (11.46)
where p, is a root of the characteristic equation and n is the reduced circumferential wave
number defined as
- ~
n=n{2£ (11.47)

Notice that because of the requirement of continuity of displacements in the circumferential


direction, the number of waves n around the circumference of the shell must be an integer.
The general solution of Eq. (ll.45c) can then be written as (see [2.69], p. 500)

(11.48)

Copyright Wiley
where the constants A,, B,, i
Provided by IHS Markit under license with WILEY = I, 2, 3, 4 must be determinedIncfrom
Licensee=McDermott the fourUser=G,
- India/8215328006, specified
Boopathi boundary
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Buckling of Circular Cylindrical Shells 877

conditions at each of the two shell edges. It has been shown in [ 11.18] that one can impose
any one of the following four simply supported boundary conditions:
SSl: N, = N". = w = w," = 0 (11.49a)
SS2: u = N" = 1v = w,,. = 0 (11.49b)
SS3: N, = v = w = w,,, = 0 (11.49c)
SS4: u = v = w = W," = 0 (11.49d)
or any one of the following four clamped boundary conditions:
Cl: N, = N,. = w = w,, = 0 (11.50a)
C2: u = N,,. = w = w,, = 0 (11.50b)
C3: N, = v = w = w,, = 0 (11.50c)
C4: u = v = w = w,, = 0 (11.50d)
One may remember these eight boundary conditions from Table 2.1 of Chapter 2.
In addition, the end of the shell can be free of all stress and moment resultants except
those caused by the applied axial stress a, 0 • It was shown in [2.69] that if one assumes that
the direction and magnitude of a,0 does not change when the shell deforms at buckling, then
these conditions imply
N, = N" = 0; w,,_,. + (2 - v)w,,dxt> + 2Aw,, = 0; w," + vw,dxb = 0 (11.51)
Since these boundary conditions are homogeneous, nontrivial solutions of the resulting
8th order system of equations exist only for discrete values of the applied axial load A. These
eigenvalues are solutions of the characteristic equation obtained by setting the determinant
of the A,,B, terms equal to zero. The critical eigenvalue is the lowest one. It is determined
by an n-search, whereby, as mentioned above, n must be an integer.
Further, it was shown in [2.69] that if the same boundary conditions are applied at both
ends of the shell, say at x* = ± L/2, then one can specify symmetric or antisymmetric
buckling with respect to x* = 0, the mid-plane of the shell. This reduces the characteristic
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

equation to a 4th order system.


The critical eigenvalues and the con·csponding eigenvectors (the buckling modes) have
been evaluated in [2.69] for a wide variety of cases. The results were summarized quite
clearly in figures and tables, which shall be presented in the following with short comments.
For the SS 1 boundary conditions, as can be seen in Figure 11.12, the critical buckling
load is always close to A, = 0.5. Whether the critical buckling mode is symmetric or anti-
symmetric depends on the value of the reduced length parameter LlvRh. The shapes of a
few symmetric and antisymmetric buckling modes are plotted in Figure 11.13. Notice that
these boundary conditions always trigger an edge buckling-type behavior with relatively few
waves in the circumferential direction. The values of the parameters used to plot these curves
are listed in Table 11.1.
As can be seen in Figure 11.14, for the SS2 boundary conditions the critical buckling
load is also always close to A, = 0.5. The corresponding critical buckling modes are most
likely to be symmetric unless the reduced length parameter LlvRh > 400. The shapes of
the few symmetric and antisymmetric buckling modes plotted in Figure 11.15 are, at least
approximately, sinusoidal with relatively few waves in the circumferential direction. The
values of the parameters used to plot these curves are listed in Table 11.2.
Solutions for the SS3 boundary conditions are discussed in Chapter 2. Notice that the
assumed buckling mode shapes given by Eq. (2.104) satisfy the conditions specified by Eq.
(11.49c) identically. The corresponding eigenvalue A,,, is given by Eq. (2.105) and depends
not only on geometric parameters but also on the axial and circumferential wave numbers
Copyright Wiley
Provided m and
by IHS Markitn. The
under smallest
license with WILEY eigenvalue \ = 1 is a repeated eigenvalue,
Licensee=McDermott whereUser=G,
Inc - India/8215328006, the Boopathi
locus of the
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
878 Boundary Conditions and Loading Conditions

Lt/Rh
Figure 11.12 Critical buckling load -\ plotted against reduced length LIVRii (from [2.69]) (SSl
boundary conditions: N, = N" = w = M, = 0)

family of eigenvectors belonging to this eigenvalue constitutes the well-known Koiter circle
given by Eq. (2.109). For further details the reader should consult Section 2.1.11a.
Investigating the critical buckling loads for the SS4 boundary conditions, it was shown in
[2.69], that for a prescribed value of the reduced length parameter LtVRh the buckling load
A is a function of the circumferential wave number n. Further, it was found that there always
n
exists a value of the reduced wave number such as to make A, = 1.0, unless the shell is
very short. As can be seen from Figure I 1.16, whether the critical buckling mode is sym-
metric or antisymmetric depends on the specified value of the reduced length parameter. A
few symmetric and antisymmetric mode shapes are shown in Figure 11.17. The correspond-
ing values of the parameters used are listed in Table 11.3.
The results of the critical buckling load calculations for the four possible clamped bound-
ary conditions are summarized in Figure 11.18, whereby both symmetric and antisymmetric
buckling are considered. For all combinations the critical buckling load A, = 1, unless the
shell is very short. In the latter case the critical buckling load A, rises very rapidly with
decreasing length. A few selected symmetric and antisymmetric buckling modes for very
short shells are shown in Figure 11.19. The values of the parameters used are listed in Table
11.4.

I
~
SYiotlotETIIIC

0 ~'---'--'--'~~_L_j__....c.• ~-'l_j'L_ j' - -'- ' ~. -l.~_L_L


0 ~ M
--'-'
M M ~
__

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`--- x*/(L/2)

Figure 11.13
Copyright Wiley
Buckling modes for SSJ boundary conditions (from [2.69]) (the values of the constants
for the curves are given in Table 11.1)
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Buckling of Circular Cylindrical Shells 879

Table 11.1 Values of constants for Figure 11.13 (from [2.69])


LlvRh n= n Vaci/2E A,

A 15.8 0.08 0.539


B 31.6 0.08 0.521
c 95.0 0.08 0.509
D 31.6 0.1 0.505
E 47.4 0.1 0.505
F 15.8 0.085 0.511
G 95.0 0.085 0.504

Finally, let us consider the free edges case, where it is assumed that the shell edges are
perfectly free of all tractions except those induced by a uniformly distributed axial load,
whose direction remains constant in space while the shell deforms during buckling. The
results of the critical buckling load calculations carried out in [2.69] are shown in Figure
11.20 for symmetrical and in Figure 11.21 for antisymmetrical buckling. Notice that when
the shell buckles symmetrically, the critical buckling load is ,\ = 0.38 for long shells and
the number of circumferential waves n is relatively high. Further, as can be seen in Figure
11.20, for shorter shells both the critical buckling load A, and the reduced wave number n
(and hence the number of circumferential waves n) increase slightly.
For antisymmetric buckling the critical buckling load of very long shells is also A, = 0.38.
However, as can be seen in Figure 11.21, shorter shells buckle at very low values of \.
Further, whereas very long shells buckle in several waves in the circumferential direction,
shorter shells buckle in the n = 2 mode. A few selected symmetric and antisymmetric buckle
modes are shown in Figure 11.22. The values of the parameters used are listed in Table 11.5.
If one looks at the above results critically, it appears that when one applies the SS1 or
SS2 boundary conditions the critical buckling load A, = 0.5 corresponds to symmetric buck-
ling with only two waves in the circumferential direction. Similarly, in the case of free edges,
shorter shells buckle antisymmetrically at very low values of A,( <0.1) with n = 2 full waves
around the circumference. Since Donnell's equations are known to be inaccurate when the
number of full waves in the circumferential direction is small (see for example [2.2], p. 139),
Hoff and Soong [2.69] repeated their analxsis for the three cases mentioned with the more
accurate Sanders equations [11.19]. As can be seen from the results shown in Figures 11.23
and 11.24, for the simply supported boundary conditions SS 1 and SS2, the results obtained
with the two sets of equations are practically identical. For the antisymmetrical buckling of

o:ra
I \ A I I
~ ANTIS't'MWETRIC
070
<>4
- r\
oss I- ~
""'
I--
ii
\ ~FOR A,.u,_ IN ANTISY...ETRIC CASE
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Q2 -·-

~
r-- 0.1 I~ r-----.
DO.IQ~~ "-.........._ r--
0 =~· 1-
A_ SYMMETRIC

..... I
I 4 1 10
I
40 70 100 200 400
20 1000

L/.fih

Figure 11.14 Critical buckling load A, plotted against reduced length LlvRh (from [2.69]) (SS2
Copyright Wiley
boundary conditions: u = N" = w = M,
Provided by IHS Markit under license with WILEY = 0)
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
880 Boundary Conditions and Loading Conditions

w
SYMMETRIC

0~-~~~'·.,
0 ~ ~ ~ ~ ~

x*/(L/2)

Figure 11.15 Buckling modes for SS2 boundary conditions (from [2.69]) (the values of the constants
for the curves are given in Table 11.2)

shells with free edges two cases were investigated. As can be seen from the results of the
computations shown in Figure 11.25, there are relatively small but noticeable deviations
when the reduced wave number (and hence n) is small. For larger values of (and hence n n
n) the agreement between the two theories is good.
Summarizing the results of [2.69] dealing with the effect of boundary conditions with
membrane prebuckling, it appears that for isotropic shells the critical buckling load ,\ is
almost independent of the length of the shell, except when the shell is very short and rela-
tively thick-walled. For all four types of clamped boundary conditions the critical buckling
load of not-too-short cylinders is ,\ = 1.0. The same is true for two of the four simply
supported boundary conditions, namely SS3 and SS4. However, for simply supported shells
the in-plane boundary condition in the circumferential direction has a strong effect on the
critical buckling load. If the condition v '= 0 is replaced by the condition that the shell edge
be free of traction in the circumferential direction, that is, if N"" = N"' = 0, then the critical
buckling load is reduced to only half of its classical value, that is ,\ = 0.5. The authors of
[2.69] felt that this strong influence of a detail in the definition of one of the boundary
support conditions upon the critical buckling load ,'\. could very well be one of the reasons
for the large scatter of the experimental results in Figure 3 .1.
Thus, Hoff and Soong pursued this question further in a later paper [ 11.20]. Assuming
symmetry with respect to x* = 0, the mid-plane of the shell, they represented the solution
of Eq. (11.45c) by the following series:
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

w = Re {,~2 [A, cosh p, x + B, cosh p, x


1 12
(11.52)

+ C, cosh p,,x + D, cosh P,ax] cos n,cp}


where n, is the reduced wave number
-n, = m {if;;
Ia: m = 2, 3, 4, ... (11.53)

Further, they assumed that Am is a real number, while Bm, Cm and D, are complex coefficients
to be determined from the boundary conditions. The complex numbers p, 1, p, 2 and their
conjugates Pmt• p, 2 are the roots of the characteristic equation of Eq. (11.45c). The symbol
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Buckling of Circular Cylindrical Shells 881

Table 11.2 Values of constants for Figure 11.15 (from [2.69])


u\/Rii n= n Vacti2E A,
A 95.0 0.12 0.516
B 63.2 0.16 0.524
c 31.6 0.21 0.548
D 95.0 0.06 0.5023
E 63.2 0.06 0.5021
F 31.6 0.06 0.5019

Re indicates that the real part of the expression in the bracket is to be used for w, the
nondimensional radial displacement.
The ratios (Bm/A,), (Cm/Am) and (D,!Am) are then determined from the three boundary
conditions that are to be enforced at x* = L/2 around the entire circumference of the shell.
Notice that with these ratios known the expression for w given by Eq. (11.52) reduces to a
series with a singly infinite set of real coefficients Am. The values of the coefficients Am are
then calculated from the fourth boundary condition, which changes stepwise around the
circumference [see for example Eqs. (11.54b)-(11.54c)]. By satisfying this boundary con-
dition approximately in the sense of Galerkin, one obtains the appropriate number of alge-
braic equations needed to evaluate the constants Am. Since these equations are homogeneous,
according to Cramer's rule they have a nontrivial solution if and only if the determinant of
the coefficients of A, vanishes. This occurs only for discrete values of A. The critical eigen-
value A, is the lowest one. Details of the computations are given in the Appendices to [11.20].
Actually, the critical value of A should be calculated from a large enough set of equations
so that convergence to the correct value of A,. is achieved. In [11.20] Hoff and Soong dem-
onstrated the convergence of this approach for the two extreme cases of cjJt) = 0 and ¢r; =
1r, where cjJj denotes the location of the stepwise change around the circumference. To
calculate the results plotted in Figure 11.26, at x* = L/2, the following boundary conditions
were used
For the curve labeled SS I and SS3
Nx = W = W,xx = 0 0 ::::: ¢* ::::: 1T ( 11.54a)
(over the entire circumference of the edge)
v = 0 0 ::::: ¢* < c/Jt: (11.54b)

N"" = 0 ¢?: ::::: ¢* < 1T (11.54c)


--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(stepwise change around the circumference)


For the curve labeled SS2 and C2
u = Nn = w = 0 0 ::::: ¢* ::::: 1T (11.55a)
(over the entire circumference of the edge)
w,, = 0 0::::: ¢* < c/Jri (11.55b)
M, = 0 c/J!S ::::: ¢* < 1T (11.55c)
(stepwise change around the circumference)
It is interesting to see that according to the results given in Figure 11.26, a relaxation of
the respective constraints specified by Eqs. (11.54b) and (11.55b) around one-tenth of the
perimeter of the boundary is sufficient to lower the critical buckling load of axially com-
pressed isotropic shells by about a factor of one-half. The dimensions of the shell used for
the numerical example were so chosen by the authors of [2.69] that for the SS3 boundary
conditions the family of eigenvectors belonging to the repeated critical eigenvalue of A, =
1.0 also contains a buckling pattern with 10 buckles around the circumference. This implies
that in the example shown the faulty restraint must extend only over a distance equal to the
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
882 Boundary Conditions and Loading Conditions

0.1

Ol!

-...........
x ~
/NTISYMWETRIC
;YI414ETRIC

\
ii \\ I
/
S>1\'
1'\ /v v
03
I
\ \
Q2 ~ ,........__
V""-- f.--/
o. I

0
2 4 7 10 20 40 TO 100 200 400 TOO

L/.,fi.h

Figure 11.16 n
Plot of = n Y cr,J2E against reduced length LIYRh for ,\ 1.001 (from [2.69])
(SS4 boundary conditions: u = v = w = M, = 0)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

wavelength of the buckling pattern in order to produce a 50 percent decrease in the otherwise
expected critical buckling load of A, = 1.
As a continuation of this investigation, Kobayashi in 1966, in a little-known publication
[11.16], considered elastically supported boundary conditions. In his approach he extended
the work of Ohira [11.15] and [11.21] on semi-infinite shells to include an uncoupled spring-
like linear elastic support at the boundary x = 0 of the type

(11.56a)
(11.56b)
(11.56c)
where C1, C2 , and C3 are the spring constants of the rotational, axial and circumferential
constraints, respectively. Neglecting the effect of axial constraint (C2 = 0), he concluded that
if an experimental test setup has end-fixtures rigid enough such that the inequality

0 o~~~-o~.2~~~~~~~~o~G~~~~o~e~~~~~D-­

x*/{L/2}

Figure 11.17 Buckling modes for SS4 boundary conditions (from [2.69]) (the values of the constants
for the curves are given in Table 11.3)

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Buckling of Circular Cylindrical Shells 883

Table 11.3 Values of constants for Figure 11.17 (from [2.69])


LIVRh n = n Y a, 1 12E A,
A 2.4 0.1 1.305
B 4.4 0.35 1.009
c 2.4 0.1 1.244
D 4.4 0.35 1.122

(~) (c') ~
2
e~r >

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
12(1 2
v ) ( 11.57)
Elr E n h

where n is the number of circumferential full waves of the buckling pattern. is satisfied, then
the critical buckling load is A, = 1. That is, in such a case there is no decrease in the critical
buckling load due to the elastically supported boundaries.
At the time that Hoff and Soong [11.20] published their results, showing that for the
simply supported case by replacing the in-plane boundary condition v = 0 by N" = 0 over
only part of the perimeter at the shell edges, reduces the value of the critical normalized
buckling load by 50% to --\ = 0.5, it appeared that the corresponding SS 1 and SS2 boundary
conditions might offer an explanation for the low experimental buckling loads found in many
test programs. However, these results were obtained by using a membrane prebuckling anal-
ysis, thereby neglecting the effect of edge restraints. The effect of this simplification will be
investigated further in Section 11.3.2. Also, in practical applications one often encounters
some sort of elastic edge support. The effect of a rigorous prebuckling analysis on Kobay-
ashi's results dealing with elastic edge support [11.16] will be presented in Section 11.3.3.
Sobel [11.22] considered the effects of boundary conditions on the stability of cylindrical
shells subjected to hydrostatic-pressure loading. Later the pioneering work of Hoff and Soong
[2.69] was extended by Stavsky and his co-workers to edge-damaged filament-wound com-
posite shells under combined torsional and axial loads [11.23] and [11.24]. Unfortunately,
the remarks made in the preceding paragraph apply also here. The authors used a membrane
prebuckling analysis. Thus, the validity of their results and conclusions remain doubtful.
Confirmation by a rigorous analysis, where the etiect of edge restraint is properly accounted
for, is needed.

L/.fih

Figure 11.18 Critical buckling load,\ plotted against reduced length LIVRh for shells with clamped
edges (from [2.69])

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
884 Boundary Conditions and Le>ading Conditions

0.2 0.4 0.6 J.O

1.0

x*/(L/2)

Figure 11.19 Buckling modes of very short shells with clamped edges (from [2.69]) (the values of the
constants for the curves are given in Table 11.4)

11.3.2 Effect of Boundary Conditions Using Rigorous Prebuckling


Searching for an explanation of the serious disagreement between the predictions of the
classical theory and the experimental results of axially compressed isotropic cylindrical
shells, in 1963 almost simultaneously M. Stein [2.66] and G. Fischer [2.67] presented so-
lutions where for the first time the prebuckling displacements and stresses induced by the
edge support were rigorously taken into account. Using Foppl's solution [see Eqs. (2.203)-
(2.205)] for w 0 (x) they solved the resulting linearized stability equations [Eqs. (2.207a)-
(2.207b)] numerically. The results of the two analyses differ significantly since Stein used
SSl boundary conditions whereas Fischer worked with the SS3 boundary conditions. B.O.
Almroth in a later paper [2.68] presented the results of a more complete investigation, where
not only simply supported but also clamped edges were considered. His results for the well-
known eight different sets of boundary conditions are reprinted in Table 11.6. In this table

Table 11.4 Values of constants for Figure 11.19 (from [2.69])


L/YRh n= n Y a,.,I2E A,
A C1 3.16 0.1 1.435
C2 3.16 0.1 1.557
C3 3.16 0.1 1.564
C4 3.16 0.1 1.565
C1 6.32 0.35 1.006
C2 6.32 0.42 1.002
B C3 6.32 0.44 1.065
C4 6.32 0.45 1.071
c C3 6.32 0.42 1.008
D C3 3.16 0.45 1.307
C4 3.16 0.45 1.328
E C1 3.16 0.1 1.019
C2 3.16 0.1 1.019
Cl 6.32 0.42 1.002
C4 6.32 0.43 1.084
F
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`--- C2 6.32 0.43 1.004
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Buckling of Circular Cylindrical Shells 885

$YWMETRIC

c [-J \
['\_
).. SYMMETRIC

\'--- v
""'
OAII

Q30 ~OM " [',

I I....
""'
o.3&

r---- ~ ii FOR ~
v
QJO
"-..
Cl2ll

0.2!5 I
7 10 20 40 70 100 200 400 1000

Lt.fih
Figure 11.20 Critical buckling load,\ plotted against reduced length L!Viifi (from [2.69]) (shell edges
free, symmetric buckling)

for boundary condition identification the numbering given by Eqs. (11.49) and (11.50) is
used. The reader must be aware that in his original paper [2.68] Almroth employed another
numbering system to identify his boundary conditions. Note also that parts of Table 11.6
appear as Table 2.1 in Chapter 2.
Comparing the results of Table 11.6, where a rigorous prebuckling analysis was employed,
with those obtained by Hoff and Soong [2.69], who used a membrane prebuckling analysis,
it is evident that the cases SS1 and SS2 differ very little. However, for the other two simply
supported and the four clamped cases, within the parameter range considered, the critical
buckling loads with nonlinear prebuckling are significantly lower (up to 20% for the simply
supported and somewhat less for the clamped cases) than the corresponding critical buckling
loads of ,\ = 1.0 obtained using membrane prebuckling analysis.
Further, it appears from the results listed in Table 11.6 that in all cases, with the exception
of the very short shells (Z < 50), the critical buckling load varies little with the different
values of the parameters Ll R and R I h. However, contrary to what one would expect, lower
values of the critical buckling load A, are found in some cases for the very short shells.
Investigating this effect further, Figure 11.27 displays the dependence of the critical buckling
load A, on the parameter Ll R for an isotropic cylinder with Rl h = 100 with C3 boundary
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

conditions (N, = v = w = w,, = 0). Here the number of circumferential waves n = 8 was
kept fixed. When one looks at the results shown in Figure 11.27 it is interesting that whereas

~------~~~---,-------r------,-----~
L!Jii;
ANTISYMMETRIC
0.6 1--------,+':.::_158'-'0-",9'-'4-=-8-~---1--------1-------j

7.9

0.1 o.z Q3 0.4 0.5

Figure 11.21 Critical buckling load A, plotted against reduced wave number n= n V u,,I2E (from
[2.69]) (shell edges free, antisymmetric buckling)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
886 Boundary Conditions and Loading Conditions

:~*/ (L/2)

Figure 11.22 Buckling modes for shells with free edges (from [2.69]) (the values of the constants for
the curves are given in Table 11.5)

for long shells the critical buckling load A,. is independent of shell length and for very short
shells the critical buckling load A,. increases with decreasing shell length, both of these results
being as expected, in the intermediate range an oscillatory behavior is displayed. Notice that
the lowest critical buckling load A,. occurs for shells with an LIR = 0.7, say. For further
interesting results the reader should consult Alrnroth's original paper, [2.68].
An additional important difference that often occurs when critical buckling load calcula-
tions are carried out using both membrane and rigorous prebuckling analysis has to do with
the shape of the respective buckling modes. To illustrate this point consider the isotropic
shell with Rl h = 400 and LtVRh = 100, used by Hoff and Soong [2.69] to calculate the
results shown in Figure 11.26. Its critical buckling mode using membrane prebuckling and
SS3 boundary conditions (Nx = v = w = M, = 0) is shown in Figure 11.28a, and the
corresponding critical buckling mode, when rigorous prebuckling is employed, is displayed
in Figure 11.28b. Notice that in this case not only that the value of the critical buckling load
and the number of full waves in the circumferential direction have changed significantly, but
the axial shapes of the critical buckling modes are completely different. With membrane
prebuckling the axial shape is sinusoidal with five half-waves in the axial direction, whereas
with rigorous prebuckling it has a typical edge buckling-type shape where over half of the
meridional remains seemingly undeformed.
Thus, in concluding this section, whereas buckling load calculations based on the use of
membrane prebuckling analysis are in most cases appropriate for parametric and optimization
studies, the critical buckling load and the critical buckling mode shape of the final design
should always be verified by a rigorous buckling load calculation whereby the nonlinear
prebuckling deformations caused by the edge constraints are accounted for.

11.3.3 Effect of Elastic Boundal)r Conditions Using Rigorous Prebuckling


It is well known that using the variational approach, in addition to the equilibrium equations
one also obtains the set of natural boundary conditions that may be imposed at the shell

Table 11.5 Values of constants for Figure 11.22 (from [2.69])


LIVRh n= n YrJ"!2E A,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

A 31.6 0.1 0.0183


B 63.2 0.1 0.0546
c 31.6 0.3 0.363
D 63.2 0.3 0.378
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Buckling of Circular Cylindrical Shells 887

0.75

OJ 0
--DONNELL
----SANDERS

1\.'\ I
~ISYNNETRIC, n • 0.06
0.156 1

0.5
SYNNETRIC,
L'1--r-
n • 0.06
I

4 10 20 40 70 100 zoo 400 1000

Lt.fih
Figure 11.23 Comparison of critical buckling loads A, using Donnell and Sanders equations (from
[2.69]) (SSl boundary conditions: N, = N, = w = M, = 0)

edges. Almroth [11.25] assumed that for axially compressed isotropic cylindrical shells the
admissible sets of boundary conditions can be put into the form
(ll.58a)

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
(ll.58b)
(ll.58c)
(ll.58d)
where the Aii's are elastic spring constants and H is the reduced transverse shear stress
resultant, defined as
(11.59)
As a first approximation one may assume that only the diagonal terms are significant, that

0.75

010

066
[\ ---DONNELL

----SANDERS

~TI~YIIMETR!C

0.55

SYMNETRIC
" ~
- 'I'-
I
ll • 0.06r--
I

I
0 ..
·- -~-

0.4
2 4 7 10 20 40 70 100 200 400 1000

L/,fih

Figure 11.24 Comparison of critical buckling loads A, using Donnell and Sanders equations (from
[2.69]) (SS2 boundary conditions: u = N" = w = M, = 0)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
888 Boundary Conditions and Loading Conditions

ANTISYMNETRIC

.4
_,..-....
63.21 7(- - -
Q3

Q2 I 1~ jl
L/,fi.h • 31.6

0 .I
V1
I! If
I
I
--DONNELL
-----SANDERS

0
0 0.1 1l2 0.4 0.5

Figure 11.25 Comparison of critical buckling loads A,. using Donnell and Sanders equations (from
[2.69]) (shell edges free, antisymmetric buckling)

is one imposes the condition that A;1 = 0 if i =f. j. Introducing the following nondimensional
spring constants
1 R
A, (11.60a)
A11 Eh
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

I 1
A2 = An Eh2 (11.60b)

1 R
A=-- (11.60c)
3 A33 Eh

1 R
A=-- (11.60d)
4 A44 Eh
then, according to the sign convention used in [11.25], the above elastic boundary conditions
can be written (for total displacements)
- R
A1w + - H = 0 (11.6la)
1
E1

(11.6lb)

R
A3 u- Eh N" = 0 (11.61c)

R
A4 u- Eh N, = 0 (11.6ld)

In [11.25] the last elastic boundary condition was replaced by


N, = -N0 (11.62)
This replacement may only be used if the choice of boundary condition for loads or dis-
placements in the axial direction has little effect on the critical buckling load. This substi-
tution must be used with great care, for it may be incorrect. For instance, for axially com-
pressed stringer-stiffened shells, a prescribed end-shortening may produce an about 50%
higher buckling load than a prescribed axial load.

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Buckling of Circular Cylindrical Shells 889

1.0
R/h ~400 ! I /
0.9
' t-
LI/Rh ~100
0.8 ! I
Ac I
ss~ and SS3 - - - - .
07 !-------- r-~

06 -,~--
I
t
,_~
l- ~--

0 5~
! i
!
I i
---1 : t -
0 4 I
0 0 I 0 2 0.3 04 0.5 0.6 0 7 0.8 0 9 1.0

<l>o/1!

Figure 11.26 Effect of variations in the boundary conditions around the circumference of the shell on
the critical buckling load A, (from r11.201)

The above formulation of elastic boundary conditions can be used to verify the results of
Hoff and Soong [11.20] displayed in Figure 11.26. The curve labeled "SS 1 and SS3" can
be simulated by the following boundary conditions:
R
N, = -N0 , A3 u - Eh N,, = 0, w = M, = 0 (11.63)

Notice that_in the limit as A3 -+ 0 we have N, = 0, thus the SS1 boundary conditions;
whereas if A3 -+ x we get u = 0, thus the SS3 boundary conditions. The vari<~_tion of the
critical buckling load A, with the nondimensional tangential restraint parameter A3 is shown
in Figure 11.29 for an axially compressed isotropic shell with Rlh = 100 and LIR = 0.7
for different values of the number of circumferential waves n. Figure 11.30 displays similar
results for an isotropic shell with Rl h = I 00 and Ll R = 2.2. As can be seen from the results
displayed in these figures, already for relatively small values of the tangential restraint pa-
rameter A1 (A 3 > 0.7, say) the shells behave as if they were fully restrained in the circum-
ferential direction.
To simulate the other curve from Figure 11.26, the one labeled "SS2 and C2," one must
employ the following boundary conditions:

(11.64)

Notice that in the limit as A"-+ 0 one gets M, = 0, thus the SS1 boundary conditions,
whereas if A"-+ x one obtains w,, = 0, thus the Cl boundary conditions. As can be seen
from Table 11.6, for R I h = I 00 and Ll R = 0. 7 the difference between the critical buckling
loads for the SS1 boundary conditions (A, = 0.508) where N, = -N0 and the SS2 boundary
conditions (A, = 0.513) where u = -u0 is very small. Thus, one may assume that for this
case the two boundary conditions yield essentially identical results.
The variation of the critical buckling load A, with the nondimensional rotational restraint
parameter A" is shown in Figure 11.31 for an axially compressed isotropic shell with Rlh =
100 and Ll R = 0.7 for different values of the number of circumferential waves n. Figure
11.32 displays similar results for an isotropic shell with R I h = 100 and L/ R = 2.4. As can
be seen from the results displayed in these figures, already for relatively small values of the
rotational restraint parameters A2 (A 2 > 0.3, say) the shells behave as if they were clamped
all the way around the circumference.
In observing Figures 11.29-11.32 it is seen that the curves for the different numbers of
waves in the circumferential direction are sometimes arranged in a disorderly manner. The
reason for this is that the values of the critical buckling load as a function of the circum-
ferential wave number n may have two minima, as for the case shown in Figure 11.33. It is
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
~

Table 11.6 Critical buckling load .\ for different boundary conditions using rigorous prebuckling analysis (from [2.68])
Critical buckling load A, and number of waves in the circumferential direction n

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Rlh LIR z SSJ SS2 SS3 SS4 Cl C2 C3 C4
10 4
0.07 47 0.499(2) 0.507(2) 0.805(79) 0.872(85) 0.862(82) 0.908(83) 0.862(82) 0.908(83)
1 0.16 244 0.501(2) 0.504(2) 0.846(84) 0.864(86) 0.913(86) 0.926(88) 0.914(86) 0.926(88)
1 0.24 549 0.501(2) 0.503(2) 0.843(79) 0.868(86) 0.906(78) 0.927(87) 0.908(78) 0.927(87)
]04 0.32 977 0.501(2) 0.503(2) 0.844(82) 0.867(84) 0.908(82) 0.926(86) 0.910(82) 0.926(86)
10' 0.222 47 0.500(2) 0.507(2) 0.807(25) 0.874(27) 0.861(26) 0.907(27) 0.863(25) 0.907(27)
1 0.506 244 0.501(2) 0.505(2) 0.846(27) 0.863(27) 0.913(27) 0.927(27) 0.914(27) 0.927(27)
1 0.760 551 0.502(2) 0.503(2) 0.842(25) 0.867(27) 0.906(25) 0.930(28) 0.908(25) 0.930(28)
10' 1.014 981 0.502(2) 0.503(2) 0.844(26) 0.867(27) 0.908(26) 0.926(27) 0.910(26) 0.926(27)
102 0.7 47 0.508(2) 0.513(2) 0.806(8) 0.876(9) 0.862(8) 0.910(9) 0.863(8) 0.910(9)
1 1.6 244 0.511(2) 0.512(2) 0.849(8) 0.858(8) 0.916(9) 0.929(9) 0.917(9) 0.929(9)
1 2.4 549 0.511(2) 0.511(2) 0.843(8) 0.868(9) 0.903(8) 0.930(9) 0.908(8) 0.930(9)
102 3.2 977 0.510(2) 0.510(2) 0.844(8) 0.868(9) 0.909(8) 0.928(9) 0.911(8) 0.928(9)

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Buckling of Circular Cylindrical Shells 891

Effect of rigorous prebuckling on the critical buckling load


isotropic shell - R/h= 1 OO.n=B - C3 boundary condition
0
0

u
0
u
llO
E~
0

1J
0
0
·..
Ol' \'b-o-G,
.~ 6 Q

.3. ·····a.
Y. ···o-·-'(1 ·-G-e. e-· o ..o- ···.e-.-···-··<&
u
J
1l

0.
-~ tC
....: 0

1J

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

]~
0
z

1.20 2.00
Length over radius ratio L/R

Figure 11.27 Effect of shell length on the critical buckling load -\ (C3 boundary conditions:
N, = u = w = w,, = 0, rigorous prebuckling)

fwv+W 0 fwv+W 0
-.3

-.l

r
-.2
-.2

-.1
-.1

.0 .0
~~;(mo•J•L/2 ~ ,.(mo~J-L/2 -+
o. Prebuckling Shape a. Prebucklin9 Shape
.1 IOI!d .1 10ild

-1.0 -1.0

A
. v.
0.0 0.0 /"-

•(mo•)•l/2

b. Buckling Component Modes


- \ ..V i(mo~)•L/2

b. Buckling Component Modes


-
a. Membrane prebuckling b. Nonlinear prebuckling
Ac = 1.003, n =10 1cc = 0.844, n = 16

Figure 11.28 Effect of the prebuckling analysis used on the shape of the critical buckling mode (SS3
boundary conditions: N, = u = w = M, = 0)

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
892 Boundary Conditions and Loading Conditions

Effect of elastic tangentiol ~estr-aint on the critical buckling load


isotr-opic shell - R/h=100 - L/R=0.7

a0 0~ ti

" 0
"E n a 5
2

"a
n K 2
n /= & L.. .-.... -· n • 6 n = 7

7~--------------~-~·8
2 0
.,.~
-~ 0

_;"
a
E
a' o0
z0

1.60
Nondimensionol elas·:ic tangential ,..estraint lornbdc;J3-bor

Figure 11.29 Critical buckling loads of cylinders with elastic tangential restraint (boundary conditions:
N, = -N0 , X,v- (R/Eh) N" •= 0, w = M, = 0)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Cffect of e>astic tonger>tiol -estraint on the cr-itical buckling loaa


isotropic shell - R/h= 100 - L/R=2.2
0 r-----~------r-------·-----,,-----------~------~------
~

0 0 n • &
o m n • 2 n
" 0
"E
~

"~a 0 n

.,.• /
/
/

' 0
/
i
0 //
'
"0 /
/
/

-~ ~ /
/

"fi 0 /
/
/

"-~ /
/
/

0 • I
:"/
E
' 0
0 0
z 0

o.oo
Nondi.....,onslonol elastic tonguntiol restroint lombdo.3

Figure 11.30 Critical buckling loads of cylinders with elastic tangential restraint (boundary conditions:
N, = -N0 , X,v- (R/Eh) N,, ·= 0, w = M, = 0)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Buckling of Circular Cylindrical Shells 893

Effect of elastic r-otational restraint on the critical buckling load


isotropic shell - R/h=-100 - L/R=0.7

n ~ 6
0 0
0 ~
TI 0 n ~ 7
D
E
~ n ~ 8
TI
0
~ 0
~~
~ 0

ii

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
0 0

'"5 '0
TI

~0
E
' 0
0 ~
z 0

0.02 0.10
Nondirnensionol elostic rototionol restraint lo...,.,bdo2-bor

Figure 11.31 Critical buckling loads of cylinders with elastic rotational restraint (boundary conditions:
N, = -N0 , N,. = w = 0, A2 w,, + (I/Eh 2 ) M, = 0)

Effect of elastic rotational restraint on t.-,e critical buckling load


isotropic shell - R/h=-100 - L/R=2.4

n __ ....... __ _
. . D. ;;; 6
---------------~~-! ________ _
u 0
0 ~ n = 8
0 0
D
E
~ 3
0
0
~ 0
"'~
.S o

"
0 L---~----~--~-----L----~--~----~--~----~--~
0.04 o.oe

Figure 11.32 Critical buckling loads of cylinders with elastic rotational restraint (boundary conditions:
N, = -N0 , N,, = w = 0, A2 w, + (l!Eh 2 ) M, = 0)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
894 Boundary Conditions and Loading Conditions

Variation of critical buckling load with wove nunober C1 B.C.


isotropic shell - R/h= 100 - L/R=2.4

•.
u
c
"0
Do
Eci
0

"0
0
0

[Jl~
(.
·- 0
~
u
J
D

0 •
-~ 0!
~ 0

"0

N

p0
z

4.00 8.00 ,2.00


Nurnber of circumferential full

Figure 11.33 Variation of the critical buckling load with the circumferential wave number n (Cl
boundary conditions: N, = -N0 , N, = w = w,, = 0)

interesting to investigate the reason for this further. As can be seen from the buckling mode
shapes displayed in Figure 11.34, the two minima at n = 5 and n = 8 have completely
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

ditierent meridional shapes. The meridional shape of the buckling mode at n = 6 represents,
on the other hand, a transition from the half-wave sine-dominated shape at n = 5 to the
edge-buckling profile at n = 8.
Using Alrnroth's results shown in Figures 11.29-11.32, the critical buckling load of axially
compressed perfect cylindrical shells can be determined if the stiffness of the supporting

t·:~~f
I "i'(mc~)-L/2 -

a. Prebuckling Shape c. Prebuckling Shope

111'1 ooltd

b. Buckling Component Modes b. Buckling Component ~odes

a. "-c= 0.937, n"'5 b. <c• 0 950, n•6 c. >-c = 0.907, n .. a

Figure 11.34 Variation of the critical buckling modes with the circumferential wave number n (Cl
boundary conditions: N, = -N0 , N" = w = w,, = 0)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Buckling of Circular Cylindrical Shells 895

substructure is known. In order to facilitate a quantitative understanding of the above-men-


tioned curves, Almroth computed the critical buckling loads of cylinders supported by cen-
trally placed end rings with a square cross-section. If the side length is vCh, then the ring
area is equal to Ch 2 and the previously defined nondimensional elastic constants become

(11.65a)

(11.65b)

(11.65c)

Referring to [2.68], the variation of the critical buckling load with the size of the end ring
is illustrated by two different cases for an axially compressed isotropic shell with R I h =
100 and L/ R = 0.7. For the first case shown in Figure 11.35 the end ring is radially supported
(w = 0) and its torsional stiffness is omitted (M, = 0) so that it provides support only against
~ngential displacement. Notice that in the limit as the end ring becomes vanishingly small
A, ......, 0 agd one gets N,,. = 0, thus the SS 1 boundary conditions, whereas for a very large
end ring A1 ......, x and one obtains v = 0, thus the SS3 boundary conditions.
For the second case displayed in Figure 11.36 the torsional stiffness of the end ring is
included but now the radial support is removed, thus modeling a rigg-~pport~ free end.
Notice that in the limit as the end ring becomes vanishingly small A1, A2 and A3 ......, 0 and
one gets H = M, = N" = 0, the free edge case. In the other limit as the end ring becomes
very large A1, A2 and X,......, x and one obtains w = w,, = v = 0, thus the clamped C3
boundary conditions.
From the results shown in the last two figures it appears that a very small end ring is
sufficient to make the critical buckling load of the shell supported by end rings approximately

Effect of supported end-ring on the critical buckling load


isot.--opic shell - R/h= 1 00 - L/R= 0. 7
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

0
0
1J
Do
Eci n ~ 2
0

n • 8
-~~
" 0
J
D

0 0
-~ ~
-~ 0

1J

N

0~
E=
0
z

Nondimensional elastic end-r-ir.g coefficient C

Figure 11.35 Critical buckling loads of cylinder with radially supported end ring (boundary conditions:
N, = -N0 • A1 u- (R!Eh) N" = 0, w = M, = 0)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
896 Boundary Conditions and Loading Conditions

variation of critical buckling load with size of the end-ring


isotropic shell - R/h""" 100 - L/R=O. 7

0
0 0
Uo
n =2 (sym)

Eci n = 8 (sym)
0

1J
oo
Q'!
-o
"'
c
:i
0 0
J'
.Oci
0
-~
·c
o
0.
ci
1J

N

0
Eg I
c . I
0 0 1
z

o.oo 4.00 8.00 12.00 1 6.00 20.00


Nondimensionol e astic end-ring coefficient C

Figure 11.36 Critical buckling loads of cylinder with end ring (boundary conditions: N, -N0 ,
X,v- (R/Eh) N" = 0, X,w + (R/Eh) H = 0, A2 w,, + (l/Eh 2 ) M, = 0)

equal to the critical buckling load of a shell with full constraint. Thus, in Figure 11.35 for
a quadratic end ring with a side of 2.25h (i.e. C = 5.06), one almost attains the full critical
buckling load for SS3 boundary conditions, instead of buckling at the much lower value
corresponding to the SSl boundary conditions. Similarly, according to the results of Figure
11.36 it takes only a somewhat larger quadratic end ring with a side of 2.95h (i.e. C = 8.70)
to approximate closely the critical buckling load for the clamped C3 boundary conditions,
instead of buckling at the relatively low values corresponding to the free edges boundary
conditions.
The results displayed in Figure 11.36 were later corrected by Cohen [ 11.26], who pointed
out that apparently Almroth erroneously used symmetry halfway along the shell at x = 0
for the free edge case for both the n = 2 and n = 8 curves. As pointed out by Hoff and
Soong in Figure 11.21, the use of antisymmetry at x = 0 yields very low critical buckling
loads for the free edge case. Using Cohen's SRA program [3.19] the corrected curves shown
in Figure 11.37 were calculated. Notice that there exists a critical size of the end ring, below
which the ring strain energy controls the buckling, and as such using rings weaker than this
critical value results in edge buckling at relatively low buckling loads, involving an inexten-
sional buckling mode with n = 2 full waves in the circumferential direction with large
deformations of the end ring itself. Further, it must be mentioned that the computer program
SRA works with discrete rings. Thus, the results shown involve a rigorous solution of
Cohen's nonlinear ring equations [11.27] including the effect of all the off-diagonal terms
of the ring stiffness matrix G,1 [11.28] representing coupling between the shell displacements
and rotations. The need to include these off-diagonal terms of Gu has been confirmed by
Dixon et al. [11.28], who have shown that if the Gu matrix represents the restraint of an end
ring, the use of only the diagonal terms in Gu will lead to serious inaccuracies in the predicted
buckling load and buckling mode (see also Figure 11.38). Finally, notice that in Figure 11.37,
according to this rigorous approach it takes a moderate size quadratic ring with a side of
11.6 h (i.e. C = 135) to approximate closely the critical buckling load for the clamped C3
boundary conditions instead of buckling at relatively low values corresponding to the inex-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

tensional n = 2 buckling modes.


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Buckling of Circular Cylindrical Shells 897

Variation of critical buckling load with size of the end-ring


isotropic shell - R/h= 100 - L/R=O. 7

0
0 0 n s: 2 (anti-sym)
"Do
.oc
E
.!'
~-~-~--~--------
"0 .-(Y_..#ir-..0--e-- n = 8 (sym)
0 0
0 ~

l""
- 0
U>

•~•
c
:i
0 0
J •
.Oci ~
0 ~
-~
·,: 0
0 •
0
"0

N

t~
0 0
z

4-0.oo !!lo.oo 12o.oo 1 e.o.oo 2oo.oo


Nondirnensionol elastic end-ring coefficient C

Figure 11.37 Critical buckling loads for axially compressed isotropic shells with symmetrically placed
symmetrical end ring

11.3.4 Load Eccentricity Effects


Load eccentricity is usually defined as the radial distance between the line of application of
the axial load and the midsurface of the shell wall. In early shell buckling investigations one
assumed that the axial load was applied through the midsurface of the skin. However, it was
soon realized that the end moments due to the eccentricity of the load introduction may
affect the buckling load noticeably.
In 1966 Stuhlman, de Luzio and Almroth [11.29] presented a theoretical evaluation of a
test program at Lockheed Missiles and Space company dealing with the influence of stiffener

Ret. 11.27

Rel.11.25

1.0

1'
/
~,..

~ Ref.11.27
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(Gjj diagonal)

Ref. 11.27
(G;j full)

.I

,<@.tr·_.,v Section A-1\


.01~~----~--~----~--------J
0 4 d 12
l'i

Figure 11.38 Effect of edge-restraint coupling on buckling of a cylinder with square end rings subjected
Copyright Wiley to axial compression (from [11.28]-L/R = 0.7, Rlh = 100)
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
898 Boundary Conditions and Loading Conditions

eccentricity and end moment on the stability of axially compressed cylindrical shells. This
work was motivated because of the poor correlation between van der Neut's theory [11.30]
and the test results mentioned above.
It was felt that the manner of load introduction, shown in Figure 11.39, was the reason
for the discrepancy. By the load being introduced in the midsurface of the skin, the stringer-
stiffened shell was actually loaded by the combination of an axial load applied at the centroid
of the cross-section plus an axisymmetrical end moment. In order to account for the beam-
column effect caused by the interaction of axial loads and lateral displacements, Stuhlman,
de Luzio and Almroth based the calculation of the prebuckling deformations on nonlinear
equations and the end moments were introduced into the problem via the boundary condi-
tions. Since the prebuckling state is axisymmetric, it is possible to obtain a closed-form
solution. The resulting variable coefficients equations were then solved by applying Galer-
kin's method. As can be seen from Figure 11.40, the correlation between test results and
theoretical predictions is good. The maximum discrepancy is less than 15%. Thus, it appears
that the inclusion of end moments is indeed necessary for a better correlation, especially for
shorter shells (LIR < 0.7 for R = 198.0 in., say) as can be seen in Figure 11.41.
More recent studies [13.33], [13.38], [13.42] and [13.51] have indicated qualitatively that
end moments, which tend to bend a shell into a barrel shape (and thus give rise to tensile
prebuckling hoop stresses), increase the compressive buckling load in the axial direction.
This means, for example, that for extemally stringer-stiffened shells loading through the
midsurface of the skin (thus zero load eccentricity) yields higher buckling loads than appli-
cation of the load at the stringer centroid. For internally stringer-stiffened shells the opposite
is true. In all the above-mentioned theoretical studies similar trends of increase or decrease
in buckling load with load eccentricity were found, except by P. Seggelke and Geier [3.11],
who arrived at opposite trends.
The influence of load eccentricity on the critical axial buckling load suggests that the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

design of end fittings especially for simply supported stringer-stiffened shells may have a
strong influence on the load-carrying capacity of a shell. Thus, one must be aware of that
the rigidity of the actual end fittings may become an important factor since end moments
may not be transmitted to the shell but carried largely by the end fitting itself.
In the seventies a very extensive test program using 7075-T6 aluminium alloy stringer-
stiffened cylindrical shells was carried out at the Technion [13.43]. The purpose of this
program was to investigate the load eccentricity effects experimentally. The dimensions of
the test specimens were based on the results of a parametric study carried out using the
nonlinear theories of D.L. Block [1.38] and Almroth, Bushnell and Sobel [13.33]. As can
be seen from the results displayed in Figures 11.42 and 11.43, for lightly stringer-stiffened
shells the highest critical load is achieved when the load is introduced through the neutral
surface of the skin-stringer combination, whereas for heavily stringer-stiffened shells the
highest critical load occurs when the load is applied at the shell midsurface. In summary,
the line of load application that yields the highest critical load of an externally stringer
stiffened shell is dependent on the dimensions of the stringers. This was a new result not
noted in any of the earlier studies mentioned above.
The specimens were manufactured as a long shell and then cut into a number of shorter
specimens with the ends machined so as to satisfy one of the end conditions shown in Figure
11.44. Thus, the effect of the different manner of load introduction could be examined for

TESTING
MACHINE

Copyright Wiley Figure 11.39 Load end configuration (from [ 11.29])


Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Concluding Remarks 899

OUTSIDE ~
1

STIFFENING
0

G~ ECCENTRICITY AND
END MOMENT

~~
~
ECCENTRICITY

INSIDE t-----1--
STIFFEN I t : ~TRICITY

I
ECCENTRICITY AND
END MOMENT
LEGEJD,
r-o TESTS WITH END MOMENT AND
ECCENTRICITY EFFECT
TEST WITH ECCENTRICITY
EFFECT ONLY
0 I
80 90 100 110 120 130
LENGTH, L (IN.)

Figure 11.40 Comparison of test results and theoretical plots (from [ 11.29])

shells of practically identical geometry. For further details of the test setup and the test
procedure the interested reader should consult [13.43].
The experimental results confirmed in general the influence of the eccentricity of loading
as predicted by the nonlinear theories used for the parametric study. However, it also became
clear that the manner of load application and the details of the end fixtures are of great
importance in determining which of a variety of possible behavior patterns will occur. Thus,
further experimental and numerical investigations are warranted.

11.4 Concluding Remarks


To summarize the main features of this chapter, it has been shown that the buckling behavior
of axially compressed columns is mainly influenced by the out-of-plane boundary conditions.
In the same way the prebuckling and buckling behavior of plates carrying in-plane loads
depend heavily on the out-of-plane boundary conditions, especially the ones along the un-
loaded edges. However, the postbuckling behavior of plates carrying in-plane loads is
strongly influenced by the in-plane boundary conditions. For shells the prebuckling behavior
depends mainly on the out-of-plane, whereas the buckling and postbuckling response is
influenced by both the in-plane and the out-of-plane boundary conditions.
The importance of using the appropriate boundary conditions can best be illustrated by
the following facts.
From the results presented in Section 11.1 it becomes evident that the buckling load of a
column clamped at both ends is 8 times higher than the buckling load of a column with one
end clamped and the other end free. Considering plates loaded by in-plane compression. it
can be seen from Figure 2.5 that the buckling load of a plate with clamped boundary con-
ditions along the unloaded edges is 17.5 times higher than the buckling load of an identical
plate with simple support on one and free boundary condition on the other unloaded edge.
For axially compressed isotropic shells, as can be seen from the results shown in Table
11.6, the in-plane force boundary condition N" = 0 results in an about 50% decrease in the
critical buckling load as compared to zero displacement edge condition (v = 0) in the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

circumferential direction. On the other hand, as can be seen in Table 11.7, for the lightly
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
900 Boundary Conditions and Loading Conditions

LENGTH, l (IN.)

Figure 11.41 Critical load vs. length for test specimen geometry (from [11.29])

stringer-stiffened shell AS-2 prescribing the end-shortening (u = Llu) instead of the axial
load N, = -N0 can result in an about 50% increase of the critical axial buckling load.
Another important consideration is that a variation may exist in the circumferential direc-
tion of the boundary conditions imposed at the shell edges. For instance, as one can see
from the results displayed in Figure 11.26, relaxing the constraint v = 0 over about one-
tenth of the perimeter of the boundary is sufficient to lower the critical buckling load of

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
axially compressed isotropic shells by about a factor of one-half.
In particular applications one will in many cases encounter some kind of elastic support
at the edges. This has been recognized for many years and papers dealing with (partial)
elastic restraints have been published regularly (see for example [2.4], [11.6], [11.16] and
[11.25]). Nowadays, with the tremendous increase in computer power, investigators can
model not only the plate or shell in question very accurately but also the adjacent structural
elements that provide the elastic edge support, making a realistic simulation of the buckling
process possible. Correlation of these simulations with experiments however, is still missing.

0 BLOCK ss 3 N11. IKt/mm)


0 BOSOR SS3
e BLOCK SS4
• BOSOR SS4

-IOh 0 +5h ... IOh


INTERNAL ECCENTRICITY EXTERNAL ECCENTRICITY

Figure 11.42 Effect of eccentricity of loading for a lightly stringer-stiffened shell (from [13.43])
(AJd,h = 0.78; ElJdp = 30.1; eJh = 3.76; Rlh = 510; LIR = 1.01)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Concluding Remarks 901

50

0 BLOCK SS3 ..& - ..1fJ


0 BOSOR SS3 /
__...-fl/
_..8/ ~0
s-~

20

10

.; 50 40 30 20 10 0 l\'lm 10 20 30 40 50 +i
INTERNAL ECCENTRICITY EXTERNAL ECCENTRICITY

Figure 11.43 Effect of eccentricity of loading for a heavily stringer-stiffened shell (from [13.43])
(AJd,.h = 1.44; ElJdp = 910.0; e)h = 10.5; Rlh = 2000; L/R = 1.0)

The effects of different types of loading conditions on the critical buckling loads are treated
in Chapter 8 for plates, in Chapter 9 for shells and in Chapter 12 for stiffened shells. Further
details of the influence of eccentricity of loading on the critical buckling load are presented
in Chapter 15. The effect of the sequence of loading on interaction curves for buckling of
stiffened shells is discussed in [11.31]. The influence of nonuniform axial loads on the
buckling of cylindrical shells was investigated by Durban and Libai in [11.32] and [11.33].
The effect of the unevenness of end supports was treated by Esslinger, Geier and Heidemann
in [11.34].
Finally, it must be clearly stated that in order to be able to model accurately the boundary
conditions that occur in practice, one must carry out more carefully planned and ingenious
(see for example [11.35]) combined experimental and numerical research projects. The ex-
perimentalist must provide the numerical analyst with the necessary input values, and the
results of the numerical predictions must be verified by test results in order to establish user's
confidence in the procedures used. One promising approach to experimental determination
of boundary conditions for shells, the vibration correlation technique (VCT), is discussed in
Chapter 15.
Further, there is a dire need for more experimental studies of the effects of boundary
conditions on the buckling and in particular the postbuckling behavior of plates and shells.
Only when more of such investigations have been performed will the much-needed verifi-
cation of the various numerical studies become feasible, leading to a more precise evaluation
of the important boundary effects. What a challenge to experimenters!

JStrlnger

u~·
LoadinQP and supporting discs
ielded portion
of stringer
i -Eccentricity
of loading
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

DETAIL A, Load DETAIL B, Lood DETAIL C: Load


applied through applied through applied through
shell mid-skin stringer tips end-ring

~~:;:-· ;;J~~ ~~:-·


p p p
DETAIL D: Restrained
radial displacement- Restrained radial displacement of striAgers-
load applied through load applied through strino•r tipa
mid-skin

Figure 11.44 Details of load application (from [13.43])


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
902 Boundary Conditions and L.oading Conditions

Table 11.7 Critical buckling load A:>~ for different boundary conditions using rigorous prebuckling
analysis (lightly stringer-stiffened shell AS-2 of [1.25])
SSl: A:>~= !.39755 (n = 10)- sym. at x = L/2
SS2: 1.65546 (n = 13) - antisym. at x = L/2
SS3: 1.42990 :n = 10)- sym. at x = L/2
SS4: !.76862 (n = 14) - antisym. at x = L/2
C1: !.61389 (n = 10)- sym. at L/2
C2: !.99989 (n = 14)- antisym. at L/2
C3: 1.63503 (n = 10) - sym. at L/2
C4: 1.99996 (n = 14)- antisym. at L/2
For the definition of the boundary conditions see Eqs. ( 11.49) and ( 11.50). A;'' = N, IN, 1 where N, 1 = Eh' I cR =
~90.644lblin. = ~158.742 Nlcm

References
11.1 European Convention for Constructional Steel Work, Guide to Design Criteria for Metal Com-
pression Members, European Recommendations, Brussels, Belgium, 1988.
11.2 Houbolt, J.C. and Stowell, E.Z., Critical Stress of Plate Columns, NACA Technical Note 2163,
August 1950.
11.3 Lundquist, E.E. and Stowell. E.Z., Critical Compressive Stress for Outstanding Flanges, NACA
Report 734, 1942.
11.4 Hill, H.N., Chart for Critical Compre~.sive Stress of Flat Rectangular Plates, NACA Technical
Note 733, 1940.
11.5 Dunn, L.G., An Investigation of Sheet-Stiffener Panels Subjected to Compression Loads with
Particular Reference to Torsional Weak Stiffeners, NACA Technical Note 752, 1940.
11.6 Lundquist, E.E. and Stowell, E.Z., Critical Compressive Stress for Flat Rectangular Plates Sup-
ported along All Edges and Elastically Restrained Against Rotation along the Unloaded Edges,
NACA Report 733, 1942.
11.7 Lundquist, E.E. and Stowell, E.Z., Restraint Provided a Flat Rectangular Plate by a Sturdy
Stiffener along an Edge of the Plate, NACA Report 735. 1942.
11.8 Gerard, G. and Becker, H., Handbook of Structural Stability. Part I-Buckling of Flat Plates,
NACA Technical Note 3781, July 1957.
11.9 Stein, M., Analytical Results for Post-buckling Behaviour of Plates in Compression and Shear,
in: Aspects of the Analysis of Plate Structures, D.J. Dawe, R.W. Horsington, A.G. Kamtekar and
G.H. Little, eds., Clarendon Press, Oxford, 1985, 205~223.
11.10 Stein, M .. Postbuckling of Long Orthotropic Plates in Combined Shear and Compression, AIAA
Journal, 23, (5), May 1985, 788~ 794.
11.11 Lentini, M. and Pereyra, V., An Adaptive Finite Difference Solver for Nonlinear Two-Point
Boundary Value Problems with Mild Boundary Layers. SIAM Journal of Numerical Analysis,
14, March 1977.
11.12 Stein, M., Post buckling of Orthotropic Composite Plates Loaded in Compression, AIAA Journal.
21, (12), Dec. 1983, 1729~1735.
11.13 Brogan, F.A., Rankin, C.C. and Cabiness, H.D., STAGS User Manual, LMSC P032594, Lock-
heed Palo Alto Research Laboratory, Palo Alto, Calif., March 1994.
11.14 Hoff, N., Buckling of Thin Shells, in: Proceedings of an Aero5pace Symposium of Distinguished
Lecturers in Honor of Dr. Theodore von Kdrmcin on His 80th Anniversary, Institute of Aerospace
Sciences, New York, 1961.
11.15 Ohira, H., Local Buckling Theory of Axially Compressed Cylinders, in: Proceedings of the
Eleventh Japan National Congress for Applied Mechanics, 1961, 37~41.
11.16 Kobayashi, S.. The Influence of the Boundary Conditions on the Buckling Load of Cylindrical
Shells under Axial Compression, GALCIT SM 66-3, California Institute of Technology, Pasa-
dena, Calif., March 1966.
11.17 Nachbar, W. and Hoff, J., On Edge Buckling of Axially Compressed Circular Cylindrical Shells,
Quarterly of Applied Mathematics, 20, (3), Oct. 1962, 267~277.
11.18 Hoff, N. and Rehfield, L.W., Buckling of Axially Compressed Circular Cylindrical Shells at
Stresses Smaller than the Classical Value, Journal of Applied Mechanics, 32, (3), Sept. 1965,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

533~537.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 903

11.19 Sanders, J.L Jr,, An Improved First Approximation Theory for Thin shells, NASA TR-R24,
1959.
1L20 Hoff, N. and Soong, T.C., Buckling of Axially Compressed Circular Cylindrical Shells with
Non-Uniform Boundary Conditions, in: Proceedings Symposium of Thin-walled Steel Struc-
tures-Their Design and Use in Building, University College of Swansea, Sept. 1967, 61-80.
11.21 Ohira, H., Linear Local Buckling theory of Axially Compressed Cylinders and Various Eigen-
values, in: Proceedings 5th International Symposium on Space Technology and Science, Tokyo,
1963, 511-526.
11.22 Sobel, LH., Etiects of Boundary Conditions on the Stability of Cylinders Subject to Lateral and
Axial Pressure, AIAA Journal, 2, (8), August 1964, 1437-1440.
11.23 Stavsky, Y., Greenberg, J.B. and Sabag, M., Buckling of Edge-Damaged Filament-Wound Com-
posite Cylindrical Shells under Combined Torsional/ Axial Loads, Composite Structures, 13,
1989, 21-34.
11.24 Sabag, M .. Stavsky, Y. and Greenberg, J.B., Buckling of Edge Damaged Cylindrical Composite
Shells, ASME Journal of Applied Mechanics, 56, (1), March 1989, 121-126.
11.25 Almroth, B.O., Influence of Imperfections and Edge Restraint on the Buckling of Axially Com-
pressed Cylinders, NASA CR-432, April 1966.
11.26 Cohen, G.A., Buckling of Axially Compressed Cylindrical Shells with Ring Stiffened Edges,
AIAA Journal, 4, (l 0), Oct. 1966, 1859-1862.
11.27 Cohen, G.A., Computer Analysis of Ring-Stiffened Shells of Revolution, NASA CR-2085, Au-
gust 1973.
11.28 Dixon, S.C., Weeks, G.E. and Anderson, M.S., EJiect of Edge Restraint Coupling on Buckling
of Ring-Supported Cylinders, AIAA Journal, 6, (8), August 1968, 1602-1604.
11.29 Stuhlman, C., De Luzio, A. and Almroth, B.O., Influence of Stiffener Eccentricity and End
Moment on Stability of Cylinders in Compression, AIAA Journal, 4, (5), May 1966, 872-877.
11.30 Van der Neut, A., The General Instability of StiJiened Cylindrical Shells under Axial Compres-
sion, Nationaal Luchtvaartlaboratorium Report 314, 1947.
11.31 Weller, T. and Abramovitch, H., Effect of Sequence of Loading on Interaction Curves for Buck-
ling of Stiffened Shells, TAE Report No. 536, Technion-Israel Institute of Technology, De-
partment of Aeronautical Engineering, Sept. 1984.
11.32 Durban, D. and Libai, A., Influence of Thickness on the Stability of Circular Cylindrical Shells
Subjected to Nonuniform Axial Compression, Israel Journal ofTechnologv, 14, 1976,9-17.
11.33 Libai, A. and Durban, D., Buckling of Cylindrical Shells Subjected to Nonuniform Axial Loads,
ASME Journal of Applied Mechanics, 44, Dec. 1977, 714-720.
11.34 Esslinger. M., Geier, B. and Heidemann, V., Comments on the Paper Some Complements to the
ECCS Design Code Concerning Isotropic Cylinders, in: Preliminary Report for the Liege Col-
loquium, April 13-15, 1977, Interner Bericht IB 152-77/06, DFVLR Braunschweig, April 1977.
11.35 Lancaster, E.R., Calladinc, C.R. and Palmer, S.C., Experimental Observations on the Buckling
of a Thin Cylindrical Shell Subjected to Axial Compression, Report CUED/D-Struct!TR162,
Department of Engineering, University of Cambridge, May 1966.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
12
Stiffened Plates

12.1 Built-up Structures, Local and General Instability


Structures are usually built up of ditierent elements, each of which can fail separately. The
whole structure can, however, also fail globally, either before any local failure has occurred,
or after local failures of elements as a cumulative result. In thin-walled structures the pre-
dominant failure modes are related to instability, and hence in built-up structures there is
local buckling and global buckling, often called overall buckling or general instability.
The simplest case is when the failure is either by local buckling or general instability, as
for example in the compression of a strut of angle section, discussed in Chapter 4 (see Figure
4.18 and [2.1] and [4.31 )). There the local instability is plate buckling of the flanges and it
occurs first when the flanges are wide (on the right-hand side of Figure 4.18); while general
instability is column buckling, which occurs when the flanges are narrow (on the left-hand
side of Figure 4.18).
In general, however, there is significant nonlinear interaction between the local and general
instability modes, and failure of a structure may involve a combination of local and global
buckling. Early optimization studies called for simultaneity of the different failure modes,
but, as pointed out in Section 2.1.4 of Chapter 2, in the late sixties and early seventies it
was shown by several authors that simultaneous or nearly simultaneous buckling modes can
lead to dangerous compound modes of failure (see for example [2.17), [2.18], [6.102),
[6.103], [12.1) and [12.2)).
Mode interaction for columns and beams was considered in Section 6.5 of Chapter 6,
where typical interactive buckling tests were also discussed. For stiffened plates the mode
interaction can be more complicated, and since they are important structural elements they
warrant more detailed consideration.

12.2 Buckling and Postbuckling Strength of Stiffened Plates


12.2.1 Introduction
Nearly all plates used in real structures are stiffened plates, plates reinforced by ribs, or
stiffeners, longitudinally and or transversely. Such plates then buckle locally, as unstiffened
plates with boundary conditions determined by the stiffness of the ribs and the structure
beyond the local plate, or alternatively by general instability, with both plate and stitieners
buckling simultaneously. In fact, often the instability commences as local buckling of the
plate elements between the ribs, which continue to carry increasing loads in the postbuckling
regime (as discussed in Sections 2.1.2 and 8.12 of Chapters 2 and 8, respectively), till the
supporting structure of ribs buckles too, when general instability or collapse occurs.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
Buckling
No reproduction Experiments:
or networking Experimental
permitted without Methods
in Buckling of Thin-Walled
license from IHS Structures:
Not for Resale, 02/13/2019Shells,
01:32:58Built-Up
MST Structures, Composites
and Additional Topics – Volume 2. J. Singer, J. Arbocz and T. Weller Copyright © 2002 John Wiley & Sons, Inc.
906 Stiffened Plates

Since stiffeners or ribs very significantly increase the buckling loads of plates, stiffened
plates have been extensively studied, starting with Timoshenko's investigations in the first
decades of the 20th century (see [8.93], [2.1] and [6.46]). In the thirties, forties and fifties
many studies were primarily motivated by aircraft design (for example [12.3]-[12.7]) as well
as by civil engineering problems (for example [12.8]-[12.11]). Marine engineering then
stimulated much of the research in the sixties and seventies (for example [8.141], [12.12]
and [12.13]), and the 1970-71 bridge disasters in Europe and Australia added their induce-
ment to studies that continued into the nineties and beyond (for example [12.14]-[12.20]
and [6.3], chap. 4).

12.2.2 Analysis of Stiffened Plates


One of the approaches to the analysis of the general instability of stiffened plates has been
to consider them as orthotropic plates, assuming that the stiffeners are longitudinal and/ or
transverse and hence orthogonal to each other. In this approach the stiffeners are "smeared"
over the plate, with the simplest model neglecting their eccentricity, i.e. assuming that the

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
stiffeners are symmetrical above and below the plate. Later improved models take the ec-
centricity of the stiffeners into account (see for example [12.21]). Many studies have been
carried out with orthotropic plate theory and have been reviewed and discussed in some
well-known books (for example [12.22]-[12.24]). They have yielded fairly simple and con-
venient design methods (see for example [12.25]-[12.27]).
Corrugated plates and corrugated sandwich plates can also be treated as orthotropic plates
(see for example [12.28] and [12.29]).
The orthotropic plate approach has, however, some disadvantages: it cannot deal with non-
uniform plates or stiffeners, nor with unevenly spaced stiffeners, it cannot take into account
the discreteness of the stiffeners and it does not consider local plate buckling between stiff-
eners or of the stiffeners.
Another approach for analysis of stiffened plates is to consider the stiffeners as linear
discontinuities, represented by the Dirac delta function, instead of smearing them. This ap-
proach, developed for stiffened shells (see for example [13.35]-[13.37]), can also be applied
to stiffened plates and has the advantage that it also includes consideration of local buckling
between stiffeners. The Dirac delta function representation of the stiffeners in this approach
is satisfactory as long as the width of the stiffeners is not comparable to the distance between
them.
A third approach to the elastic buckling analysis of a stiffened plate is to divide the cross-
section into its elements of plate and stiffener sections and solve this assembly of plates as
a series of elements that are rigidly connected along common longitudinal edges. The trans-
verse variations are solved exactly, instead of employing finite element approximations (see
for example [12.30]-[12.32]). This approach has yielded a better understanding of the be-
havior of stiffened plates and has resulted in a much more efficient computational procedure
[12.33] than the employment of general-purpose finite element codes. The program VIPASA
(developed by Wittrick and Williams in the U.K.) is so economical that it was adopted by
NASA and the U.S. aerospace industry and incorporated into a NASA optimal panel design
code PASCO (see [12.34]).

12.2.3 Mode Interaction in Stiffened Panels


Mode interaction in stiffened panels has been the subject of many studies in the last two
decades because of its significant influence on their optimal design (see for example [12.35]-
[12.41]). Most of the analyses aimed at establishing the imperfection sensitivity of stiffened
panels, in particular for near optimal con:figurations. Some of them then concluded that in
the presence of typical imperfections the optimum does indeed shift from the concept of
simultaneous buckling towards a range where the overall buckling (or general instability)
stress is smaller than the local buckling stress.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Plates Subjected to Axial Compression 907

In the early seventies a careful experimental investigation of the interactive buckling of


uniformly eccentrically stiffened plates was carried out at University College, London [5.64]
and [12.42]. In these tests the overall imperfections of the panels were controlled by altering
the line of application of the loading, and small local imperfections of shapes resembling
the local buckling modes were also introduced. The results confirmed predicted buckling
loads and imperfection sensitivities and have since been used to verify some of the analyses
referenced above.
Further to their elastic buckling, the postbuckling behavior and collapse of stiffened plates
has also been widely studied, both experimentally and theoretically (see for example [12.16],
[12.36], [ 12.37] and [ 12.39]). Though the design methods have predominantly relied on
experiments, discussed below, many analytical and finite element studies have also been
reported (see for example [12.43] and [12.44]).

12.3 Experiments on Stiffened Plates Subjected to Axial


Compression
12.3.1 Local Buckling Tests
In the discussion on plate buckling experiments in Chapter 8 it was often mentioned that
some of the test rigs were also employed for buckling tests on stiffened plates. There is
similarity in the testing of unstitiened and stiffened plates, and the emphasis in this chapter
will therefore be on those elements particular to stiffened panels or on those test setups
specifically designed for stiffened plates. For example, the discussion on the choice of model
scale at the end of Chapter 8, in Section 8.7.1, and in Chapter 5 applies also to stiffened
plates and need not be repeated here. But there are some special problems relating to stiffened
plate experiments that warrant distinct consideration, as will be pointed out in the discussion
below of some typical test setups.
In many of the earlier experiments on stiffened plates, like those discussed in Chapter 8
(see Figure 8.41, [8.27] and Figure 8.107, [8.23]), the objective was the study of the buckling
of the plate between the stiffeners, of local buckling. The widely spaced and relatively heavy
stiffeners served only as supports for the test plate, while general instability was usually not
considered. As a matter of fact, even today, in some lightly loaded structural clements in
aerospace applications, design constraints lead to thin skins with relatively heavy stiffeners.
For example, the dorsal fin of the SAAB 2000 commuter aircraft (which entered service in
1994) is of such a construction, as shown in Figure 12.1, where the buckles in the skin were
photographed in a test flight and can be clearly seen. The panel can carry loads much beyond
the buckling load of the skin.

12.3.2 Early Small-Scale Tests


In the mid-forties the emphasis shifted to the general instability of closely stiffened panels,
primarily in the aerospace industry. As part of an extensive study of structural elements
suitable for the compression surface of wings, comprehensive investigations of the com-
pressive strength of flat, longitudinally stiffened panels were carried out at NACA Langley
Research Center (see for example [12.45]-[12.48]). A typical example of such a closely
stiffened 75S-T6 aluminum-alloy flat panel, with Z-section stiffeners, after failure is shown
in Figure 12.2. The panels were tested flat-ended without side supports (essentially as col-
umns) in a large testing machine. Prior to the test, the ends of the panels were ground
accurately flat and parallel and aligned in the testing machine to ensure uniform bearing.
The panel shown is one with more closely spaced stiffeners, out of the very large number
of specimens tested with varying stiffener spacing, stiffener size and panel length. The failure
in Figure 12.2 was clearly overall buckling (general instability).
In these NACA experiments large numbers of specimens were tested, which made the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
data
Provided by rather
IHS Markit reliable.
under license For example, in the 1944 experiments
with WILEY Licensee=McDermott of Kotanchik User=G,
Inc - India/8215328006, et al.Boopathi
[12.45] 247
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
908 Stiffened Plates

Figure 12.1 Local buckling of a transversely stiffened panel under axial compressions: The lightly
loaded dorsal fin of the Saab 2000 commuter aircraft has a thin skin that is allowed to
buckle under load between the :;tiffening ribs-photographed during a test flight when a
rudder deflection induced compressive loads in the left side panel of the dorsal fin (cour-
tesy of Saab Aircraft AB)

panels with Z-stiffeners and 304 panels with hat-section stitleners were tested, or Schuette
eta!. in 1946 [12.46] tested 64 panels w:ith hat-section stiffeners.
Simultaneously, similar studies were carried out by Cox and his co-workers at the National
Physical Laboratory (NPL) in England. They performed one of the earlier series of tests on
the buckling of stringer-stiffened panels with emphasis on the load carrying capacity of the
whole panel [12.3]. This report begins with an interesting discussion on design philosophies
of efficient stiffened panels for wing structures, worth reading even today. Seven aluminum-
alloy panels stiffened by three to five lipped Z-section stringers were fabricated by a British
aircraft company as examples of the then new design concept of thick wing covers, stiffened
by fairly closely spaced, relatively light stringers. The stringer area ratios were (AJ bt) =
0.104-0.196 and the stringer spacings (spans of the plate between stringers divided by the
plate thickness) were (bit) = 40-60. Except for the two specimens with the larger stringer

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Figure 12.2 General instability failure of a 75 S-T6 aluminum alloy fiat panel with closely spaced Z-
stiffeners tested under compression at NACA Langley Research Center in the forties (from
Copyright Wiley
Provided by IHS Markit under license[12.47])
with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Plates Subjected to Axial Compression 909

spacing, (bl t) = 49 and 62, local buckling did not significantly precede overall buckling
(general instability), which involved considerable torsion and bending of the stringers. Since
in the forties electrical resistance strain gages were in their infancy and NPL was among
their developers, their use and accuracy were discussed in the report at great length.
Of special interest, and also to future experimenters, is the method employed to ensure
uniform loading, which is therefore discussed in detail. Because the panels were wide and
short, 20.5 in. X 12.5 in. ( = 520 mm X 318 mm), uniform distribution across the width
was important. But though the testing machine available had a very rigid wide lower platen,
the upper platen was only a disc of 6 in. (152 mm) diameter, "so that a special loading rig
had to be interposed between the top of the panel under test and this upper platen."
The test arrangement of a typical panel is shown in Figure 12.3. The ends of the test
panels were cast in low-melting-point alloy (1) to a depth of about I in. (25 mm), and the
ends were then ground flat and parallel. The base of the panel rested on a flat plate (2) 0.75
in. (19 mm) thick, which was placed onto the lower platen of the testing machine (3). This
flat plate carried two pairs of vertical pillars (4), along which rectangular flat plates (5) were
spaced by rubber washers (6). The edges of the test panel fitted into slots (7) cut in the
edges of these flat plates, which served as supporting combs that prevented the edges from

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
buckling, except between the closely spaced combs.
Though edge buckling was prevented, the boundary conditions represented by the sup-
porting combs were not completely defined. Since the fit in the slots would have to be quite
tight to prevent out-of-plane displacements, this probably introduced some frictional re-
straints on the in-plane displacements near the critical loads.
Across the top of the panel another cover plate (8) was laid, which slid over the ends of
the vertical side pillars (4). Extra guiding plates (9) were added above and below the top
cover plate (8), where it fitted over the pillars (4), to provide it with additional restraint
against rotation about a horizontal axis parallel to the test panel width.
On top of the top cover plate was laid a steel beam ( 10) of length equal to that of the
test panel and of !-section 5 in. (12.7 mm) deep by 4.5 in. (11.4 mm) wide. The load was
applied via two l-in. (25.4-mm) diameter rollers (11 ), and these rollers were loaded through

Figure 12.3 National Physical Laboratory, England, early test rig for buckling of closely stringer stiff-
ened llat panels (from information in [ 12.3]): (1) low-melting alloy cast end, (2) bottom
plate, (3) lower platen of testing machine, (4) vertical support pillar, (5) supporting comb,
(6) rubber washer, (7) slot for plate in supporting comb, (8) top cover plate, (9) guiding
plates, (10) steel !-section beam, (11) roller, (12) additional !-section beam, (13) top roller,
Copyright Wiley (14) top platen of testing machine
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
910 Stiffened Plates

another !-section beam (12), which itself was loaded at its center through a third transverse
roller (13) bearing against the top platen of the testing machine (14). During assembly, two
jigs were used to locate the three rollers in their correct position. The optimum spacing (2a)
of the two rollers ( 11) for the most uniform lateral distribution of the applied load was
calculated for each test panel.
This fairly elaborate rig was necessary to ensure uniform lateral load distribution. Unfor-
tunately, while very effective for uniform distribution of load laterally, the pyramidal con-
struction of the rig "rendered it far from stiff to rotation about an axis parallel to the width
of the panel." The extra guiding plates (9), which were fitted after one panel had failed
prematurely, improved matters, but it was still necessary in setting up a test panel to adjust
the position of the top loading beams so that the stringers and plate were loaded uniformly.
Furthermore, when, as a result of initial buckling, the stiffness of the plate or stringers tended
to be reduced, the load was still partially free, by rotation, to follow up the weakened
component and thus overload it.
Note that though this effect, when apparent in some of the tests, was here found not to
be important, it should be watched out for in other test setups that use similar loading
pyramids!

12.3.3 University College, London, Small-Scale Tests


As mentioned in Section 12.2.3, a series of experiments on axially compressed stiffened
plates was initiated about three decades later at University College, London [5.64], [12.42]
and [12.49]-[12.51]), that represent typical present-day small-scale testing concepts. In the
intervening years, the experimental studies on buckling of stitiened plates were mainly con-
fined to a few large-scale tests related to specific structures (see for example [12.52] and
[12.53]). The researchers at University College, however, believed that small-scale mode
tests would provide a means of exploring the behavior of new structural configurations and
investigating the interactive buckling of eccentrically stiffened plates. The more economical
and versatile small-scale experiments would permit extensive parametric studies, clarify the
phenomena and indicate problem areas. At a later stage the many small-scale tests would of
course have to be supplemented by large--scale tests, but these would then be limited to the
critical domains.
They therefore set out to develop small-scale models that would correctly reproduce the
relevant characteristics of the full-scale structure. Initially they focused on elastic buckling
(see [5.64] and [12.42]) and therefore looked for a model material that would remain elastic
up to large strains and thus avoid yielding. As mentioned in Chapter 5, they made their
models from an epoxy plastic, Araldite 219, that had a very high elastic strain capacity
(maximum elongation of up to 5 percent) and thus permitted repeated testing well into the
postbuckling regime without permanent deformations. The Araldite, which has also been
used in many other structural models, comes in liquid form, and sheets were made by casting
it between glass plates separated by metal shims. The stiffeners were machined from an
Araldite sheet and then bonded to the plate with Araldite as adhesive to form the stiffened
panel. By the use of the same material for components and adhesive, the cured models were
practically homogeneous and free of residual stresses. The method of fabrication of the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

models permitted close control of their geometry and was also inexpensive, while their low
modulus allowed testing in a simple hand-powered loading rig.
The rig is shown in Figure 12.4. It consisted of fixed and movable loading beams, both
machined into knife edges that apply the load. The movable beam was mounted on a pair
of linear ball-bushings to facilitate sliding. Two screw-jacks were connected to the movable
beam to draw it towards the fixed beam and thus compress the panel placed between them.
To protect the loaded edges, the stiffened panels were mounted in steel shoes with V-grooves
machined at their back. The knife edges of the beams placed in these V-grooves provided
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Plates Subjected to Axial Compression 911

Figure 12.4 University College. London. simple hand-operated test tig for buckling of stiffened plates
(from ( I2.421)

approximate simple supporL~. A number of parallel V-groovcs were machined in each shoe
to allow eccentric load application (loading ofT-set from the panel centroid).
The overall load ing concept of thi s si mple test rig for stiffened plates was si milar to that
of the earli er University College, London. test setup fo r unstiffened plates discussed in Chap-
ter 8 and shown in Fig ure 8.16. Here, however, the more sophisticated needle bearings,
simulating simpl e supports at the loaded edges. used there were abandoned and replaced by
a simpler "knife edges in V-grooves" approximation to simple supports in order to facilitate
the variable eccentric loading (with respect to the centroidal plane). of importance in the
testing of sti ffened plates.
By changing the line of application of the compressive loading (its offset from the plane
of the neutral axis), either by placing the knife edges in a different V-groove o r by adding
shims under the panel in the steel shoe. the overall imperfections of the st iffened panels
cou ld be controlled. One may recall that the change in load offset is equivalent to initial
curvature in the column-type overall buc kling.
Local plate imperfections were introduced by compressing the panel with a low compres-
sive load and applying steady moderate heat (from a photography floodlamp). The Araldite
softened slightly and after some time local buckles appeared in the plate due to creep, wh ich
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

became permanent imperfections when the resin cooled under constant load. These perma-
nent local deformations had the geometric fo tm corresponding to the local buckl ing mode
and therefore presented the most severe type of local imperfections.
The out-of-plane deformations, inc luding initial imperfectio ns. were measured by a pair
of LVDT's mounted on a gantry that could slide along the length of the pane l. The outputs
from these transducers were recorded o n an X- Y plotter after separation of local and overall
deflection shapes. The axial loads applied to the panel were measured by strain gages bo nded
to the screw jacks and therefore included any inadvertent load sharing due to friction in the
test rig.
The stiffened plate configuration invest igated is shown in Figure 12.5, together wi th the
relevant buckling modes. The panels consisted of a thin plate with a number of rectangular
sti ffeners auached to one side. As mentioned above. the loaded edges were simple suppo rted,
whi le the longitudinal edges were free. The overall buckling (general instability) mode was
therefore essentially that of a broad Euler column. The University College investigation
focused on the interaction between this Euler mode and the local plate buckling mode, as
well as with the third mode, a mode of local torsiona l buckling of the sti ffeners, important
primarily when they are thin. As shown in Figure 12.5a, the test specimens included two
stiffener configurations. a stocky one. where the interaction occurred between the local plate
mode and the overall Euler mode, and a thin o ne, where local torsional buckli ng of the
stiffener appeared as a th ird mode (see [5.64] and 112.421. respectively).
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
912 Stiffened Plates

Slillener Oi~ensiols in jm
conliguralion h b il' d
.. Stock( 0.75/57.5 4.90 I 10.0
,.lhin" 0.75 46.5 1 0. 75 14.4
(a)

I~
I
I
I
I

:c) (d)

Figure 12.5 University College, London, Araldite stiffened test panel (from [12.42]): (a) panel ge-
ometry and dimensions, (b) Euler general instability mode, (c) local plate buckling mode,
(d) local stiffener torsional buckling mode

Since the local buckling stress ucr of the panels was determined by the plate thickness
and the spacing and torsional rigidity of the stiffeners, while the overall (Euler) buckling
stress ur; depended on the overall flexural stiffness of the panels and their length, it was
possible, by successively reducing this length, to vary systematically the ratio (a,,J u£) with-
out changing any of the other dimensions, thus permitting direct comparisons. The specimens
were therefore manufactured with a length of 400 mm, for which (Pj P £) > 1. After ex-
haustive testing up to maximum carrying capacity, with two local imperfection amplitudes
(w1/t) and various overall imperfections (W0 /t), represented by the load offset (load eccen-
tricity), the panel was shortened by 20 mm and the tests were repeated. This process was
continued till the length was about 200 mm and (PjPF:) < 1. The many repeated tests were
feasible since the Araldite models remained entirely in their elastic domain.
The tests yielded experimental impeifection-sensitivity plots, some of which are shown in
Figures 12.6 and 12.7. As pointed out in Section 12.2.3. these 1974 experimental results
have since been used to verify new analyses and the theoretical results of one of these [ 12.41]
appear in Figures 12.6 and 12.7 and exhibit good agreement. In the experimental imperfec-
tion-sensitivity plots for plates with stocky stiffeners, shown in Figure 12.6 for three values
of the critical load ratio Per! P E = 1.50, 1.02 and 0.64, PM represents the observed maximum
loads for different equivalent overall (Euler) imperfection amplitudes W0 (which were ob-
tained by different load offsets). The plots present results for two local plate imperfection
amplitudes w1/t = 0.01 and 0.14.
Since in many applications structural efficiency would dictate thinner and deeper stiffeners
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

than the stocky ones tested, some panels with thin outstand stiffeners were also tested. This
introduced a third mode of buckling associated with torsional buckling of the thin outstands
(see Figure 12.5d). The results indicated rhat the coupling between this torsional mode and
the overall Euler mode could be very severe. A typical experimental imperfection-sensitivity
plot for this case is presented in Figure 12.7, again showing a comparison with the more
recent predictions of Sridharan and Peng [ 12.41]. There were two local to Euler mode ratios

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Plates Subjected to Axial Compression 913

10

0.9
lol
\,,~ i ;;_,I P,. , 1.50

l-.-~ ~
r>~~ '1---
I ----

o w0 ft• 0.01 i
0.6
- , W0 it'l'" '
0.5
0 f f f
0.5 05 Waft 1.0 15 2.0
I0

0.9
)b)
\~ I ~,IPE '1.02

''
0.8 '• F\_ I
r<~--~
..... _--d-o.~-
I
0 Wo/t; 001
0.6 f- •Waft ~ 0.14
I
f f
0 0.5Woit 10 1.5 2.0

lol ~'I Pe, 0.64


0.9
I W 0 /t : 0.01
f\..o I I

06

05
9o.s
' 0 1.5 2.0
0.5 Wait 1.0

Figure 12.6 University College, London, experimental imperfection-sensitivity plots for three panels
with stocky stiffeners, versus overall Euler mode imperfection amplitude W,,l t, and
comparison with more recent theoretical predictions (from [12.41]): (a) L = 372 mm,
PjP, = 1.50; (b) L = 307 mm, P,,/Pr: = 1.02; (c) L = 243 mm, P,,IP, = 0.64; con-
tinuous and dashed lines represent theoretical results for local imperfection amplitudes
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

w0 /t = 0.01 and w,,ft = 0.14, respectively

1.0 I .1_
{,1Pe•l.05
0.9
o~-'G,

-
o-' ~

. · / ·.
0 "<..,
__ ... , .. c' !Ca.:.: p k-- --
....o 0
- v
--~- ~ local mode imperfecroora_~
0.6
~ v ow0 /t•002
• w 0 /t • 0.10
I ~
0.5
f f f f j' f
'21.50 -1.00 -0.50 0 0.50 1.00 1.50
Wo/t
Figure 12.7 University College, London, experimental imperfection-sensitivity plots for a panel with
thin stiffeners, versus overall Euler mode imperfection amplitude and comparison with
more recent theoretical predictions (from [12.41]): L = 395 mm, PjP, = 2.20 for local
stiffener mode and Pj PE = 1.05 for local plate buckling mode; the dashed and continuous
lines represent theoretical results for local imperfection amplitudes w 0 / t = 0.02 and 0.10,
respectively

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
914 Stiffened Plates

in this case: Pj P10 = 2.20 for the local s1~iffener mode and Pj P10 = 1.05 for the local plate
mode. The local mode imperfections here were actually a mixture of plate and stiffener
buckling modes, but only the plate imperfection magnitudes (w 0 /t = 0.2 and w 0 /t = 0.10)
were measured and presented in the figure.
These University College. London, experiments stand out as unusual carefully controlled
tests of stiffened panels, with an unusual emphasis on the experimental determination of the
imperfection sensitivity, a task usually relegated to analytical treatments.
The UC London researchers then turned to the manufacture and testing of small-scale
steel stiffened panels [12.49]-[12.51], focusing on the correct production of practical levels
of welding stresses and out-of-plane imperfections. They had concluded from dimensional
analysis (as discussed in Chapter 5) that, provided the material and boundary conditions of
model and prototype were similar, linear scaling of the geometries would ensure similar
behavior. Hence the steel used in the smalll-scale models was specially prepared to correspond
to a grade often employed in British structural practice, and each model was manufactured
from a single sheet. They also noted that small-scale models had a real advantage over their
large-scale brethren in the better consistency of material properties throughout the test spec-
imens. Their "models were fabricated from individual stiffener and plate components, which
were welded together using a tungsten inert gas welding (TIG) process, with argon as the
inert gas. To reduce the level of welding stresses the plate and stiffener elements were
clamped to copper heat sinks." The weld-induced stresses and out-of-plane imperfections
were carefully measured and found to be of the same levels as those measured in full-scale
steel stiffened panels. Both single-bay and double-bay stiffened panels were manufactured
and tested.
All the specimens were tested in uniaxial compression on a Denison universal testing
machine equipped with auxiliary instrumentation that permitted direct plots of the load-end-
shortening while the test was in progress. The rate of loading was generally slow, not ex-
ceeding 10 kN per minute.
The ends of the panels were cast into blocks of a low-melting point alloy. Cenobend, and
then machined fiat, parallel and perpendicular to the longitudinal axis. These blocks were
then placed into loading shoes. "The single-bay simply-supported panels were loaded in a
rig, which consisted of two (vertical) pillars (1 ), with grooves accurately machined to house
the ball-bearings (2), attached to the loading shoes." The ball bearings (see Figure 12.8c)
facilitated the sliding (or rolling) of the loading shoes (3) and "ensured continual alignment
of the load as the panel deflected in the overall mode. Note that since the overall mode is
a wide column mode, the supports were similar to the column supports discussed in Chapter
6. The load was applied through a circular bar, which in e±Iect acted very much like a knife-
edge."
The same rig was used for the two-bay tests. except for the addition of "a cross-bar which
had ball-bearings that fitted into the same grooves as the loading shoes. again ensuring
alignment. The bearings were almost entirely free from friction so that the cross-bar provided
a rigid support against lateral deflection but no restrain against rotation," thus closely ap-
proximating simple support boundary conditions.
As the panels were placed in rather rigid end blocks, the influence of their rigidity was
investigated and found to be usually less than 2 percent. Similar results were earlier obtained
for columns by Chilver [6.16] and discussed in Chapter 6, Section 6.1.5.
Since most of the design information for stiffened panels available at the time was based
on single-bay tests, the UC London tesr program on small-scale steel plates [12.50] and
[12.51] focused on a comparison between single-span and multi-span panels. Their experi-
mental results indeed confirmed the predicted change in mode with the increase in the
number of bays.
For validation of the main premise of the University College researchers, that very small-
scale modeling can reproduce the behavior of full-scale stiffened panels, they compared their
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Plates Subjected to Axial Compression 915

results with those of si milar quarter full -~cale specimens, tested at the Univcr>t ty of Man-
chester [ 12. 14) in the seventies and di scussed below, and found excel lent correspondence.

12.3.4 Manchester University Experiments


In the mid-seventie' an extensive experimental study of the Mrcngth of stiffened plates of
the type used in steel box girders wa~ carried o ut at the Simon Engineering Lahoratories of
the University of Manchester (see [1 2. 141 and [ 12.54] and the reports referred to there). A
total of 86 stiffened panels was tested. 52 panels devoted to an investigatio n of the plate
fai lu re mode. 26 to a study of the sti ffener failure mode ;md & to an exam ination of the
eiTect of various straightening procedures. A~ before. the terms plate jai111re and stiffener
j11ilure indicate the component that buckle' first and trigger; the failure of the stiffened panel.
plate failure occurring when the stiffener.. are stocky and stiffener failure "hen they arc
sufficiently slender for torsional buckling to take place.
The sizes of the panels tested were determined "by the scale required to reproduce real-
i ~tical l y the e ffects of weldi ng and initial imperfections and to observe satisfacto ri ly the post-
buckl ing behavior." The specimens were plates of thickness 6.5 or 9.5 mm . compared wi th
prJctical full-scale thicknesses of. say. 10- 25 mm. or in other words one-quarter to one-half
~ale. The corresponding University College. London. ~mall-scale specimens. ui\cussed
above. were 1112-1130 scale. and The London reo;earcher\ claimed that the prJctical full-
scale welding and imperfection en·ccts and postbuckling behavior were also satisfactorily
~i mulated in their very small panels. They supported their claim by compari,on with the
larger-scale Manchester tests, which showed very good agreement of results. The Manchester

(b)

(a) (C)

Fij!ure 12.8 Unher..lly College. London. 'maii-\Cale muhi-ba) >teel stiffened panelte>t,-test rig and
buckled 'pecimens (courte,y or Proressor A.C. Walker and rrom inrormauon tn [ 12.51 J):
(a) tC\1 ng. (b) detail or lower loading shoe with buckled specimen. (c) 'cction through
loading 'hoc (where I is vertical pillar. 2 a ball bearing rolling in a grove and 3 a loading
shoe). (d) typical postbuckling pauc.-n of an axially compressed specimen (the coin indi-
cating it' "nallness) showing the postbuckling dcftmnation of the sti iTencl"\
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
916
Experiments on Stiffened Plates Subjected to Axial Compression 917

specimens, however, probably reproduced the actual fabrication techniques of the full-scale
structures more faithfully, and had the advantage of the larger test specimens.
A high-capacity rigid test rig was built specially for this series of tests (see Figures 12.9
and 12.10). Designed by Professor L.K. Stevens of the University of Melbourne, it was
capable of applying a maximum load of 10,000 kN to stiffened plate specimens up to 1.8
m wide and 3.3 m long. The length could be extended to 9 m, the limit imposed by the
laboratory strong floor on which it was mounted. The rig depended on the strong floor of
the laboratory for vertical and horizontal stiffness, but the main horizontal force was taken
by eight tension straps, two pairs on each side (B in Figure 12.9).
These straps are connected by high strength friction grip bolts at the anchor (fixed) end A to a
heavy cross-plate (C) and at the free end F to a second cross-plate (D). While the cross-plate C is
fixed in position to a braced frame (E) connecting it to the strong floor, the plate D is free to slide
in a horizontal plane between the two pairs of horizontal steel beams (G). The beams G are
themselves fixed through columns (I) to the strong floor [of the laboratory].
The free end D takes the reaction from the stiffened panel under test (J). The other end of the
test panel is fixed to a yoke (K), which again slides between two pairs of horizontal members (L),
fixed through columns (M) to the strong floor. However, the yoke K is free to slide only in a
direction parallel to the longitudinal axis of the test rig, as it is restrained against horizontal lateral
motion by a long link arm (N) pivoted to the yoke at one end and to a rigid support (0) at the
other [12.14].

The longitudinal load can be applied to the yoke in one of two ways: either as a constant
load, by operating an electric pump under constant pressure, or as a constant strain, by using
the two hand-operated pumps connected to the two 5000 kN hydraulic jacks (Q) to give
controlled longitudinal deformation. "The electric and manual system are connected in par-

Figure 12.9 Manchester University stiffened plate test rig (from [12.14]): (A) anchor end, (B) tension
straps, (C) heavy anchor end cross-plate, (D) free end (reaction) cross-plate, (£)braced
anchor end frame, (F) fi·ee end, (G) horizontal steel guide beams between which D can
slide, (!) vertical support columns, fixing beams G to laboratory strong floor, (1) stiffened
panel under test, (K) yoke, free to move longitudinally but not laterally, (L) horizontal
guide beams between which the yoke K slides, (M) vertical support columns, fixing beams
L to laboratory strong floor, (N) link arm, restraint against horizontal lateral movement
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
forK, (0) rigid support of link arm end. (P) load cells, (Q) hydraulic jacks (2 of 5000
kn, motorized or manual), (R) cylindrical end bearings of jacks
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
918 Stiffened Plates

figure 12..10 Manchester Unl\cl'\ity stiffened plate test rig- ,iew from top with \pe<:tmen no. 9 in
position (coune>y of Professor R. Nara)anan)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

allel so that it is possible to apply the loads initially through the electrically operated pump
and. later on, switch over to manual control."
The free end cmss-platc D was allowed to float l;ucrally in o rder to avoid unknown shear
forces being applied to the plate. Thus the stiffened test panel was loaded o nly by a known
overall longitudinal load.
""The jack load is measured by two load cells P which react against the anchored cross-
plate C." The strains in the load cells were measured by four strain gages attached to each
of them and were recorded by a data logger. "The jacks have cylindrical end bearings (R)
which will ensure central loading of the jacks. although some accidental eccentricity is
allowable because the load cells have been designed to he accurate to 0.05'lf with a maximum
misal ignment of 25 nun in 5,000 kN. As the yoke K is free to rotate about a venical axis,
the jack loads can be applied unequally. thus perm itting the a pplication of a lateral stress
gradient to the test panel'' (12.14].
The system of high-capacity hydraulic jacks used n large volume of oi l. which increased
the overall llexibility of the test. In order to minimize the volume of oi l under pressure and
thus the reduction in rigidity. the tests were staned with a jack back clearance as small as
possible (less than 10 mm). Thi, was achieved by adjusting the position or the free cross-
plate (reaction plate) D on the side straps (8). and by introducing a packing piece between
the cross-plate D and test ,pccimen. The higher rigidity increased the chance' of ohservi ng
the unloading of the specimen after it had reached its maximum load. lndecd. for plate
failure cases the unloading versus axial shortening rela tionship could be observed, hut not
for the sti ffener failure cases.
A typical test specimen is shown in Figure 12. 11. The test panels had three bays and were
1523 mm (60 in.) wide. with slightly longer end plates. about 1625 mm (6-1 in.) wide. Their
plate thickness was 6.5-10 mm and three sizes or longitudinal Hat outstand stiffeners were
used with each plate thickness: 150 mm x 16 mm. 150 mm x 9.5 mm and 80 mm x 12
mm. The stiffeners were 457 mm (18 in.) apan so that the corresponding (b/t) values were
70-48.
The panels were of two lengths. 9 15 mm (36 in.) and 1830 mm (72 in.). The short ones
were tested with nomina l fixed ends. o btained by bolt ing the end plates of the specimens
d irectly to correspondi ng J)lates o n the free-end (react ion) cross-plate (Din Figure 12.9) and
the floati ng cross-plate (the yoke, K in the figure). When necessary, shim~ were inserted

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Plates Subjected to Axial Compression 919

r---------_1523 mm (60'_')__ -------1


I 457m~, 1

g;_:;~.-L~±~_t=-~~·-_=:~~~n=d
end piate
I '

ia"( I
~T


9.5mm( 3
'
fillet welds

length(12")
I
I weld
length (4")
6-5 mm (1/4")
fillet welds
/
I [mm("')
fillet welds
I

l
I
I
~.

continuous mterm1ttent
welding welding

Figure 12.11 Manchester University stiffened plate tests-typical test specimen (from [12.54]): length
L = 1830 mm (72 in.) or 915 mm (36 in.), plate thickness t = 9.5 mm (3/8 in.) or 6.5
mm (1/4 in.), stiffener sizes d, = 150 mm (5.9 in.) or 80 mm (3.15 in.) and t, = 16
mm (5/8 in.), 12 mm (4.7 in.) and 9.5 mm (3/8 in.)

between the end plates to avoid the introduction of transverse bending moments into the test
specimens.
The longer specimens were tested with nominal pin-ends (simple supports). These were
obtained by attaching two 38-mm diameter pins of hardened steel, which extended over the
entire width of the specimen, to a pair of face plates, fixed to the faces of the movable yoke
(Kin Figure 12.9) and the reaction cross-plate (D in Figure 12.9) along the centerlines of
the faceplates. These two pins were carefully aligned and checked for parallelism. The load
was applied through the pins to a pair of faceplates, with hardened steel bearing surfaces
built into them, which were bolted to the end plates of the specimen.
This specimen assembly was then rested on four corner screws, which permitted adjust-
ment of its position vertically to eliminate any transverse bending stresses as measured by
the strain gages attached to both sides of the plate and stiffeners. Any desired eccentricity
of loading was later measured from this (thus obtained) neutral loading plane.
The actual realization of these nominal pinned and fixed ends by the simple boundary
conditions employed was not reported by the Manchester researchers. Today, this could be
conveniently measured by vibration correlation techniques, as discussed in Chapter 15. Be-
cause the unloaded edges are free, the overall buckling modes of the stiffened panels are
those of wide columns, which makes the panels very amenable to vibration correlation.
The test setup was well instrumented.
Residual strains were measured by drilling 1 mm dia. holes 1 mm deep, at 50 mm or I 00 mm
apart in plate panels and 25 mm in stiffeners, and using a Demec gauge. These measurements were
made on the plate and on the stiffeners along the centre of each representative panel.

Axial shortening of the specimen and the vertical displacement of the plate and the stiffener at
various points were measured using transducer gauges and read by the data logger.
Three scanners 457 mm, 915 mm and 1830 mm long were used to measure the plate imperfec-
tions, the overall imperfections of stiffeners normal to plate and the imperfections of stiffener
outstand parallel to plate .... The ripples formed along the plate and along a stiffener due to the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
920 Stiffened Plates

application of load were observed and plotted using a ripple scanner and an X- Y plotter. The scanner
consisted of a motorized carriage moving on an aluminium beam over 3 m long which could be
fixed at any position over the panel on the test rig. Two transducer gauges were mounted to this
carriage at right angles to each other: one to measure the plate deflexions and the other to measure
the torsional deflexions of the stiffener. The output from these gauges was fed directly to an X-Y
plotter and the deflected shapes of the plate and of the stiffener were directly plotted.

As mentioned earlier, the size of the panels sufficed to reproduce realistically the fabri-
cation techniques of full-scale structures, and therefore the specimens were suitable for in-
vestigation of different fabrication effects in addition to the main test program. Examples of
these were a study of the effects on buckling and collapse of residual welding stresses due
to continuous versus intermittent welding, of geometric imperfections and of straightening
procedures.
The typical continuous vvelding and intermittent welding employed in different specimens
are shown on the two sides of the panel in Figure 12.11. Measured mean residual stresses
induced in the plate panels continuously welded to the stiffeners varied between 90 and 150
N/mm 2 , compared with values deduced from the Merrison rules [12.55] between 42 and 68
N/mm 2 • The corresponding figures for intermittent welding were 20 to 40 N/mm 2 measured,
compared with 13 to 21 N I mm 2 • Hence the Merrison rules underestimated the welding
residual stresses by a factor of the order of 2.
The comparison of continuous welding with intermittent welding indicated that the higher
residual stresses due to continuous welding did not have any large consistent effect on the
collapse loads of the panels, although there appeared to be some benefit from the lower
residual stresses induced by intermittent welding for the panels of higher slenderness. The
absence of any benefit from the lower residual stresses in intermittently welded panels of
low ,slenderness was apparently due to separation between plate and stiffeners as the failure
load was approached. Additional tests on panels with various forms of intermittent welding
also showed that, even with very short miss-lengths, intermittent welding seemed to have no
real beneficial effect for short panels.
"The post-buckling behaviour of the intermittently welded panels showed a more rapid
fall-off of load with axial shortening than was the case for continuously welded panels,
[whereas] the 'softer' behaviour of the continuously welded panel was an advantage in al-
lowing a greater degree of equalization of longitudinal stress at failure in stiffened boxes"
(see [12.54]). Hence, in spite of its signilicantly lower residual stresses, intermittent welding
appeared to have no noticeable advantage and even revealed a lower ductility.
The effect of geometric initial imperfections of three types were investigated: of plate
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

panel imperfections, local torsional imperfections and overall panel imperfections. For all
three types, imperfections of large magmtudes were artificially imparted to the specimens.
For plate failure cases the imperfections
were induced by using a specially built dishing rig. The rig was fitted with three hydraulic jacks
(of 10 ton capacity) which could be fixed in any desired position over the plate. The specimen was
positioned in the rig and dish-shaped bearing plates were located at places where imperfections
were to be induced. Dial gauges were fixed such that the deflexion of the plate under the load
could be measured. The jacks were carefully operated so that the plate was strained a little beyond
the yield point and the desired degree of imperfection induced [somewhat similar to the technique
used for unstiffened plates at Monash University, shown in Figure 8.54, though without recourse
to heating]. The procedure was repeated so that a wavy imperfection (square chequer-board pattern)
similar to the expected elastic critical mode was induced.

For stiffener failure cases the stiffeners were crimped so that they would have a wavy
imperfection at the outstand.
The crimping tool consisted of a seating plate with two split conical supports and a loading plate
with its loading surface having the shape of the desired imperfection. The stiffener was supported
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Plates Subjected to Axial Compression 921

by the seating plate and a hydraulic jack (100 kN capacity) was used to apply a carefully controlled
load on the outstand through the loading plate. When the stiffener tip was loaded just beyond the
yield point, there was permanent set and an imperfection was induced at the outstand. A dial gauge
was fixed behind the point of application of the load so that the extent of induced imperfections
could be carefully controlled. This procedure was repeated by shifting the crimping tool to the
different positions on the stiffeners so that an imperfection with a wavelength similar to that of the
expected mode of failure was induced in all the stiffeners.

Overall panel imperfections were imposed by eccentric application of the end loads, the
load eccentricity representing also the lack of straightness.
In all three cases, the magnitude of the imposed imperfections was three or more times
the tolerance prescribed by the Merrison rules [ 12.55], and of the most severe shape, that of
the respective buckling mode. The effects on the collapse load of the panels were, however,
found to be not very significant. For plate imperfections, the threefold increase caused only
a small drop in collapse load., though for intermittently welded panels the detrimental effect
of the imperfections was larger. For local (torsional) stiffener imperfections the threefold
increase resulted in an 8 percent reduction in collapse load. Only for overall imperfections
did the fivefold increase produce significant decreases in collapse load, of about 25 percent.
Except for overall imperfections and eccentric loading, the axially compressed stiffened fiat
panels of the geometries tested (remote from the perfect optimum) appear therefore not to
be imperfection sensitive.
A special investigation was carried out on the effect of straightening procedures on the
strength of stiffened panels (see [12.54] and [12.56]).
In order to overcome imperfections introduced by shrinkage of welded joints or otherwise, various
procedures are in common use by fabricators. These may be categorized as follows [12.54]:

(a) Localized heating, to a sufficient temperature to cause plastic upsetting, followed by cooling
and consequent contraction which pulls the panel back to the required shape.
(b) Clamping of parts during welding to initial curvatures and choice of welding sequences such
that, on release after welding, the stiffened panel springs back to the required shape.
(c) Mechanical loading (stretching or bending) causing some degree of plastic deformation such
that, on release of the applied load, the required shape is obtained.

To study the influence of these straightening procedures, eight panels were tested, which
were of two types, with fiat outstand stiffeners or with T-stiffeners, but all designed to
collapse by stiffener failure. The details are discussed in [12.54] and [12.56], the general
conclusions being that the decrease in strength of the panel was correlated to the sense and
magnitude of the residual stresses induced in the stiffeners during the straightening proce-
dure. When these stresses were compressive they caused some reduction in the strength,
whereas when they were tensile they were not detrimental.
To conclude this subsection, the carefully designed Manchester University tests, though
performed two decades ago, still represent the state of the art, except for some updating of
instrumentation and computerization of data acquisition and reduction.

12.3.5 Welded Steel Ship Grillage Tests


Plated grillages occur frequently in deck and hull structures of ships. In the fifties and sixties
welding replaced riveting in ship construction and longitudinal framing was adopted in most
large ship hulls. This focused attention on the strength of welded plate grillages, with closely
spaced longitudinal girders and wide space frames under longitudinal compression, and mo-
tivated the extensive theoretical and experimental studies carried out by naval architects well
into the eighties.
Among the earlier investigations should be mentioned the extensive series of compression
tests on single-bay welded steel uniaxially stiffened panels carried out in the late sixties by
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
922 Stiffened Plates

Faulkner and his students at the Royal Naval College, Greenwich (see [12.13]). In this
program 65 models, representing at approximately 116-114 scale a range of geometries of
interest to ship designers, were tested to failure. The test series included 13 nominal identical
models to examine statistical aspects, 9 models to examine residual welding stress effects,
19 student tests (the geometry of which covered only a narrow range) and 24 models of
varying practical and optimal geometries to provide design data.
All the models except two had five stiffeners each to ensure a minimum of edge effects.
The unsupported edge widths of plating were chosen to be b/3, where h is the stiffener
spacing, to give a buckling stress slightly above that of the simply supported plate between
the stiffeners. Most models were stiffened with T-bars and some with fiat outstands. The
material was mild steel and, as far as possible, normal shipyard construction techniques were
used in the fabrication of the specimens, to provide realistic design data. The overall di-
mensions were length L = 348 to 1224 mm, width 4.67 b = 383-1069 mm, plate thickness
t = 2.5-4.9 mm, and stiffener depth d =' 23.6-63.5 mm.
The testing machine (Figure 12.12) "consisted of a self-reacting bolted frame with a
variable cross-beam position to accommodate differing length models .... A total load of
900 kN was provided by three jacks supplied from a MN ( ~ 100 ton) Amsler machine. The
central jack provided 500 kN and the other two 200 kN each. The hydraulic pumping ca-
pacity was low, so that the test rig effectively applied dead load with a low strain rate.
"The scantlings of the loading T-beams were heavy to ensure uniformity of load. This
was checked by load-deflection tests on three pre-calibrated load cells and by strain mea-
surement on the first models." The entire setup was intentionally rather rigid, to provide
well-defined boundary conditions.
"Special end box-fittings were designed to provide the desired pinned-end condition and
to allow freedom to adjust the plane of load action relative to the neutral plane. The two V
knife-edges [see Figure 12.13] were made from toughened HTS and examined periodically.
Only slight signs of wear were seen, insufficient to affect the pinned-end condition."
The end fittings of the panels (seen in Figure 12.13) were also rather rigid. The influence
of their rigidity on the effective length of the specimens was therefore taken into account
by Chilver's approach [6.16] discussed in Chapter 6, Section 6.1.5.
The instrumentation was kept fairly simple. During tests, measurements were confined to
magnetically mounted dial gages to record lateral deflection of the central stiffener as an

ADJUSTABLE
CROSS·BE1IM

HODEL

SECTION THROUGH XX

Figure 12.12 Royal Naval College, Greenwich, test setup for compression tests on stiffened plate panels
(from [12.13])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Plates Subjected to Axial Compression 923

Fiuure 12.1 3 Royal Naval College. Greenwich, comprc,sion t~sts on welded ccccJUrically Miffened
plate panels-a typic;tl failure. model F2, ' howing abo the end finings (from r 12.13 ])

indication of incipient failure and two dial gages to check unifonnity of crush. The load was
recorded on the jack and wa~ increased step by step until the model failed. The load and
mode of failure were recorded.
The two primary failure modes of the stiffened paneb considered and ohsened were
column- like failures of the stiffener-plate panel (essentially the plate failure mode of Section
12.3.4) and sideways tripping of the stiffeners about their line of attachment to the plates
(the stiffener fai lure mode of Section 12.3.4). These two primary failure modes arc shown
in Figures 12.15 and 12. 16. respectively (from tests in the ARE test rig depi cted in Figure
12.14), as well as in Figure 12.2 1, taken from another series of tests discussed below [ 12.17].
Both modes could and did involve compl icated interaction with plate buckling phenomena.
One of the special stiffened plate problems that worried Faulkner and defied ~ati;factory
experimental treatment was the effects of alignment of the loading plane relative to the initial
neutral axis. or. in other word~. the uncenaimies arriving from load eccentricitie\. Calcula-

LOA.O U.t. ... t


OISTAit!Jfl \.0"0

r" _j\ INTO l,.f t


;.c.JIV( .IACICS
,-LOAD CELLS
~~~"J~n, PAUl \I( .U.CKS -
.......... lllANt-fG;
~

'- 'r~·I..~:LUI&U A.El.t.NNG WA.l£R• '"-I..t:D7


IVAI Pl..t.tr:(iOTH OC)$) RUUE.R '"G
.l.

f"='
rl
w
T ~ JII ......,AT
P\,_.t.TFOIUol
II lh
/ L TF

~
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Fi~:urc 12.14 ARE (NCRE) Dumfcrmline tests on steel welded grillages-diagram of tc't l"ig, built in
their large testing frame. f<lr compressive loading or combined compre«ion load with
lateral pressure (from [8.148])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
924 Stiffened Plat es

F ig ure 12. 15 ARE (NCRE) Dumfennl ine tests o n steel welded g rillages-view of test rig with grillage
3b after collapse by interframe Oexural buckling (courtesy of Dr. C.S. Smith)

tions indicated that when the load was aligned between the initial neu tral axis and the plate,
as was done in some of the tests. appreciable compressive bending stresses cou ld develop
with increase of loadi ng. which could load to premature plate collapse and be the cause of
disagreements between prediction and test.
To somehow deal with the chang ing load eccent ricities in the 24 parametric design data
tests, the load plane was aligned midway between the initial ne utral ax is and fi nal one
(assumi ng a width of 40 ti mes the thickness of plat ing to be then still effective). Thi s gave
good agreement wi th predictions, but it still involved a strong clement of uncertainty.
A possible solution to the load eccentricity uncertainty appeared to be mu ltibay tests. like
those employed in the Admiralty Research Establish ment (formerly Naval Construction Re-
search Establishment) Dumfennl ine. Scotland, large-scale tesL' [8.1 48] detai led below, or in
some of the University Col lege. London. small-scale experiments discussed in Se-c tion 12.3.3
( I2.50] and ( I2.5 I]. Multi bay tests allow interaction between behavior in adjacent bays and
provide greater realism. if early failure of the end bays can be prevented. Of course, because
of the greater complexity and costs, fewer models can then be tested. There fo re ex periments
on single-bay specimens continue in parallel wi th mult ibay ones.
Nowadays, the load eccentricity uncertainty cou ld also be overcome by vibration corre-
lation techni ques (VCT), d iscussed in Chapter I5. which can assess the actual load eccen-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 12. 16 ARE (NCRE) Dumfennline tests on steel welded grillages-specimen Ia afler collapse
Copyright Wiley by lateral to rsional buckling of the longitudinal stiffeners ( fro m [8. 148])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Plates Subjected to Axial Compression 925

tricity present. The load eccentricity uncertainty is also a serious problem in stiffened shells,
where it provided one of the motivations for the development of VCT.
In the early seventies extensive tests on large-scale grillages, representing warship single-
bottom, and deck structures under compressive load, were carried out by Smith and his co-
workers at ARE Dumfermline [8.148], [12.57] and [12.58].
Compressive failure of longitudinally stiffened deck and single-bottom structures in war-
ship hulls usually takes the form of local, inelastic buckling of longitudinal girders and
attached plating. The dominant geometrical parameters are therefore:
1. the plate slenderness ratio (hit), or the plate slenderness parameter f3 = (bit)~,
encountered already as Eq. (8.8) in Chapter 8
2. the longitudinal girder slenderness ratio (a/ k), or the longitudinal girder slenderness
parameter
(12.1)
where
a = spacing of transverse frames
b = spacing of longitudinal girders
t = plating thickness
k = radius of gyration of longitudinal girders acting with assumed effective width of plating
a 0 = material yield strength
E = Young's modulus
A survey of deck and single-bottom designs in British warships operating in the seventies
indicated values of f3 = 1.0-4.5 and y = 0.15-0.9. The test grillages were designed in mild
steel with f3 and y falling within these ranges and thus to represent the then current British
design practice, which probably also represented that of other countries at the time.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

The 12 test specimens included four pairs of nominally identical structures representing
possible ship-bottom configurations, two grillages representing frigate strength decks, one
grillage corresponding to a light superstructure deck and one constructed specifically for
evaluation of weld-induced residual stresses. The overall dimensions of each grillage were
6.096 m (240 in.) long by 3.20 m ( 126 in.) wide, with plating thickness = 6.30-8.97 mm
(0.248-0.353 in.). The specimens had nondimensional grillage slenderness parameters f3 =
1.41-3.65 and y = 0.23-0.70, which covered their design ranges fairly well. Except for two
grillages with fabricated girders, all stiffeners were standard Admiralty tee bars. The speci-
mens had 6-13 longitudinal girders each, spaced at b = 254-610 mm (10-24 in.), and 5-
6 heavier transverse frames, spaced at a = 1219-1524 mm (48-60 in.).
[The] grillages were constructed in the ARE (NCRE) workshop, following as far as possible normal
shipyard fabrication and welding procedures. The plating of each structure was formed by two
strakes jointed by a longitudinal butt weld at (or close to) the grillage centerline. All stiffeners
were continuously welded to the plating. During fabrication. plates were initially tack-welded to-
gether and stiffeners tacked to the plating, typically by 3 in runs of weld at 18 in spacing. Welding
was then completed manually, normally by a single welder operating first on one side of each
stiffener and then the other, giving in effect a two-pass welded attachment. ... An attempt was
made to achieve reasonable uniformity of plate thickness, stiffener dimensions and yield strengths
by ordering twice the required quantity of steel and selecting plates and stiffeners for grillage
construction on the basis of measured dimensions and material properties.

Material properties were obtained from extensive standard tensile and compression tests.
The tests were carried out in the Large Testing Frame (LTF) at ARE (NCRE) in a specially
designed test rig shown in Figure 12.14.
Grillages were supported horizontally on a 20 ft X I 0 ft ( ~6 m X 3 m) steel platform and were
held down at their ends against lateral pressure by light flexure plates and along their sides by tie-
bars with 6 in. (152 mm) spacing. The tie-bars [seen clearly in Figure 12.151 incorporated bottle
Copyright Wiley
screws
Provided by IHS to allow
Markit under vertical
license with WILEY adjustment. This form of Licensee=McDermott
support was intended to impose
Inc - India/8215328006, conditions
User=G, Boopathi of
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
926 Stiffened Plates

zero vertical displacement and zero rotatiollal restraint (simple support) at the ends and sides of
test structures, with the ends and sides free to move longitudinally in the plane of the grillage and
the sides also free to move transversely in-plane.
The tie-bars were connected to the support platforms at their bottom ends and to the edges
of the test panels at their tops by pins located in loose-fit holes, which provided low friction
joints. Also, the flexural stiffness of the tie-bars was very small compared with that of the
transverse stiffeners. These two factors ensured a close approximation to simple support
conditions.
Longitudinal compression was provided by six (or in some cases seven) hydraulic jacks evenly
spaced across the ends of the grillages. Compression was applied by activating the jacks at one
end of the rig from a common pressure source and reacting loads through a passive set of jacks,
also connected to a common (closed) hydraulic line, at the other end. At the active end loads were
applied through calibrated hydraulic load 1:ransducers giving an indication of each applied jack
force with better than 2% accuracy. Differences between jack loads, attributable to friction in the
jacks, were always less than 10% and in most cases less than 5%. Differences between total loads
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

at active and passive ends, attributable to friction between grillages and the support platform and
to horizontal components of force in tie-bars and flexure-plates, were also normally less than 5%.
Jacks and load-transducers were mounted on cradles incorporating vertical screw adjustment to
provide accurate control over the line of actmn of compressive forces. As shown in [Figure 12.14],
jack loads were applied through ball-bearings in order to minimize rotational restrain at the ends
of test structures. Jack forces were reacted through an arrangement of load beams into the walls
of the LTF.
In order to distribute jack loads into the test structure in a reasonably uniform manner, avoiding
the risk of premature failure close to the ends, each grillage was reinforced by fitting double plates
to the webs and tables of longitudinal girders and to plate panels between girders over 2/3 of the
span of end bays. Because of the conditions of simple support at the ends, reinforcement in these
regions had negligible influence on overall behaviour of the grillages and proved a successful means
of ensuring that interframe collapse occurred in a representative region of each structure away from
the ends.
In some of the tests the compressive load was combined with lateral pressure.
[This] was applied by means of a water-filled rubber bag contained between test panel and support
platform as shown in [Figure 12.14]. Provision was made for lateral pressures of up to 30 psi
(206.9 X 103 Pa). Lateral loads were monitored by a calibrated pressure gauge and checked by a
water manometer.
Before testing, the initial defonnations of each grillage were thoroughly surveyed. The overall
vertical deformation of each grillage relative to its ends was measured at stiffener intersection
positions (and at intermediate positions along selected stringers) by means of stretched wires. Local,
interframe vertical deformations of longitudinal girders were also measured, together in some cases
with sideways distortion of stiffener tables. Extensive measurements were made of plating distortion
along the center-lines of plate panels between girders [using an array of dial gages mounted on a
portable bar].
In the light of subsequent research Smith considered these plate distortion measurements
inadequate [12.59]. He felt that, although less important than in shells, the effect of initial
geometric imperfections in flat plates was quite subtle and warranted more extensive mea-
surements and study. The authors concur with this view, which is also expounded in Chapter
10.
Weld-induced residual strains were measured during the construction of the grillages by
taking extensometer readings on both sides of the plating and stiffeners before and after
welding using a Demec mechanical extensometer. Measurements on the plating were usually
confined to longitudinal strains at the centerlines of plate panels. As a result of the unex-
pectedly high level of residual stresses recorded in all the specimens (except the first two,
in which residual stresses were not measured), it was decided to construct specially and test
two additional grillages for better evaluation of the weld-induced residual stresses.
Seven grillages were tested to collapse under compressive load alone, while four (the
Copyright Wiley
remaining
Provided by IHS mates
Markit under license of the four pairs) were subjected
with WILEY to compressive
Licensee=McDermott load combined
Inc - India/8215328006, with lateral
User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Plates Subjected to Axial Compression 927

pressure. The tests on the nominally identical pairs provided a direct indication of the influ-
ence of lateral pressure on compressive collapse.
Before commencement of the collapse test. each grillage was subjected within the elastic range to
various combinations of in-plane and lateral loads. During collapse tests compressive loads were
applied incrementally with frequent returns to zero to allow evaluation of permanent set of deflec-
tions and strains.
In adjusting the position of jacks for application of compressive load an attempt was made to
anticipate loss of plating effectiveness (caused by buckling of slender panels and yielding associated
with weld-induced residual stress). with consequent upward shift of the plate-stiffener neutral axis,
by introducing a slight initial upward eccentricity of jack loads. It was found, however, that while
this initial eccentricity caused downward bending of stiffeners in reinforced end bays, the effect of
jack eccentricity had largely disappeared (and even in some cases caused slight upward bending)
in adjacent, unreinforced bays as a result of "grillage action" by transverse frames. Jack loads
were therefore finally applied at positions corresponding approximately to the initial neutral axis
of each plate-stiffener cross-section.
In multi-bay grillages with stiff transverse frames and reinforced end-bays, the effect of load
eccentricity on interframe collapse is likely to be small; tests on orthogonally stiffened grillages
should in this respect give a satisfactory representation of actual ship bottom and deck structures
in which the effective line of action of compressive load will to some extent follow any shift of
neutral axis caused by change of plate effectiveness. However, in experiments on single-bay panels
having longitudinal stiffeners only, fixed alignment of axial load in the presence of a shifting neutral
axis may in some cases cause serious experimental error [as was also pointed out by Faulkner
discussing his tests].

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
During tests, extensive measurements of stiffener and plating deflections were carried out using
dial gauges. Overall vertical deflections of each grillage were measured at stiffener intersection
positions by gauges mounted on an independent datum frame. Vertical interframe deflections of
longitudinal girders relative to transverse frames were recorded in selected regions, together in
some cases with horizontal deflections of girder tables. Detailed measurements of vertical deflection
were also made along the centre-lines of selected plate panels relative to transverse frames.

Strains in stiffeners and plating were also measured at many points. In the first tests over
270 strain gages were attached to each specimen. The interpretation of the recorded plating
strains, which were very irregular and clearly influenced by local yielding at an early stage
of loading, was found to be difficult, and hence strain gage readings were relegated to lower
priority in subsequent tests. With modern automatic data reduction techniques, more consis-
tent strain records could probably be obtained, which would take the local yielding into
account.
Both the elastic and the collapse behavior were carefully studied and meticulously reported
in [8.148]. For example, Figure 12.15 shows the test rig with grillage 3b in it after collapse
by interframe flexural buckling, whereas Figure 12.16 shows the fully developed collapse
by lateral torsional buckling of the longitudinal stiffeners of grillage 1a. The buckling and
collapse behavior of each of the grillages was assessed and discussed in detail, including the
plating and stiffener deformations, as for example the deformations of grillage 3b (of Figure
12.15) in Figure 12.17. The observed elastic behavior under lateral pressure and axial com-
pression and the buckling modes were compared with computations by simple approximate
formulae, checked selectively by folded plate analysis [12.60] and finite-element analysis
[12.61].
The testing of 11 full-scale grillages was an enormous task, but the care taken by Dr.
Smith and his co-workers and the insight conveyed in Smith's 1975 paper [8.148] make it
indeed a notable effort and achievement that has been widely cited since.
Near the end of the paper Smith reiterates the importance of extensive measurements and
evaluation of plate and stiffener distortions in ship hulls and recommends the use of "ripple-
scanning" devices, like that developed at Cambridge University [8.66] shown in Figure 8.51
and discussed in Chapter 8, Section 8.2.9.
He concluded with a "Note on Experimental Procedure" that still applies today and there-
Copyright Wiley
fore warrants recapitulation:
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
928 Stiffened Plates

PAKEL R PA.J-lEL C

cz~~TION ~~
z
"
-0: 7

0·2 [ PERMA!C'OT SET ·.~

~
0·2 r JJ>;ITIAL ~
1 DEFORMATlOl\ j_,..........-- ---..........__
~ 0~~~1 T4
""" -o·2 L~ ------ T3

Figure 12.17 ARE (NCRE) Dumfermline tests on steel welded grillages-plating and stiffener defor-
mations for specimen 3b (from [8.148])

"1. Design of Test Structures: in order to avoid experimental error caused by non-
uniformity and eccentricity of compressive load and by unrealistic support conditions,
orthogonally stiffened grillages containing at least three and preferably four or more
interframe spans should be tested rather than single-span stiffened panels; in such test
structures realistic allowance will be made for interaction between adjoining spans.
End-bays should be reinforced (or reduced in length or manufactured in higher strength
material) in order to avoid premature failure in these regions.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

2. Material Properties: Compressive as well as tensile yield strength should be measured


using standard test methods with reasonably low strain rates (e.g. in accordance with
BS 18); test specimens should be sufficient in number to provide a statistical definition
of material variability.
3. Initial Deformations: plating and stiffener distortions should be thoroughly surveyed,
including horizontal deformation of stiffener tables; continuous profiles of plating dis-
tortion should be recorded using a scanning device in order to allow evaluation of
modal deflection components.
4. Residual Stresses: high priority should be given to measurements of residual stresses
in stiffeners as well as plating, ideally at various stages during the welding procedure.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Plates Subjected to Axial Compression 929

Weld-induced strains are readily measured using a small portable extensometer (e.g. a
Demec extensometer); readings should be checked for repeatability and corrected for
temperature effects. Particulars of the welding procedure should be recorded, including
the amount of weld material deposited and/ or the welding voltage and current em-
ployed.
5. Deflection Measurements: a thorough record of plating and stiffener deformations
should be obtained, including horizontal deflection of stiffeners wherever tripping is
likely to occur and including permanent sets.
6. Strain Measurement: measurement of strains at carefully selected positions is clearly
desirable as a means of monitoring local plating and stiffener deformations and of
checking the uniformity of compression across test grillages; it should however be
borne in mind that, particularly during the latter stages of load application when parts
of the structure have yielded extensively causing permanent set of strains, evaluation
of stresses may be impossible.
7. Load Application: ideally test structures should be subjected to controlled end-
shortening up to and beyond collapse rather than to controlled loads; in this way
deformations and strains may be correlated accurately with applied load in the sensitive
range close to collapse where stiffness is low and may be negative, and a record may
be obtained of post-collapse load-carrying capacity and stiffness (which are of interest,
for example, in evaluating the strength of a multi-deck ship where collapse of one deck
may not precipitate hull failure). Controlled application of end-shortening is likely to
be achieved more effectively using a mechanical, screw-jack type of testing machine
than using hydraulic load application unless a sensitive, travel-related servo-control is
incorporated in the hydraulic system. Loading procedures should include frequent re-
turns to zero load to provide a record of permanent set of deflections and strains" [8.148).
Indeed, sound advice, which could today be slightly improved upon by use of more
sophisticated and automated measurement techniques, as used for example in later ARE
tests, discussed below.
Experiments on buckling of stiffened plates continued at ARE in the eighties with the
introduction of aluminum alloys in the structure of British warships, such as frigate super-
structures, hydrofoils and surface effect ships. A series of measurements and tests were
therefore initiated to extend the understanding of steel structures to that of stiffened alumi-
num plating [12.62].
Compression tests on five stiffened N8 aluminum alloy plates representing typical full-
scale warships scantlings were carried out. The test panels of different plate and stiffener
sizes were manufactured using normal shipyard procedures to ensure typical initial imper-
fections and residual welding stresses. The latter were extensively measured during construc-
tion. The original 3.7 X 2.7 m panels were cut into two each, and end and side supports
were welded on to give eight compression test specimens, each having three longitudinals
and two full-frame spaces (see Figure 12.18), five of which were tested. The subdivision
resulted also in each grillage containing two different plates joined by a longitudinal butt
weld offset from the centerline, as shown in the figure.
The experiments on the aluminum alloy panels were similar to the earlier ARE ones on
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

welded steel grillages, except that the sides of the specimens were supported here only at
the frames so that failure mode was interframe buckling. Also, the applied loading here was
displacement controlled so that the postbuckling behavior could be monitored.
The test frame employed is shown in Figure 12.19. The grillages were loaded by four 500
kN servo-hydraulic jacks operated in the displacement mode. These jacks "incorporated load
cells. but as a check on the longitudinal loading at the supports the load was also measured
at the reaction end (left side of diagram) using six load cells. The supports at the frame ends
were double bottle screws [no. 4 in Figure 12.18] connected to bars which enabled the top
and bottom connections to slide longitudinally under load." For one of the specimens the
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
930 Stiffened Plates

"~

I I
1QOO 610

-~J-,
A~)
I ',--
'7~ 9,'\"
"''"'

t
] f-
-
"e- - - - -- ---------
PLAT[

'WHO
--------- -- --- - :E,
BBO
1-
-- PlATE B

" f-.-
·- }
,~"dbe dbb ;;),!£; (0" -
~
\
,3j

Figure 12.18 ARE Dumfermline tests on aluminum alloy stiffened plating-typical test specimen with
nominal dimensions (from Ret·. [12.62)): (IJ 12 mm end plates welded to grillage, (2)
152 x 89 steel horizontal !-beam. with four 6-mm webs each side, bolted to end plates,
(3) 152 X 89 steel horizontal 1-bcam, with six 6-mm webs each side, bolted to end plates,
(4) vertical supports at transverse frames, (5) 5 mm doubler plates (both ends) over initial
300 mm, both sides of stiffener webs and both sides of plate (except one side of plate
only for two specimens)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

failure load was estimated theoretically to exceed the capacity of the four servo-hydraulic
jacks. Two additional 250 kN pressure-controlled jacks were therefore added, which were
switched in when the total load reached 600 kN.
As mentioned earlier, more sophisticated distortion measurement techniques were em-
ployed in the aluminum alloy tests. Longitudinal, transverse and vertical plate and stiffener
deflections were monitored by many ARE deflection transducers covering the test specimen.
These transducers were supported on a light datum frame and their signals were recorded
on a data logger. An additional measure of plate deflection was obtained from a transducer
that was scanned along the centerline of the plate panels (on the unweldcd side). The output
of this scan was recorded on line on an X- Y plotter and subsequently digitized to produce
scaled plots.
The initial distortions were also obtained in the same manner. The stiffener distortions
were measured by three different method': (1) survey of the original panels with a displace-
ment transducer mounted on a bar, (2) using a swept laser beam as a reference plane, and
(3) measurement of the final test grillages using a laser delleetion method (also employed
in the Braunschweig tests, discussed in Subsection 12.3.6; sec [12.65]).
The strains in the plating and longitudinals were monitored by many foil resistance strain
gages distributed over the two middle h2:ys of the grillages and were recorded via a strain
gage logger for subsequent computer analysis.

4 • 50 T JACKS
[lOS • 89 CHANMEL

6 LOAO CHLS

Figure 12.19
------------------7010

ARE Dumfermline tests on aluminum alloy stiffened plating-diagram of test frame


.I
(from [12.621)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Plates Subjected to Axial Compression 931

Load-deflection curves could be obtained automatically from the recorded measurements.


During the acquisition of this data, the deflection was held for at least three minutes before
logging, to minimize short-term creep effects. The experimenters also pointed out that at the
higher deflections care had to be taken to ensure that the grillage did not move the light
datum frame and that deflection transducers which were nearing their limits were reset or
removed.
The tests were continued until the overall behavior could no longer be accurately moni-
tored. The experimental maximum loads and postbuckling load-deflection results were com-
pared with the theoretical predictions using the ARE computer program Nl06C [12.63]. The
postbuckling load-carrying capability was, however, underestimated by the two-frame finite-
element model used for all the specimens. This was probably primarily due to the additional
rotational restraint at the frame ends or to the additional restraint provided by the short
sections with doubler plates. As often, incomplete definition of the boundary conditions is
the main villain!
Before we leave the noteworthy ARE tests, it may be noted that they were closely coor-
dinated with parallel studies at Cambridge University, discussed in Chapter 8 [8.58] and
[8.59] and at Imperial College, London [12.16].
It should also be pointed out that recently a large-scale testing system for stiffened plates
under combined axial and lateral loads was built in Edmonton, Alberta, Canada (see
[12.102]). The construction of the test system and the experimental program of tests on 12
2000 X 500 mm stiffened plates were jointly sponsored by the U.S. Ship Structure Com-
mittee and the Defense Research Establishment of Atlantic Canada. The test setup, the Centre
for Engineering Research (C-FER) laboratory Tubular Testing System (TTS), is a high-
capacity vertical computer controlled servo-hydraulic system (15,000 kN axial and 5000 kN
lateral static loading), suitable for combined loading buckling tests of stiffened steel plate
components, representative of those used in ship structures. Further buckling studies in this
experimental facility are planned.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
12.3.6 Large-Scale Civil Engineering Tests
Axially compressed stiffened plate panels are often employed in civil engineering structures,
in particular as bridge deck elements, box girder flanges or panels in crane construction.
Hence they have also been the subject of many studies in civil engineering laboratories. In
order to account correctly for fabrication imperfections and residual stresses, most of the
stiffened plate tests in civil engineering were rather large scale.
A good example are the experiments carried out by Barbre, Scheer and their co-workers
at the Institut fiir Stahlbau of the Technical University Braunschweig in the late seventies
and early eighties [12.17] and [12.18]. The comprehensive test program (performed in the
Braunschweig test rig, see Figures 12.20 and 12.21) included 65 large longitudinally stiffened
steel plates, 45 with free unloaded ends and 20 with simply supported ones. The test spec-
imens, 7-mm thick plates with four flat-lipped oustand stringers welded on to each, were
0.88-5.33 m long and 0.7-2.1 m wide. They were not completely full scale, but represented
about I: 1.4 scale models of typical building structures, the plate thickness of 7 mm repre-
senting a reasonable compromise between actual structures and manageable model sizes that
would still realistically simulate imperfections and residual stresses.
The specimens were designed to cover the two primary failure modes, stiffener failure
mode and plate failure mode (see Figure 12.22). The first two test series, with free long
ends [ 12.17], were divided into five groups of nine models each, one group of near perfect
ones, two groups of exaggerated geometrical imperfections and two groups of both exag-
gerated geometrical imperfections and exaggerated residual weld stresses. The predetermined
exaggerated geometrical imperfections included, for specimens biased towards a stiffener
failure mode, typical short wave torsional deformations as well as a global sinusoidal one.
For specimens biased towards a plate failure mode, the predetermined geometrical imper-
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
932 Stiffened Plates

~. I C ROSSSECTIOMS .. B ·~
Oetolt ot

r
tooding tdgt CROSS SECliOI\ A·A
OElAIL C

DETAILD

Jbfi
Figure 12.20 TU Braunschweig large-scale tests on axially compressed Stiffened platcs- 8000 kN-
test rig (from [12.18]. courtesy of Professor J. Scheer): ( I) test specimen, (2) longitudinal
beam, (3) horizontal loading edge. (4) inductive displacement transducer, (5) inductive
displacemelll transducer, (6) upper cross-beam. cruT)•ing longitudinal beams (2), (7) sti ff
movable cross-plate, (8) guides for movable cross-plate (7), (9) 2000 kN loading jack
with load cell, ( I0) fixed stiffened cross-plate. ( I I) longitudinal main beam of base frame.
( 12) pendulum bar, (13) vertical support of upper cross-beam
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 12.21 TU Braunschweig large-scale tests on axially compressed sliiJened plates- view of test
rig (courtesy of Professor J. Scheer)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Plates Subjected to Axial Compression 933

figure 12.22 TU Braun;chweig large-scale tcM• on axially comprcs;cd stiffened plmes- failurc modes:

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
(a) sti fTcncr failure mode. (b) plate failure mode (courtesy of Professor J. Scheer)

fcctions included typical short wave chessboard defonnations and a global sinusoidal one.
The exaggerated welding stresses were obtained by heat-treating the first three groups to
yield practically stre~s- free models. while omitting such heat trcaunem in the last two groups.
The predetennined imperfections of both kinds facilitated calibrated calculations that pro-
vided. by interpolation. approximate design data for realistic practical structures. Note. how-
ever, that for more accurate predicti ons, more extensive measurements and red uction of
typical geometric imperfections occulTing in practical structures would be necessary, as dis-
cussed in Chapter I 0.
The second two test series. with simply supported long ends [12.18), were divided into
two groups of 10 specimens each. one group with unifonn axial compression and one with
a triangular axial loading.
End plates were welded onto all the test specimens for load appl ication, wi th their outside
faces milled to a plane and perpendicu lar surface.
A lithe tests were carried out in a horizontal 8000 kN test rig (Figure 12.20) at the lnstitut
fiir Stahlbau of the Technical University Braunschweig, that could accommodate plates of
up to 9 m length and 2.7 m width. The base frame of the test rig consists of two longitudinal
main beams (no. II in the figure) connected by two cross-stiffened plates (I 0). Four 2000
kN jacks (9) react agai nst these stiffened cross-plates ( 10) and apply their loads via load
cells to the stiff movable cross-plates (7), which move in horizontal gu ides (8) attached to
the main beams (I I). The jacks and cross-plates are arranged in pairs, at both ends of the
test specimen (I). Or in other words, Figure 12.20 shows only the right-hand half of the test
rig. with the symmetrical left-hand half omitted. The movable cross-plates (7) apply the load
to the horizontal loading edge (3) of the te~t model (I). which represents simple supportS.
a~ shown in cro~~-section BB.
In the first two test series [12.17]the unloaded ends of the specimens were free edges. In
the ~econd two test series [ 12.18). simply supported unloaded edges were represented by
discrete point supports 70 mm apart. These points of the test model were supported vett icall y
by pendulum hars ( 12). attached to the specimen and to the longitudi nal beams (2) with
spherically shaped nuts. as shown in detail C of Figure 12.20. T hus, the unloaded edges of
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
934 Stiffened Plates

the test specimens were free to rotate in any direction or move freely in the plane of the
plate, as specified by simple suppolts. This type of edge support essentially resembles the
support "fingers" developed independently (and earlier) at Cambridge University for unstif-
fened plates and discussed in Chapter 8 (see Figure 8.34). The longitudinal beams (2), to
which the tops of the pendu lum bars were connected, were held by the upper cross-beams
(6) that were fixed to the main beams (II) by vertical suppolts ( 13).
The servohydraul ic loading system operated in a controlled displacement (controlled end
shortening) mode in order to enable the load-displacement behavior to be studied well into
the postbuckling regime. The end shortening of the specimen was measured by the two pairs
of inductive displacement transducers (nos. 5 and 6 in Figure I 2.20). Axial loading was
appl ied by a controlled end shorteni ng at rates of 0.025 - 0.25 mm/min. Since this did not
correspond exactly to the required quasi-static loadi ng, the continuous end shorteni ng was
interrupted at certain intervals and the end shortening was kept constant. In the clastic-plastic
region, the load then dropped due to plastic Row and relaxation. When these had subsided,
measurements were taken. In the first three test series the same oil pressure was supplied to
aU four pairs of jacks by the control system, providing the uniform ax ial compression. In
the fourth test series the desired triangu lar loading was obtained by applying pressure with
the electronic control system only to the fi rst two pairs of jacks on one side, not using the
third jack, and operating the fourth pair of jacks manually with a hand pump. The end

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
shortening of the strained side, measured by one displacement transducer (no. 5 in Figure
12.20), was the reference value for the servo-hydrau lic control. The strain measured at the
other unloaded end of the model by the second transducer (no. 4 in Figure 12.20) should
be zero. and was kept zero by continuous correct ions with the hand pump.
The test specimens were mounted in the test rig with the stiffeners facing downwards. As
usual, measurements included axial loads (measu red by the load cells), strains along the
middle line of the specimen, vertical deflections under the sti ffeners and in the centers of
the subpancls and horizontal deflections of the stiffeners. The deflections were measured
with the aid of two longitudinally movable, bridge-like measurement carriages, one under-
neath the specimens (with ni ne displacement transducers) for vertical plate deformations (see
Figure 12.23) and a simi lar one above for horizontal stiffener deformations.
The test results were compared with the permissible loads of the German code DASt-
Richtlinie 012, and it was found that more attention should be paid to stiffener failures,
which are sometimes not assessed correctly by the code.
The TU Braunschweig experimental investigation was also extended to similar stiffened
plates subjected to combined axial compression and in-plane shear loads, using the same

Figure 12.23 TU Braunschweig large-scale tests on axially compressed stiffened plates-lower mea-
suring carriage for plate deformations (from I 12.17], courtesy of Professor J. Scheer)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Plates Subjected to Axial Compression 935

horizontal test rig (shown in Figures 12.20 and 12.21), which had been designed also for
application of shear along the longitudinal edges [12.65].
The TU Braunschweig studies also included other tests on axially compressed stiffened
plates with trapezoidal hollow stiffening ribs [12.66] and [12.67] and were the stiffened plate
part of a broader plate buckling test program ("Plattenbeulversuche") sponsored by the
German Steel Construction Committee (Deutschen Ausschuss ftir Stahlbau) that involved
also TU Darmstadt, TU Karlsruhe, TU Hanover and Stuttgart University (see [12.68] for a
summary of the whole program).
Another interesting example of large-scale collapse experiments on stiffened civil engi-
neering plate structures were the investigations carried out on an orthotropic steel bridge
deck at the Department of Civil Engineering of the University of Calgary, Alberta, Canada
in the late sixties and seventies [12.69]. An experimental and analytical study of the bending
collapse of an approximately half-scale model of an actual orthotropic (ribbed) bridge deck
was performed. The test specimen was a three-bay steel deck plate of overall dimensions
4.98 m (196 in.) X 2.90 m (114 in.) stiffened by floor beams and nine rounded closed-
section ribs welded to the underside (see Figure 12.24).
Though the loading in these tests was by two concentrated lateral loads applied at midspan,
and hence they were mainly bending and not buckling tests; and though, except for buckling
of the flanges of the main girders and the edge of the deck plate, failure was primarily by
yielding, these experiments present an interesting point worth noting, to which the discussion
will be limited:
Pilot tests indicated that in order to reach the ultimate load of this deck panel, the rectan-
gular loading frame of the Calgary laboratory had to be redesigned and reinforced to a double
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

SECTION A-A

PLAN OF 0£CK PLAT£

Figure 12.24 University of Calgary bending tests on an orthotropic steel bridge deck-test specimen
(from [12.69])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
936 Stiffened Plates

capacity of 2400 kN (540 kips) and two :WO-ton capacity hydraulic rams had to be installed.
The experimenters, however, wanted to ensure that their test frame preserved sufficient ri-
gidity up to the ultimate load of the bridge deck plate. They therefore mounted 19 strain
gages at critical locations on the frame and thus monitored its behavior during the test.
The careful monitoring of the test frame behavior during a test is rather unusual. but it is
certainly worth considering in cases of doubts about the relative rigidity of the test frame.

12.3.7 Single-Stiffener Panels


Single-stiffener panels are widely used in civil engineering applications ranging from com-
pression flanges of large steel box girder bridges to light-gage steel roofing sections. They
have an economic advantage over wide-stiffened panels in that they can be used without
cross-framing. Hence the ultimate strength behavior of single-stitiened panels has been ex-
tensively studied in the last decade, both experimentally and theoretically (see for example
[12.70]-[ 12.72]).
Single-stiffener panels can take a number of different forms as shown in Figure 12.25.
The stiffener can either be placed symmetrically about the midplane of the plate, or only on
one side. It can either be welded or extruded onto the section, or folded, and in some cases
also glued on, as in plates with many stiffeners.
Single-stiffener panels have two predominant modes of failure (sec Figure 12.26):
l. Overall buckling, mode 1, where both the plate and the stitiencr elements buckle out-
of-plane, i.e. the junction between plate and stiffener deflects (sec Figure 12.26a)
2. Local buckling, when the individual plate elements buckle while the junction remains
essentially straight. Local buckling can appear in three forms: mode 2, purely local
buckling (see Figure 12.26b ), where there arc no deflections in the plane of the stiffener;
mode 3, local buckling when the stiffener has a very high torsional rigidity (see Figure
12.26c), where the plates buckle on either side of the stiffener-plate junction, which
neither deflects nor rotates; and mode 4, local buckling of the stificner (see Figure
12.26d), a clamped outstand mode, which is possible when the stiffener is slender.

DOUBLE SIDED

ONE- SID ED -
OPEN FOLDED
-----·· --

Figure 12.25
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`--- Examples of single-stiffener panels
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Plates Subjected to Axial Compression 937

I a)

(b)~

Figure 12.26 Buckling modes of single-stiffener panels, shown here for


one-sided open panels, the displayed traverse mode in each
lcl~ case being modulated sinusoidally along the length of the
panel (from [12.72]): (a) mode !-overall buckling, (b)
mode 2-local buckling of the plate (zero torsional rigidity
of the stiffener), (c) mode 3-local buckling of the plate
(high torsional rigidity of the stiffener). (d) mode 4-local
(d) 0 _ _ _..._( _ _ _0
buckling of the stiffener (clamped outstand buckling)

Furthermore, interactive buckling modes involving two or more of those modes may
also occur.
The buckling and strength tests on single-stiffener flat panels carried out at Cambridge
University in the eighties represent a good example of such experiments. The investigation
included two test series of one-sided single-stiffener panels: one preliminary one of 10 alu-
minum alloy specimens [12.71] and one of 20 mild-steel specimens [12.72].
Except for one specimen, all the tests of both series were carried out at the Cambridge
University large horizontal test setup for axially compressed plates, discussed in Chapter 8,
Section 8.2.4 (see Figure 8.33), with support fingers providing simple supports along the
long unloaded edges of the panels. Some slight modifications were made to this test rig to
admit the single-stiffener panels. A small vertical bearing piece was bolted to each of the
end platens to allow direct axial loading of the stiflener (see Figure 12.27). This piece was
located in the central interval between the left- and right-plate clamping arrangements. It
could he shimmed from behind (between platen and bearing piece) to take up any nonvert-
icality of the machined ends of the test specimen. In order to avoid premature failure, special
reinforcements (no. 6 in Figure 12.27) were fixed to the plate on either side of the stitlener.
The specimens were partially restrained against rotation at their loaded ends by the clamp-
ing arrangement, as in the original test rig. That meant that the test panels had to be long
enough for the influence of the end clamping to be immaterial, i.e. long and slender enough

Figure 12.27 Cambridge University single-stiffener flat panel buckling tests-end clamping, reinforce-
ment and vertical bearing piece for direct loading of stiffener (from information in [12.71]
and [12.72]): (I) platen, (2) vertical bearing piece for direct loading of stiffeners, (3)
stiflener. (4) clamping arrangement of plate, (5) packing, (6) reinforcement for plate, (7)
plate of specimen --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
938 Stiffened Plates

to achieve several buckle half-wavelengths of the overall buckling mode in the 2.44-m (96
in.) test rig. To prevent lateral buckling of the specimen while still allowing the plate to pull
in with its supporting fingers, additional in-plane supports were provided at one--third points
along the length.
In the aluminum alloy specimens of the first series of tests, the stiffeners were initially
glued with Araldite to the plate. After some failures, however, all subsequent panels were
welded together. The mild-steel specimens of the second (main) series of tests were folded
and spot welded (see Figure 12.28), with the top of the stiffener and the loading faces of
the panel machined to be level and parallel respectively. The spacing of the spot welds was
typically 5 em.
Into some of the specimens deliberate residual stresses and initial out-of-plane geometrical
imperfections were introduced. In six of the mild steel specimens of the main test series the
residual stresses were introduced by laying a butt well in the groove between the two L-
sections (see Figure 12.28). The residual stresses introduced by welding were measured in
the middle of all the panels by means of a Weldscan demountable extensometer (which
employs a strain-gaged cantilever as the measuring element) developed at Cambridge [8.68]
and mentioned in Chapter 8. The deliberate geometrical imperfections were introduced in
half the specimens by bumping with a bumping rig, similar to the one employed at Cam-
bridge earlier on unstiffened plates (see [8.59]). Bumping was performed after welding, and
then the initial overall bow was measured by a simple device, discussed in Chapter 8 (see
Figure 8.52).
During the tests, the out-of-plane deflections were monitored by a "ripple scanner" (also
considered in Chapter 8), which could be moved longitudinally as well as sideways and thus
permitted deflection measurements both at the tip of the stiffener and on the plate on either
side of the stiffener. The record of these vertical deflections on an X- Y plotter exhibited the
buckle profile and showed the interaction between overall and local buckling (see Figure
12.29).
Among the conclusions of the study it may be noted that the experimental results indicated
that the buckling mode, overall buckling or interaction between local and overall buckling
had little influence on the ultimate strength of the panels. However, in the interactive buckling
mode the local buckles in the plate had a greater weakening effect than large initial out-of-
plane deflections. Comparison of the experimental strength results with predictions by dif- --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

ferent codes showed that, whereas the relevant ECCS Recommendations and German
Regulations provided satisfactory predictions, the British ones considerably under-
estimated the strength of single-stiffener panels-another case lending support to the exten-
sive code-coordinating efforts under way in the last decade.

•: :2t
5

"fingers"

/
\)

.,__ _ _ b ---~
2S


I: •
Figure 12.28 Cambridge University single-stiffener fiat panel buckling tests-cross-section and side
view of the mild-steel specimens of the main test series (from [12.72])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Plates Subjected to Axial Compression 939

0 0
0 0
0 0
0 0

------============---.::.::~---~-- 8
Figure 12.29 Cambridge University single-stiffener panel buckling tests-ripple scans for typical in-
teraction between overall and local buckling (from [12.72], for a specimen without butt
weld). Ripple scans taken at: (1) 0.25 ultimate load, (2) 0.80 ultimate load, (3) ultimate
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

load, (4) 0.80 ultimate, after peak load

Another example of single-stiffener panel tests is those carried out, as part of a wider
experimental program on the collapse of continuously welded stiffened plates, at the Civil
Engineering Structural and Material Laboratory of Pontifical Catholic University of Rio de
Janeiro, Brazil, in the eighties [12.73]. The whole research program included 17 welded steel
plates, stiffened by one to three longitudinal and one to two transversal stiffeners of three
different cross-sections (rectangular, L- and T-stifieners). All the plates were square, of 750
mm X 750 mm, simply supported and uniaxially compressed. Three of the test specimens
were single longitudinal stiffener panels, and three were single-stiffener panels with the
addition of one transversal stiffener, for which (bit) was about 70.
A relatively stiff test rig was employed, and the load was applied by two 500 kN jacks
to a rigid loading beam, which was constrained by sliders and rollers to move vertically in
the direction of load and in the plane of the test plate. To promote uniform load distribution
into the specimen, two base bars of 51 mm X 16 mm X 750 mm were welded to the upper
and lower (loaded) ends of the plates.
The experimental results of the entire program indicated that the type of stiffener cross-
section had relatively little influence on the ultimate stress but significantly affected the
failure mode, in particular in the single-stiffener panels where stiffener failure occurred. It
was also found that the addition of transverse stiffeners hardly changed the maximum stress
but substantially reduced the maximum out-of-pane deflections.

12.3.8 Aerospace Tests


Though the extensive experimental studies of the forties and fifties on the buckling and
strength of aerospace type stiffened plates (for example [12.3]-[12.7] and [12.45]-[12.48])
provided substantial aerospace design data, testing of aircraft-type stiffened panels has con-
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
940 Stiffened Plates

tinued to the present, to investigate new concepts for wing panels and other primary structural
elements and verify new methods of analysis.
For example, in the late seventies a series of experiments on small integral and riveted
stiffened fiat panels was carried out at the Japan National Aerospace Laboratory, comple-
menting and verifying an elastic-plastic and geometrically nonlinear finite element analysis
[12.74].
More recently, the wing panel compression tests for the Boeing 777 airliner [12.75] rep-
resent a typical modem development program of the stiffened fiat panels. The purpose of
the tests was the establishment of static design values for the model 777-200 wing upper
surface skin and stringers.
Each specimen was reinforced by six Z-stringers (see Figure 12.30a) or by three vent
(large hat-section) stringers, both plates and stiffeners being made from 7055 aluminum alloy.
All the 38 stiffened panels were tested to destruction (see for example Figure 12.30b). The
panel test data, and the data of the accompanying test coupons, were the basis for the design
values used on the 777 wing.
The stiffened panels were tested in a large 550 ton Baldwin universal testing machine.
Strain gages were attached to the panel:> to monitor the load distribution and the failure
stress. The measured strains were recorded on a digital data acquisition system, with parallel
on-line plots. The panels were centered in the test machine and shimmed at 10, 20 and 30
percent of the predicted failure load for improved loading uniformity. This shimming was
continued till all the strain gages read within 10 percent. The panels were loaded in 50 kip
( ~220 kN) steps, with data being recorded at each load increment, till about 67 percent of
the expected failure load. Then the loading and data recording proceeded continuously to
destruction.
The instrumentation employed in the tests included axial strain gages for the load align-
ment, mentioned above in connection with the shimming procedure, and for comparison with
nonlinear finite element analysis. Shadow moire techniques were employed to measure out-
of-plane deflections on all the panels. The analysis of the moire patterns for deflection versus
load data was enhanced by digital image processing. High-speed videos were used to identify
the details of the failure mode of the stringers. For example, Figure 12.30(b) shows a frame
just before failure from such a high-speed video, exhibiting the collapse mode of a typical
Z-stringer panel. The corresponding nonlinear finite element model predicted failure pattern
at the moment of failure agrees very well with that observed in the test. Similar good
agreement between prediction and experiment was obtained for the other panels.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 12.30(c) shows the moire fringe displacements for another typical Z-stringer panel
at I 00 percent maximum applied load, as well as the displacement profile along line A1-
A2. Again, the moire fringe displacements agreed well with the corresponding finite element
predictions at the various loading stages till failure.
The full-scale static test of the fuselage of the Domier Do 228 regional aircraft, is another
example of stiffened fiat-panel tests for aerospace applications. The fuselage of this multi-
purpose 20-seat aircraft has a square cross-section (see Figure 12.31) and therefore represents
a collection of fiat panels. The test setup [12.76], shown in Figure 12.3la, is typical of a
full-scale fuselage test frame for small aircraft. Figure 12.3lb shows a side panel (with
window and emergency exit cutouts) subjected to the loading of load case "level landing,
spin up," with a load factor of 1.3. The loading is primarily shear and has resulted in well-
developed shear buckling. This can clearly be seen near the window and emergency exit,
but even more distinctly on the right-hand side of the picture, where well-defined tension
fields have evolved.
Another recent example is the tests carried out on curved stiffened panels for the Airbus
A 330/340 (see [12.77], [12.78] and [12.103]). These were static tests with shear, axial
compression and combined shear-compression loading on 12 stiffened panels for the fuselage
of the large Airbus A 330/340 transport aircraft. The combined loading test setup (see Figure
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Plates Subjected to Axial Compression 941

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
(a) (b)

DISrtACHM ENT PRO PILE A I·A:!

•·"' I
•... I I
·"'
•·"' f\ f I f
• II \ I

. !( !
I \ n "'
,u \J
il 1\ I \ I
v
•.
• I
•... I I
•·"'

Figure 12.30 Boeing Co mmercial Airplane Company 777 wing panel cornpre.~sion tests-typical Z-
stringer ftat panel test: (a) test specimen. (b) still from high-speed video (for panel no.
5) at the moment of failure, (c) moire fringe displacements for another simi lar panel (no.
14) at maximum applied load. as well as displacement profi le along line A 1-A2 (courtesy
of Boeing Commercial Airplane Company)

12.32a) has been developed primarily for curved panels (and could therefore also be dis-
cussed in Section 13.6 of Chapter 13), but it can also be used for flat stiffened panels. The
emphasis here will be on some features that apply both to flat and curved panels.
ln Figure 12.32a the test setup is shown with a curved stiffened panel in posi tion. T he
curved panel (which represents an arc of circle) is kept in position by swinging supports
(whose length is the radius of the panel) that rotate about the ax is of the panel. For flat
panels this complication would not be necessary.
The combined loadi ng capability of the test rig is essential for simu lation of realistic flight
load cases. The loads, both axial compression and shear. or their desired combination are
applied with aid of a computer, integrated in the test system, that monitors the deformations
of the panel and controls the hydraulic jacks.
One feature of load application is of special interest. In the case of combined loadi ng, the
vett ical shear is supposed to be applied as a uni formly distributed loadi ng. Since the axial
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
942 Stiffened Plates

(b)

Figure 12.31 Domier fu ll·scale static te't of the Do 228 regional aircraft: (a) test setup for fuselage
tests. (b) side panel subjected to shear loading and showing well developed shear buck·
ling (courtesy of Domicr Luftfahrt GmbH. Fried1ichshafen)

compressio n shortens the panel. possible non·unifonni ty of vertical shear is min imized by
its application alo ng bo th sides of the panel through 20 pairs of hydraulic cy li nders fixed in
IOO·tmn spaci ngs. Each pair of jacks can apply a load of 60 kN. the total vertical shear
capacity on each side be ing thus 1200 kN . To obtain an approximately unifo rm d istribution
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

of the vertical shear. these 20 pairs of jacks are interconnected via a hydrauli c man ifold, into
which the oil is supplied from the servo.valve. The complicati on of the many smaller hy·
draulic jacks (not shown in Figure 12.32a) was considered to be justified by the more uni·
formly distr ibuted shear loading achieved.
During the tests, strains were measured with strain gages. the pane l shortening by ind ucti ve
displacement transducers. the forces by load cel ls and deformation patterns by photography.
All the measured data were recorded and reduced by data loggers, coupled with the computer.
Details of the measuring and control system can be found in 112.78].
Fig ures 12.32(b) and (c) show deformation patterns of a typical combined loading test of
a relatively thin· walled panel (no. I I): (b) compression only, at a load factor 1.0 (80 percent
of fai lure). exh ibiting well developed local compression buckles and (c) combined compres-
sion and shear (53 percent of ultimate shear and 80 percent of ultimate compression), at a
load factor 1.3, with deep di sto11ed shear buckles. Failure occurred at a load factor of 1.41.
Many other combi ned loading cases arc reported and well documented in [ 12.78].

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Sandwich Plates 943

(a) CURVED
PANEL

(b) (C)

Figurt' 12.32 Daimler-Ben/ Aero>pace Airbus Miffcncd CUI'cd panel tests for Airbu~ AJ30/340 (from
[12.78), counes) or Daimler-Benz Acn"Jl""cl: (a) combined loading tcm 'ctup (sche-
matic). (b) deformauon pattern for compre"ion only. at a load factor or 1.0. ~flowing
local comp"'"'on buci.Jes. (c) defonnation p:ottcrn for combined compression and >hear,
at a load factor or 1.3. exhibiting deep. cli>toncd ,hear buckles

12.4 Sandwich Plates


12.4.1 Sandwich Structures
Sandwich plates are one form of stiffened plate~. Sandwich conslruction. with it~ obvious
advantage of a large bending .,tiiTness. obtained by ~pacing far apart the face ... the load-
carrying clements. with the atd of the lightweight core. ha' been widely employed in aero-
~pace vehicles for decades. h s applications have become even more prcvalenl with the in-
lroduction o f laminated composite slructures.
The buckling and postbuckl ing behavior of sandwich struct ures has the re fo re been exte n-
sively s tudied theoretically and expe rimentally. attetH io n be ing paid to the typical sandwich
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
944 Stiffened Plates

modes of failure, such as faceplate yielding, faceplate wrinkling, intra-cell dimpling (face
dimpling) and shear crimping (see Figure 12.33). The literature on sandwich structures, and
in particular on sandwich plates, is quite voluminous: from the early studies in the forties
and fifties (for example [12.79]-[12.82]), the textbooks and handbooks of the sixties, sev-
enties and eighties (such as [12.83]-[12.88]), the reviews and conference proceedings of the
eighties and nineties (such as [12.89]-[12.92]) and the many papers and reports up to the
present (for example [12.93]-[12.101]).

12.4.2 Buckling Tests of Sandwicl1 Plates


The series of buckling and collapse tests on aluminum sandwich plates carried out at the
Aeronautical Research Institute of Sweden (FFA) in the seventies [12.96] is a good example
of a test program that focused on the special experimental problems presented by sandwich
construction. Motivated by the development of the Saab Viggen fighter aircraft, a compre-
hensive analytical and experimental program was initiated at FFA in the sixties and continued
into the seventies and early eighties.
The experimental investigation of [12.96] aimed at verification of a method of analysis
developed at Saab Aircraft Company in the sixties and seventies for sandwich panels with
a honeycomb core subjected to combinations of in-plane axial and shear forces and lateral
pressure (see for example [12.97], one of a series of Saab reports on sandwich panels, all
in Swedish).
The test specimens consisted of box columns, or cylinders with a square cross-section,
assembled from four identical honeycomb sandwich panels. As shown in Figure 12.34, this
box could be subjected to an axial force P, a torsional moment M and an internal pressure
p. The axial force was applied in a 500 kN static testing machine via a heavy axial roller
bearing, which allowed application of the torsional moment. However, in the presence of
axial compression, this torsional moment M apparently contained a small error due to the
friction of the bearing, which was estimated to be less than 5 percent of M. The internal
pressure was applied by an inflated rubber bag placed inside the box. Sixteen box specimens,

I I II II I I

:---facing

facing-

core

t t l t t t t t
(a) (b)

t I II I I I

core

I it l I t tI
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

I tI I
(c) (d) (e)

Figure 12.33 Types of buckling failures for sandwich plates (from [12.89]): (a) overall buckling (gen-
eral instability), (b) shear crimping, (c) face dimpling (intra-cell dimpling), (d) face wrin-
kling-separation from core, (e) face wrinkling-core crushing
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Sandwich Plates 945

Figure 12.34 Swedish FFA buckling experiments on sandwich panels-method of load application to
a box test specimen, schematic (from [12.96]): axial load Pis applied via an axial roller
bearing, two hydraulic jacks apply the torsional moment M, and the internal pressure p
is applied by inflation of a rubber bag placed inside the box

each with four panels 367 mm wide, nine of length 814-861 mm and seven of length 316-
361 mm, were tested. The total thickness of the panels was 10 or 15 mm and different core
dimensions and comer reinforcements were used. The sheet material of the panels was 2024
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

and 7075 Alclad aluminum alloy and that of the honeycomb core was 5052 aluminum alloy.
The boundary conditions represented one of the main experimental problems. The corners
of the test boxes were designed to simulate either clamped or simply supported edges for
the panels, as can be seen in Figure 12.35. The geometry of the corner reinforcements were
carefully designed for each type of specimen. Note, for example, the flexible corner joint
affected by the loop of the outer comer lap plate in the simply supported edges. Also note
the edge reinforcements necessary in both types of edges (see also Figure 12.36). Naturally,
however, neither of the ideal edge conditions was fully obtained, and special tests of the
corner stiffness were carried out and the measured clamping constants were recorded as a
percentage of perfect clamping (1 00 percent), while 0 percent implied simple supports.
Special test devices were employed for evaluation of the material properties of the sand-
wich plates: To obtain the compressive stress-strain curves of the facings, a special test
apparatus developed at FFA for compression testing of thin sheet coupons, in which they
were supported sideways while compression was applied by a ram guided with roller bear-
ings, was used. The shear properties of the honeycomb cores were obtained by two methods:
(1) a four point shear loading of a panel strip, and (2) a block shear test, in which only the
core was loaded by shear, yielding its stiffness and ultimate strength.
The initial panel thicknesses were measured, as were the initial curvatures of the middle
surface of the panels. Since the initial curvatures changed when the four test panels were
assembled to a test box, the initial deflections were measured and recorded before and after
the assembly. When the initial deflection magnitude of the assembled structure exceeded 0.5
mm (i.e. the initial deflection exceeded 3-10 percent of the panel thickness), corrections
were made with tapered shims or milling of the corner profiles.
Typical failure modes of a box test specimen are shown in Figure 12.37. The specimen
shown, no. 5, had boundary conditions close to simple supports, estimated at 20 percent
clamping. It was first loaded by axial compression, then by internal pressure Finally it was
loaded by a constant axial compression, a constant internal pressure, and then by increasing
torsion till the ultimate shear, resulting in failure. The failure mode indeed shows the effect
of torsional buckling. However, in most cases it was impossible to judge where the collapse
initiated and to identify precisely the predominant mode of failure, since most specimens
were heavily damaged due to the release of energy at the point of failure.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
946 Stiffened Plates

(a)

(b)

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
(c)

Figu re 12.35 Swedish FFA buckling experimenl> on sandwich panels-design of comers and edge>
simulating clamped or simply ~upported edges (from [ 12.%]): (a) clamped edges. (b)
simply supported edges. (c) reonforcemcnt> used for panels with clamped edges to reduce
local torsion at the comers

ln addition to the box specimens. eight single sandwich panels were tested in pure axial
compression. These panels were compressed in a simpler test rig (see Figure 12.38). with a
series of ball hearings ensuring simple supporLs at the loaded ends, while at the unloaded
edged pairs of knife edges supplied an approx imation to simple supports. These fairly SO·
phisticatcd edge supports were similar to those in some plate buckling test setups discussed
in Chapter 8 (sec Figures 8. 17. 8.19. 8.20. 8.45 and 8.110). At the knife edges of the
un loaded ends. prohably thin Tellon tapes or some grease were used to reduce the in-plane
frictional rcwai nK

Figure 12.36 Swedish FFA buckling experiments on sandwich panels-typical box speci men, with
Copyright Wiley clamped edges and local torsion reinforcements. in test setup (from [12.961)
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Sandwich Plates 947

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
figure 12.37 Swedish FFA buckling experiments on sandwich panels-failure modes of specimen no.
5 (from r12.961)

Figure 12.38 Swedish FFA buckling experi ments on sandwich panels- test rig for uniaxial compres·
sion of single panels (from [1 2.96)). Note the ball -bearings of the loaded edges and the
Copyright Wiley knife edges at the unloaded ones
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
948 Stiffened Plates

Sltif:R srm
ROLLER

(b)

Figure 12.39 Bristol University buckling experiments on sandwich panels-edge supports of plate
(from [12.95]): (a) side supports, (b) loaded end supports

In general the agreement between theory and experiment was found to be good. For
ultimate loads the agreement was better when collapse was initiated by yielding of the face
sheet material, than when the core strength was the limiting factor.
Another example of a test program that emphasized some of the special experimental
problems of sandwich construction is that carried out at the Department of Aeronautical
Engineering of the University of Bristol in the early seventies [12.95]. The aim of these
experiments was to determine the overall buckling and faceplate wrinkling loads of sandwich
plates, with carbon fiber-reinforced plastic (CFRP) faceplates and nonmetallic honeycomb
(Ciba-Geigy Aeroweb) cores, when subjected to uniaxial compression.
The sandwich panels were specified to be simply supported, and it may be of interest to
examine the experimental edge supports employed in the Bristol tests. Details of the side
supports, of the unloaded edges are showrr in Figure 12.39a. Rounded knife edges press on
the sandwich plate, without any local reinforcement, preventing transverse displacement
while permitting rotation, as required by simple supports. Except for the significant rounding
of the knife edges, these side supports resemble those employed for other plates, as described
in Chapter 8.
In the end supports (see Figure 12.39b) provision had to be made for the transmission of
the load to the panel. This was achieved by aluminum end blocks, as shown, which also
prevented delamination of the faceplates at the loaded ends. The end blocks, with silver steel
rollers bearing in V-blocks, imposed some unwanted rotational restraint. To minimize this
restraint, the rollers were greased before being inserted into the buckling test rig. To obtain
uniform loading, a special device, based on that used by Hoff et al. in 1948 and discussed
in Chapter 8 (see Figure 8.45), was employed below the top platen. This load adjuster fitted
over the upper V-block and transferred load by five bolts that were adjusted according to
the strains measured by strain gages during the early stages of loading (at 0.05-0.25 of the
buckling load). In the Bristol test setup it was considered sufficient to place only one load
adjuster at the top end instead of the two, one at the top and one at the bottom, used in the
original Brooklyn test rig. The uniform loading device proved to be rather successful.
The Bristol test program included both specimens designed for overall buckling, whose
test results agreed reasonably well with theory, and thicker ones designed for failure by
wrinkling, whose experimental results were somewhat equivocal.

References
12.1 Thompson, J.M.T. and Lewis, G.M., On the Optimum Design of Thin-Walled Compression
Members, Journal of Mechanics and Physics of Solids, 20, 1972, 101-109.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 949

12.2 Thompson, J.M.T. and Supple, W.J., Erosion of Optimum Designs by Compound Branching
Phenomena, Journal of Mechanics and Physics of Solids, 21, 1973, 135-144.
12.3 Cox, H.L., Thurston, F.R. and Coleman, E.P., Compression Tests on Seven Panels of Mono-
coque Construction, Aeronautical Research Council (U.K.) R&M No. 2042, May 1945.
12.4 Cox, H.L. and Riddell, J.R., Buckling of a Longitudinal Stiffened Panel, Aeronautical Quarterly,
1, 1949, 225-244.
12.5 Budiansky, B. and Seide, P., Compressive Buckling of Simply Supported Plates with Transverse
Stiffeners, NACA TN 1557, April 1948.
12.6 Seide, P. and Stein, M., Compressive Buckling of Simply Supported Plates with Longitudinal
Stiffeners, NACA TN 1825, March 1949.
12.7 Seide, P., The Effect of Longitudinal StiiTeners Located on One Side of a Plate on the Com-
pressive Buckling Stress of the Plate-Stiffener Combination, NACA TN 2873, January 1953.
12.8 Barbre, R., Beulspannungen von Rechteckplatten mit Uingssteifen bei gleichmassiger Druck-
beanspruchung, Der Bauingenieur, 17, 1936, 268-273.
12.9 Chwalla, E. and Novak, A, Theorie der einseitig angeordneten Stegblechsteife, Der Stahlbau,
Beilage zur Zeitschrift Die Bautechnik, 10, (10), May 1937, 73-76.
12.10 Kli.ippel, K. and Scheer, J., Beulwerte ausgesteifter Rechteckplatten, val. I, Verlag W. Ernst u.
Sohn, Berlin, 1960.
12.11 K!Oppel, K. and Moller, H., Beulwerte ausgesteifter Rechteckplatten, val. 2, Verlag W. Ernst u.
Sohn, Berlin, 1968.
12.12 Becker, H., Colao, A., Goldman, R. and Pozerycki, J., Compressive Strength of Ship Hull
Girders, Part II, Stiffened Plates, Ship Structure Committee (U.S.) Report SSC-223, Washington,
D.C., 1971.
12.13 Faulkner, D., Compression Tests on Welded Eccentrically Stiffened Plate Panels, in: Steel Plated
Structures, P.J. Dowling, J.E. Harding and P.A. Frieze, eds., Crosby Lockwood Staples, London,
1977, 581-617.
12.14 Horne, M.R. and Narayanan, R., Ultimate Capacity of Longitudinally Stiffened Plates Used in
Box Girders, Proceedings, Institution of Civil Engineers, Part 2, 61, June 1976, 253-280.
12.15 Murray, N.W, Ultimate Capacity of Stiffened Plates in Compression, in: Plated Structures:
Stability and Strength, R. Narayanan, ed., Applied Science Publishers, London and New York,
1983, 135-163.
12.16 Dowling, P.J., Chatterjee, S., Frieze, P.A. and Moolani, F.M., Experimental and Predicted Col-
lapse Behaviour of Rectangular Steel Box Girders, in: Proceedings, International Conference
on Steel Box Girder Bridges, Institute of Civil Engineers, London, February 13-14, 1973, 77-
94.
12.17 Barbre, R., Schmidt, H. and Riemann, S., Traglastversuche an langsgestauchten, durch Wulst-
fiachstahle versteiften Blechen mit freien Langsrandem, Der Stahlbau, 51, (II), Nov. 1982,
321-332.
12.18 Scheer, J. and Vayas, J., Traglastversuche an langsgestauchten, versteiften Rechteckplatten mit
allseitiger Lagerung, Der Stahlbau, 52, (3), March 1983, 78-84.
12.19 Danielson, D.A., Cricelli, A.S., Frenzen, C.L. and Vasudevan, N., Buckling of Stiffened Plates
Under Axial Compression and Lateral Pressure, International Journal of Solids and Structures,
30, (4), 1993, 545-551.
12.20 Hoon, K.H., Rhodes, J. and Seah, L.K., Tests on Intermediately Stiffened Plate Elements and
Beam Compression Elements, Thin-Walled Structures, 16, 1993, 111-143.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

12.21 PflUger, A., Zum Beulproblem der anisotropischen Rechteckplatte, Ingenieur Archiv, 16, 1947,
111-121.
12.22 Lekhnitskii, S.G., Anisotropic Plates, Gordon & Breach Science Publishers, New York, 1968
(translated from the second Russian edition).
12.23 Ambartsumyan, S.A., Theory of Anisotropic Plates Nauka, Moscow, 1967 (in Russian).
12.24 Troitsky, M.S., Stiffened Plates: Bending Stability and Vibrations, Elsevier, Amsterdam, New
York, 1976.
12.25 Falconer, B.H. and Chapman, J.C., Compressive Buckling of Stiffened Plates, The Engineer,
88, June 5, 1953, 789-791. June 12, 822-825.
12.26 Pelikan, W. and Esslinger, M., Die Stahlfahrbahn-Berechnung und Konstruktion, M.A.N. For-
schungsheft, (7), 1957, 114-123.
12.27 Williams, D.G., Some Examples of the Elastic Behaviour of Initially Deformed Bridge Panels,
Civil Engineering and Public Works Review, 1971, 1107-1112.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
950 Stiffened Plates

12.28 Libove, C. and Hubka. R.E., Elastic Constants for Corrugated-Core Sandwich Plates, NACA
TN 2289, 1951.
12.29 Lau, J.H., Stiffness of Corrugated Plate, ASCE Journal of Engineering Mechanics Division, 17,
(EMl), 1981, 271-275.
12.30 Wittrick, W.H., A Unified Approach to the Initial Buckling of Stiffened Panels in Compression,
Aeronautical Quarterly, 19, 1968, 265-283.
12.31 Williams, F.W., Computation of Natural Frequencies and Initial Buckling Stresses of Prismatic
Plate Assemblies, Journal of Sound and Vibration, 21, 1972, 87-106.
12.32 Williams, F.W., Wittrick, W.H. and Plank, R.J., Critical Buckling Loads of Some Prismatic
Plate Assemblies, in: Buckling of Structures, Proceedings of IUTAM Symposium, Harvard 1974,
B. Budiansky, ed., Springer-Verlag, Berlin, New York, 1976, 17-26.
12.33 Wittrick, W.H. and Williams, F.W., Buckling and Vibration of Anisotropic or Isotropic Plate
Assemblies under Combined Loadings, International Journal of Mechanical Sciences, 16, 1974,
209-239.
12.34 Davies, G.A.O., W.H. Wittric-A Short Biography, in: Aspects of the Analysis of Plate Struc-
tures, D.J. Dawe, R.W. Horsington, A.G. Kamtekar and G.H. Little, eds., Clarendon Press,
Oxford, 1985, XV-XIX.
12.35 Tvergaard, V., Imperfection Sensitivity of a Wide Integrally Stiffened Panel under Compression,
International Journal of Solids and Structures, 9, 1973, 177-192.
12.36 Tvergaard, V., Influence of Postbuckling Behavior in Optimum Design of Stiffened Panels,
International Journal of Solids and Structures, 9, 1973, 1519-1534.
12.37 Koiter, W.T. and Pignataro, M., An Alternative Approach to the Interaction Between Local and
Overall Buckling in Stiffened Panels, in: Buckling of Structures, Proceedings of IUTAM Sym-
posium, Harvard 1974, B. Budiansky. ed., Springer-Verlag, Berlin, New York, 1976, 133-148.
12.38 Van der Neut, A., Mode Interaction with Stiffened Panels, in: Buckling of Structures, Proceed-
ings of IUTAM Symposium, Harvard 1974, B. Budiansky, eel., Springer-Verlag. Berlin, New
York, 1976, 117-132.
12.39 Tvergaard, V. and Needleman, A., Mode Interaction in an Eccentrically Stiffened Elastic-Plastic
Panel Under Compression, in: Buckling of Structures, Proceedings of IUTAM Symposium,
Harvard 1974, B. Budiansky, ed., Springer-Verlag, Berlin, New York, 1976, 160--171.
12.40 Van der Neut, A., Overall Buckling of Z-Stiffened Panels in Compression, in Collapse: The
Buckling of Structures in Theory and Practice, IUTAM Symposium, J.M.T. Thompson and
G.W. Hunt, eds., Cambridge University Press, 1983, 259-268.
12.41 Sridharan, S. and Peng, M.-H., Performance of Axially Compressed Stiffened Panels, Inter-
national Journal of Solids and Structures, 25, (8), 1989, 879-899.
12.42 Thompson, J.M.T., Tulk, J.D. and Walker, A.C., An Experimental Study of Imperfection-
Sensitivity in the Interactive Buckling of Stiffened Plates, in: Buckling of Structures, Proceed-
ings of IUTAM Symposium, Harvard 1974, B. Budiansky, ed., Springer-Verlag., Berlin, New
York, 1976, 149-159.
12.43 Murray, N.W., Analysis and Design of Stiffened Plates for Collapse Load, Structural Engineer,
53, 1975, 153-158.
12.44 Crisfield, M.A., Full Range Analysis of Steel Plates and Stiffened Plating under Uniaxial Com-
pression, in: Proceedings of the Institution of Civil Engineers, Part 2, 1975, 595--624.
12.45 Kotanchik, J.N., Weinberger R.A., Zender, G.W. and Neff, J., Compressive Strength of Flat
Panels with Z- and Hat-Section Stiffeners, NACA ARR L4F01, 1944.
12.46 Schuette, E.H., Barab, S. and McCracken, H.L., Compressive Strength of 24 S-T Aluminum-
Alloy Flat Panels with Longitudinal Formed Hat-Section Stiffeners, NACA TN 1157, 1946.
12.47 Hickman, W.A. and Dow, N.F., Data on the Compressive Strength of75 S-T6 Aluminum-Alloy
Flat Panels with Longitudinal Extrud<~d Z-Section Stiffeners, NACA TN 1829, March 1949.
12.48 Hickman, W.A. and Dow, N.F., Data on the Compressive Strength of 75 S-T6 Aluminum-Alloy
Flat Panels Having Small, Thin, Widely Spaced, Longitudinal Extruded Z-Section Stiffeners,
NACA TN 1978, Nov. 1949.
12.49 Walker, A. C. and Elsharkawi, K., Small Scale Model Studies of the Buckling of Welded Thin
Plate Structures, in: Proceedings of International Conference on the Behaviour of Slender Struc-
tures, The City University, London, 1977.
12.50 Elsharkawi, K. and Walker, A.C., Buckling of Multi-Span Steel Stiffened Plate Panels, in:
Stability Problems in Engineering Structures and Components, T.H. Richards and P. Stanley,
eds., Applied Science Publisher, London, 1978, 297-313.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 951

12.51 Elsharkawi, K. and Walker, A.C., Buckling of Multibay Stiffened Plate Panels, ASCE Journal
of the Structural Div., 106, (ST8), August 1980, 1695-1714.
12.52 Vasta, J., Lessons Learned from Full-Scale Ship Structural Tests, Transactions, U.S. Society of
Naval Architects and Marine Engineers (SNAME), 66, 1958, 165-203.
12.53 Clarkson, J. and Wallace, G., Transverse Strength of a Large Steel Frigate Model, Transactions
of the Royal institution of Naval Architects (R/NA), 109, 1967, 447-486.
12.54 Home, M.R. and Narayanan, R., Ultimate Strength of Stiffened Panels under Uniaxial Com-
pression, in Steel Plated Structures, P.J. Dowling, J.E. Harding and P.E. Frieze, eds., Crosby
Lockwood Staples, London, 1977, 1-23.
12.55 Merrison Committee, Inquiry in the Basis of Design and Erection of Steel Box-Girder Bridges,
Report of the Committee, Appendix I: Interim Design and Workmanship Rules, Her Majesty's
Stationery Office, London, 1973.
12.56 Horne, M.R. and Narayanan, R., The Influence of Rectification Proceedings on the Strength of
Welded Steel Stiffened Plates, Simon Engineering Laboratories, University of Manchester,
March 1976.
12.57 Smith, C.S. and Kirkwood, W., Influence of Initial Deformations and Residual Stresses on
Inelastic Flexural Buckling of Stiffened Plates and Shells, in: Steel Plated Structures, P.J. Dowl-
ing, J.E. Harding and P.A. Frieze, eds., Crosby Lockwood Staples, London, 1977, 838-864.
12.58 Smith, C.S., The Testing of Marine Structures, in: The Future of Structural Testing, Compu-
tational Mechanics Publications, Southampton, Boston, and McGraw-Hill, New York, 1990,
127-157.
12.59 Smith, C.S., Personal Communication to J. Singer, August 1988.
12.60 Smith, C.S., Bending, Buckling and Vibration of Orthotropic Plate-Beam Structures, Journal
of Ship Research, 12, (4), Dec. 1968, 249-268.
12.61 Smith, C.S., Elastic Buckling and Beam-Column Behaviour of Ship Grillages, Transactions
RINA, 110, Jan. 1968, 127-147.
12.62 Clarke, J.D. and Swan, J.W., Interframe Buckling of Aluminium Alloy Stiffened Plating (UL),
Report AMTE(S) R85104, Admiralty Research Establishment, Dumfermline, Scotland, Oct.
1985.
12.63 Dow, R.S., N106C: A Computer Program for Elasto-Plastic Large Deflection Buckling and
Post-Buckling Behaviour of Plane Frames and Stiffened Panels, Report AMTE(S) R80726,
Admiralty Research Establishment, Dumfermline, Scotland, July 1980.
12.64 Brown, J.C., A Laser Based Distortion Measuring System, Report AMTE(S) TM82450, Ad-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

miralty Research Establishment, Dumfermline, Scotland, August 1982.


12.65 Scheer, J. and Vayas, J., Traglastversuche an liingsgestauchten und schubbeanspruchten, ver-
steiften Rechteckplatten, Der Stahlbau, 52, (7), 1983, 207-213.
12.66 Scheer, J. and Plumeyer, K., Traglastversuche an liingsversteiften Rechteckplatten mit konstruk-
tiv bedingten Storungen unter Liingsdruck, Der Bauingenieur, 58, 1983, 387-394.
12.67 Scheer, J. and Gri.iter, A., Traglastversuche an liingsgestauchten, mit trapezfOrmigen Hohlrippen
versteiften Rechteckplatten mit freien Liingsriindem, Der Stahlbau, 53, (9), 1984, 275-280.
12.68 Scheer, J., Gemeinschaftsprogramm "Plattenbeulversuche" des Deutschen Ausschusses fi.ir
Stahlbau, Der Bauingenieur, 58, 1983, 375-379.
12.69 Ghoneim, G.A.M. and Glockner, P.G., Experimental Investigation of an Orthotropic Steel
Bridge Deck up to Ultimate Loads, Research Report CE 76-3, Department of Civil Engineering,
University of Calgary, Alberta, Canada, January 1976.
12.70 Desmond, T.P., Pekoz, T. and Winter, G., Intermediate Stiffeners for Thin-Walled Members,
ASCE Journal of the Structural Division, 107, (ST4), 1981, 627-648.
12.71 Mofflin, D.S, The Strength of Single Stiffener Panels, University of Cambridge, Department of
Engineering, Report CUED/D-Struct/TR.104, 1983.
12.72 Frey, P.M., Buckling Tests on One-Sided Single-Stiffener Panels, University of Cambridge,
Department of Engineering, Report, CUED/D-Struct/TR.ll3, 1985.
12.73 Ghavami, K., The Collapse of Continuously Welded Stiffened Plates Subjected to Uniaxial
Compression Load, in Inelastic Behaviour of Plates and Shells, IUTAM Symposium, Rio de
Janeiro, 1985, L. Bevilacqua, R. Feijoo and R. Valid, eds., Springer-Verlag, Berlin, Heidelberg,
1986.
12.74 Sanbongi, S., Compressive Strength of Flat Stiffened Panels, Japan National Aerospace Labo-
ratory TR-624, 1980 (in Japanese).
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
952 Stiffened Plates

12.75 Holt, S.C., Structures Engineering, Boeing Commercial Airplane Company, Seattle, personal
correspondence to J. Singer, May 2, 1995.
12.76 Waibel, P., Abt. Strukturversuch, Dornier Luftfahrt GmbH Friedrichschafen, Fairchild-Dornier,
Personal Communication to J. Singer, April 10, 1994.
12.77 Haack, C., Airbus Division, Daimler-Benz Aerospace, Personal Communication to J. Singer,
Feb. 16, 1995.
12.78 Oesmann, W., Deutsche Aerospace, Airbus, Technical Note EVll-3470/94, 1994.
12.79 Hoff, N.J. and Mautner, S.E., Buckling of Sandwich-Type Panels, Journal of the Aeronautical
Sciences, 12, (3), July 1945, 285-297.
12.80 Bijlaard, P.P., Stability of Sandwich Plates, Journal of the Aeronautical Sciences, 16, (9), Sept.
1949, 573-574.
12.81 Hoff, N.J., Bending and Buckling of Rectangular Sandwich Plates, NACA TN 2225, Nov. 1950.
12.82 March, H.W., Sandwich Construction in the Elastic Range, in: Symposium on Structural Sand-
wich Construction, ASTM Special Technical Publication No. 118, 1952, 32-45.
12.83 Plantema, F.J., Sandwich Construction, John Wiley & Sons, New York, 1966.
12.84 U.S. Department of Defense, Structural Sandwich Composites, MIL-HDBK-23, December 30,
1968.
12.85 Sullins, R.T., Smith, G.W. and Spier, E.E., Manual for Structural Stability Analysis of Sandwich
Plates and Shells, NASA Contractor Report, CR-1457, Dec. 1969.
12.86 Allen, H. G., Analysis and Design of Structural Sandwich Panels, Pergamon Press, Oxford, 1969.
12.87 Stamm, K. and Witte H., Sandwichkonstruktionen. Berechnung, Fertigung. Ausfiihrung,
Springer-Verlag, Vienna, 1974.
12.88 Wiedemann, J., Leichtbau, Springer-Verlag, Vienna, vol. 1, Elemente, 1986, vol. 2, Konstruk-
tion, 1989.
12.89 Leissa, A.W., A Review of Laminated Composite Plate Buckling, Applied Mechanics Review,
40, (5), May 1987, 575-591.
12.90 Olsson, K.A. and Reichard, R., eds., Sandwich Constructions 1, Proceedings of the 1st Inter-
national Conference, Stockholm, June 1989, EMAS, Solihull, West Midlands, U.K., 1989.
12.91 Weissman-Berman, D. and Olsson, K.A., eels., Sandwich Constructions 2, Proceedings of the
2nd International Conference, Gainesville, Florida, July 1992, EMAS, Solihull, West Midlands,
U.K., 1992.
12.92 Allen, H.G., eel., Sandwich Construction 3, Proceedings of the 3rd International Conference on
Sandwich Construction, Southampton. September 1995, EMAS, Warley, West Midlands, U.K.,
1996.
12.93 Harris, B.J. and Crisman, W.C., Face-Wrinkling Mode of Buckling of Sandwich Panels, Journal
of the Engineering Mechanics Division, ASCE, 91, (EM3), June 1965, 93-111.
12.94 Harris, B.J. and Nordby, G.M., Local Failure of Plastic-Foam Core Sandwich Panels, Journal
of the Structural Division, ASCE, 95, (ST4), April 1969, 585-610.
12.95 Pearce, T.R.A. and Webber, J.P.H .. Experimental Buckling Loads of Sandwich Panels with
Carbon Fibre Faceplates, Aeronautical Quarterly, 24, (4), Nov. 1973, 295-312.
12.96 Gamziukas, V. and Samuelson, L.A., Buckling of Sandwich Panels Subjected to Axial Com-
pression, Shear Forces and Lateral Pressure, Part I: Experimental Investigation, The Aeronau-
tical Research Institute of Sweden (FFA), Report 132, Oct. 1977.
12.97 Jonnson, A., General Instability of Sandwich Panels with a Soft Core, Subjected to an In-Plane
Axial Force. Comparison with Test Results, Saab Report KHU-0-2145R, 1963 (in Swedish).
12.98 Norris, P., Montague, P. and Tau, K.H., All-Steel Structural Panel to Carry Lateral Load: Ex-
perimental and Theoretical Behaviour. Structural Engineering, 67, (9), 1989, 167-176.
12.99 Oikawa, M., Minamida, K., Goto, N. and Teudo, M., Development of All-Laser-Welded Hon-
eycomb Structures for High-Speed Civil Transports, in: Proceedings of the Laser Materials
Processing Conference, JCALEO '93, Orlando, Florida, 1993.
12.100 Tuhkuri, J., Sandwich Structures under Ice Loadings-Theoretical and Experimental Investi-
gations, Marine Structures, 9, (2), 1996, 259-280.
12.101 Laine, C. and Rio, G., Buckling of Sandwich Panels Used in Shipbuilding: Experimental and
Theoretical Approach, in: Sandwich Construction 3, Proceedings on the 3rd International Con-
ference on Sandwich Construction, Southampton, Sept. 1995, H.G. Allen, ed., EMAS, Warley,
West Midlands, U.K., 1996, 243-252.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 953

12.102 Chen, Q., Zimmerman, T.J.E., DeGeer, D.O. and Kennedy, B.W., Strength and Stability Testing
of Stiffened Plate Components, Ship Structure Committee (U.S.A.) Report SSC-399, 1997.
12.103 Horst, P., Buckling Behavior of Large GLARE Fuselage Shells Under Compression and Shear
Loading, in: /CAS Proceedings 1996, Proceeding of the 20th Congress of the International
Council of Aeronautical Sciences, Sorrento, Italy, September 8-13, 1996, AIAA, Reston, Va.,
1996, 2699-2708.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
13
Stiffened Shells

13.1 Global and Local Buckling of Stiffened Shells


13.1.1 Introduction
The nearly universal semimonocoque aeroplane shell structure of the thirties was typified by
a very thin skin, which often buckled in flight, supported by much stiffer longitudinal and
transverse reinforcing elements (see for example [9.48]). The structural element that buckled
was therefore a plate, or an unstiffened shell segment, supported by relatively rigid stiffeners.
Buckling was primarily local buckling, and the final strength of the structure was then es-
timated with the aid of the concepts of effective width and tension field discussed in Chapter
8. With increase in speed of flight, skins became thicker and stiffeners closer, till skin and
reinforcements were of similar rigidity and interacted during buckling. The buckling behavior
of the modern aerospace structure differs, therefore, from that of the earlier semimonocoque
in that it is primarily not a local phenomenon but involves the entire structure, being pre-
dominantly global buckling or general instability.
The efficient aerospace shell structure today is a closely stiffened shell, the stiffening being
the usual stringers and rings, sandwich construction, corrugated reinforcement, composite
material stiffening, integrally machined 4SO waffle stiffeners, or integrally machined stringers
and rings. The last type of construction, integrally stiffened shells, has gained much prom-
inence in aircraft, missiles and space launchers in recent years. Considerable effort has there-
fore been devoted to buckling analysis and experimental studies of integrally stiffened shells
and other closely stiffened shells, primarily to cylindrical shells, the most commonly used
type of shell (summarized for example in [13.1]-[13.4], [9.212], Sections 3.5 and 3.6 of
[9.74], and Chapters 4-6 of [4.48]). In recent years, closely stiffened shells, and in particular
closely stiffened cylindrical shells, have also been widely employed in marine and civil
engineering structures, leading to even more extensive theoretical and experimental investi-
gations (see for example [13.5]-[13.9], [1.31], Part III of [6.14] and Sections 14.4 and 14.5
of [6.3]).
Hence, as in stiffened plates, discussed in the previous chapter, a stiffened shell structure
can fail either by local buckling of the curved panels between the stiffeners or by general
instability involving the panels and their reinforcing members. Since optimization of shells,
in particular that of cylindrical and conical shells, for elastic and inelastic stability usually
leads to closely stiffened shells (see for example [13.1]-[13.4], [13.10], [13.11] and [9.212]),
they have been the focus of most buckling and postbuckling studies of stiffened shells.

13.1.2 Closely Stiffened Shells


For many years the usual approach to the stability analysis of a stiffened shell was to replace
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

it by an equivalent orthotropic shell (see for example [13.10]-[13.15] and [2.39]). Such an
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
956 Stiffened Shells

approach, however, did not permit taking into account the eccentricity of stitTeners (whether
they are on the outside or inside of the -;hell), whose etTect had already been observed in
the thirties and forties ([2.48] and [13.16]). As stiffeners became heavier, in particular in the
large space launch vehicles, designers became aware of the importance of these eccentricity
effects and their influence was extensively studied.
Most of these investigations dealt with stiffened cylindrical shells, the most common type
of shell used in practice, though some considered conical and spherical shells. The eccen-
tricity effects for stringer- and ring-stiffened cylindrical shells have been analyzed and given
a physical explanation, and comprehensive parametric studies have been carried out (see
[2.41], [13.17], and [13.19]). Outside stiffeners have usually been found to stiffen the shell
more than inside ones, but the opposite may be true, depending on the loading and shell
geometry, and in a certain range of shell geometries an inversion of the eccentricity effect
occurs, which is most pronounced in the case of ring-stiffened shells under external pressure
or torsion (see [13.2], [13.17] and [13.18]).
"Smeared" stiffener theory, in which the stiffeners are "smeared," or distributed, over the
entire shell, in a manner that takes into account the eccentricity of stiffeners, has been found
to be a satisfactory approach for closely stiffened shells that fail by general instability, since
the effect of the discreteness of stitieners is usually negligible, as will be indicated below.
The linear smeared stiffener theory employed in most of the studies (like [13.17]-[13.26]),
which considers perfect shells, is, however, meaningful only if the extent of its adequacy
has been determined either experimentally or by advanced numerical solutions taking the
nonlinear effects caused by the edge restraints into account. Hence considerable experimental
effort was devoted to assessing the bounds and extent of applicability of linear theory. The
smeared stiffener linear theory and some of the experiments that assess and bound its ap-
plicability will be briefly discussed in the following sections.

13.1.3 Linear Smeared Stiffener Theory


When "classical" linear theory, using a membrane prebuckling solution, can be assumed to
predict buckling loads adequately (the bounds of validity of this assumption and the inter-
pretation of "adequate" that emerge from the experimental results are discussed below), the
general instability of closely stiffened cylindrical shells can be analyzed with a simple
smeared stiffener theory, derived at the Technion in 1963 [9.299]. Because of its simplicity,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

this theory has been used by many investigators for cylindrical shells (see for example [3.12],
[13.21], [13.22]-[13.24]) and has also been applied to conical and spherical shells (for ex-
ample [9.278], [9.290], [ 13.25], [13.26] and [ 13.159]). Its results also correlate well with
those obtained by the earlier theory of van der Neut [ 13.16]. The details of the linear smeared
stiffener analysis, which assumes a membrane prebuckling state, were presented for the case
of axial compression in [2.41], for external pressure in [9.299] and [13.17], for torsion in
[13.18] and for bending in [13.23]. (A short summary is presented in Chapters 2 and 3.)
In outline, the analysis employs the linear Donnell stability equations, into which the
smeared rings and stringers are introduced through the force and moment expressions, all
equations being represented in terms of displacements. In the mathematical model the stiff-
eners are smeared to form a cut layer. For example, external rings are replaced by a layer
of many parallel rings that cover the whole exterior of the shell and touch each other but
are not connected with each other. The model slightly underestimates the torsional stiffness,
but this is preferable to its overestimation, which occurs when the stifTeners are replaced by
a concentric orthotropic continuous layer, as in [3.11] of Volume 1.
The main assumptions are therefore (as in [9.299], [2.41], [ 13.17], etc.), using the notation
of Figure 13.1:

1. The stiffeners are distributed over the whole surface of the shell.
2. The normal strains e,(z) and e,r(z) vary linearly in the stiffener as well as in the sheet.
Copyright Wiley
The
Provided by IHS Markit under normal strains in the stiffener and inLicensee=McDermott
license with WILEY the sheet are Inc -equal at their
India/8215328006, point
User=G, of contact.
Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Global and Local Buckling of Stiffened Shells 957

x•

Figure 13.1 Notation for stiffened cylindrical shells

3. The shear membrane force N"" is carried entirely by the sheet.


4. The torsional rigidity of the stiffener cross-section is added to that of the sheet (the
actual increase in torsional rigidity is larger than that assumed).

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
The middle surface of the shell is chosen as reference line, and the expressions for forces
and moments in terms of displacements are (see also [2.41]):
N, = [Eh/(1 v 2 )][u,Jl + f.LJ) + v(v,,f,- w) - x,w,H)
2
N"' = [Eh/(1 v )][(v,,1, - w)(l + fL2) + vu,, - X2W•d"t,]

N"l> = N,~>' = [Eh/2(1 + v)](u,, + v,d,)

M, = -(DIR)[w,,,(l + 7]01 ) + vw,J><P- ?,u,J


M," = -DIR)[w,d"t'(l + 7JoJ + vw," - ?2 (v,, - w)]

M,J, = +(D/R)[(l - v) + 1), 1]w,Ht,

MJ>, = -(D/R)[(l - v) + 1Jclw,,J, (13.1)


where h is here the thickness of the shell, fL 1 , fL 2 , 7]01 , 7]02 , 7], 1 and 7], 2 are the changes in
stiffnesses due to stringers and rings and x,, x2 , ( 1 and ( 2 are the changes in stiffnesses
caused by the eccentricities of the stringers and rings. u, v and w are the additional displace-
ments during buckling, and they are nondimensional, the physical displacements having been
divided by the radius of the shell. (Notice that thus x and cfy are the nondimensional axial
and circumferential coordinates, respectively.) As usual, D = Eh'l 12(1 - v 2 ), Pis the axial
compressive load and p the external pressure.
The changes in stiffnesses arc defined as follows:
The increase in effective cross-sectional area of the shell due to stringers and frames, re-
spectively,
fL 1 = (1 - v 2 )(A,E,IbhE)}
(13.2)
fL 2 = (1 - v 2 )(A 2 E 2 I ahE)

The changes in extensional stiffness caused by the eccentricities of stringers and frames,
Copyright Wiley
respectively,
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
958 Stiffened Shells

2
X1 = (1 - v )(E 1A 1e 11EbhR)}
(13.3)
X2 = (1 - v 2 )(E0 2 e 2 1EahR)

The increases in bending and twisting stiffness of the shell due to stringers and frames,
respectively,
7Jo 1 = E/JbD

7Jo2 = Ei21 aD
(13.4)

The changes in bending stiffness caused by the eccentricities of stringers and frames, re-
spectively,

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
(13.5)

In these changes of stiffnesses, in Eqs. (13.2)-(13.5), it is assumed that the stiffeners arc
closely spaced and therefore the entire shell is active. One could, however, analyze in a
similar manner a shell with wider stiffener spacing where only part of the shell can be
considered active (the effective width), by multiplying the terms in the above expressions
relating to the shell proper by an appropriate coefficient.
Substitution of Eqs. (13.1) into the Donnell stability equations transforms them into three
equations in terms of displacements:
[Eh/(1 - v 2 )]{(1 + f.L 1)u,,, + [(1 - v)/2]u,""" + [1 + v)/2]u,,<f>- X 1W,,_,- vw,,} = 0

[Eh/(1 - v 2 )]{[(1 + v)/2]u,"" + (1 + f.Lc}U"M + [(l - v)/2]u,,, -(I + f.L 2)w,,b- X 2W,,1,bdJ

=0
(-D/R){[ 1(-u,,J + [ 2 (2w,"""- u, 4,<t><t>) + (1 + 7]01 )w,,,,, + (2 + 7)11 + 7] 12 )w,,, 1
,d,

+ (l + 7J02}>v,,P•I><t><t> + 12(R/ hff(l + f.L 2 )(w - u,,b) - vu . .] + (PRI nD)(w,j2)


+ (pR'ID)[(w,n/2) + w,</>d,]) = 0 (13.6)
For classical simple supports (SS3: w = M, = N, - u = 0) the same displacements as
for isotropic cylinders (see [2.41]),
u = A, cos m{3x sin ncp
u = B, sin m{3x cos ncp (13.7)
w = C, sin m{3x sin ncp
solve also Eqs. (13.6). Notice that m,n arc integers denoting the number of half-waves in
the axial and the number of full waves in the circumferential direction, respectively, and
{3 = (1rRI L). For classical clamped ends (RF2: w = w,, = u = N,<f, = 0) a closed form
solution is not possible, but the same displacements and procedure (applying the Galerkin
method to the third equation) as for unstiffened shells could be employed [2.41]. In the late
sixties and early seventies, improved solution procedures that consider different in-plane
boundary conditions were also developed (see for example [13.27]-[13.32]). Computer pro-
grams for shells of revolution with different types of stiffening, based on finite differences,
were also generated at that time (see for example [13.33]).
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Global and Local Buckling of Stiffened Shells 959

The results of linear smeared stiffener theory were summarized (in [13.1], [2.41], [13.17]
and [ 13.18]) with conclusions about the effectiveness of stiffeners and the relative importance
of shell and stiffener geometric parameters. The shell geometry is usually represented by
one parameter, the Batdorf parameter Z = (1 - v2 ) 112 (U I Rh), and the stiffener parameters
are spacing, shape, cross-sectional area and eccentricity. Spacing is determined by local
buckling and is hence outside the realm of the smeared stiffener theory, whereas the influence
of the other parameters is discussed in detail in those references. Here only some of the
main results are reiterated.
Among the parameters determining the efficiency of stiffeners, there is strong interde-
pendence between the effects of stiffener and shell geometry and weaker interdependence
among the various stiffener geometric parameters. The cross-sectional area of the stiffeners,
or the corresponding nondimensional area ratio, is usually the prime geometric parameter
determining their effectiveness. It is, however, also the one that affects the weight directly,
and hence structural efficiency considerations may dictate relatively weak stiffeners.
As an example, consider Figure 13.2 (reproduced from [13.17]) in which the influence of
ring area on the structural efficiency in buckling of a ring-stiffened cylindrical shell under
hydrostatic pressure was studied. Though ring-stiffening is very effective in this case, it is
apparent that the heavy rings (A 2 / ah) = 0.8 are never the most efficient ones. For small Z
the weakest rings are best, and for large Z there appears to be an optimum area ratio.
It may be noted that the cross-section of the stiffeners is usually constant for the whole
shell, but noticeable weight savings of 10-20 percent were shown to be possible by variation
of the stiffener cross-section along the shell (see [13.2] and [13.34]). Also, experimentally
the stiffener cross-sectional area, or the corresponding nondimensional area ratio, has been
shown to be the prime parameter in the determination of the applicability of linear theory
(which assumes a membrane prebuckling state), which is discussed below.

13. 1.4 Eccentricity and Discreteness Effects


The eccentricity of the stiffeners has two effects on the behavior of the shell. One is an
increase in the effective moment of inertia of the stiffener that is independent of the position
of the stiffener; and the other, the eccentricity effect proper, is the result of the coupling of
membrane forces with bending moments, which is determined by the position of the stiff-
eners. This eccentricity effect varies with different types of loading and geometries. It is
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

made up of two opposing contributions: the primary effect, the change in actual bending
stiffness of the stiffener due to the membrane forces in the shell; and the secondary effect,

Figure 13.2 Influence of ring area on structural efficiency of closely ring-stiffened cylindrical shells
under hydrostatic pressure (from [13.17])

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
960 Stiffened Shells

the change in actual extensional stiffness of the shell due to bending moments in the stiff-
eners. As a result of the interplay of these opposing contributions, the eccentricity effect
depends very strongly on the geometry of the shell, which determines which bending mo-
ments and membrane forces carry the brunt of the load.
Under axial compression, stringers are the most effective stiffeners. Figure 13.3 (repro-
duced from [2.41]) shows the structural efficiencies of stringer-stiffened cylindrical shells
under axial compression. For a typical stringer geometry, the ratio of buckling load of stiff-
ened shells to that of equivalent-weight unstiffened shells is plotted versus the Batdorf pa-
rameter Z = (1 - v 2 C12 (U I Rh) for outside, inside and centrally placed stringers and for
classical simple supports and clamped ends. The effectiveness of the stringers is large in the
low Z range but appears to diminish rapidly for longer and thinner shells. However, the
curves in Figure 13.3 are conservative since the equivalent unstiffened shells were also
computed with linear theory, which is unrealistic for them. Reevaluation after application of
empirical corrections to the buckling loads of the unstiffened shells yielded much higher
efficiencies and showed that at least outside stringers are always more effective than equiv-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

alent thickening of the shell.


The eccentricity effect is very pronounced in axially compressed stringer-stiffened shells
and depends strongly on the geometry of the shell, as can be clearly seen in Figure 13.4
(reproduced from [2.41 ]), in which ( P""' I pin) is plotted against Z. The designer may note
that the eccentricity effect has a pronounced maximum that occurs for values of Z common
in aerospace practice. This maximum, the shape of the curve and the inversion of the ec-
centricity effect, which occurs here for extremely short shells, are the results of the interplay
between the opposing primary and secondary contributions that make up the total effect. For
clamped ends the eccentricity effects are similar, except for a shift in the maximum to a
higher Z.
Rings are much less effective than stringers in stiffening of cylindrical shells, subjected
to axial compression, when alone. However, they may increase the buckling load consider-
ably when they act together with stringers (see [13.2], [13.22] and [13.33]). Furthermore,
the behavior of ring-stiffened shells under axial compression is also important, since shells
primarily designed to withstand lateral pressure (which will be predominantly ring-stiffened)
may be subjected to axial loads under certain conditions. Ring stiffening was therefore dis-
cussed in detail in [ 13.1]. For example, in the case of torsion loading, rings were found to
be more efficient than stringers, except for short shells, and large eccentricity effects were
found for rings.
The discreteness effect of the stiffeners was investigated by a linear discrete stiffener
theory (see [13.35]-[13.38]). Instead of being smeared, the stiffeners are considered as linear

P
9

r----,,--,--~,-.--~-,--~,I-A-,/b_h_~-o.-5~,='~'- J
I ?-buckling load for stiffened shell !
f--+-\1~--t---1~-±c-·~..J,-----+1 Pun~ ~q.-buckling load for I
I equivalently thickened sh. ell • I

: ! l 1 II '

Figure 13.3 Structural efficiency of closely stringer-stiffened cylindrical shells under axial compression
(from [2.41])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Global and Local Buckling of Stiffened Shells 961

-------------------;

pout 1.6
P'n
1.4

1.2

1.0
1rf 10' 102 10 3 104 10 5
z
Figure 13.4 Variation of eccentricity effect with shell geometry, represented by the Batdorf parameter
Z, in closely stringer-stiffened cylindrical shells under axial compression (from [2.41])

discontinuities represented by the Dirac delta function. The force and moment expressions
of Eq. (13.1) were modified accordingly, and the remainder of the analysis is similar to the
smeared stiffener theory. The representation is, however, satisfactory only as long as the
width of the stiffener is small compared to the spacing.
For buckling under hydrostatic pressure, appreciable load reductions were found for dis-
crete rings, even when the number of rings was not small (see [13.36]). In the case of axial
compression, the discreteness effect of rings could usually be neglected (see [9.212] and
[1.25]). However, if the number of rings in the shell is very close to the number of axial
half-waves in the predicted axisymmetric buckle pattern, a recheck should be made with
discrete stiffener theory. Such a recheck (in [1.25]) confirmed the negligibility of discreteness
effects even when the number of rings was close to the number of axial half-waves. The
only exception was the case in which the rings were strong enough to make local buckling
dominant. The discrete theory then yields buckling loads very close to the local ones, and
when these are considerably lower than the general instability loads, appreciable discreteness
effects appear (see Table 3 of [1.25]). Hence in ring-stiffened shells designed for failure by
general instability, discreteness effects are negligible.
In the case of stringer-stiffened shells the discreteness effect was again found to be neg-
ligible for practical stringer spacings used in aerospace applications, which are usually very
close, while significant discreteness effects are only likely for stringers with very large tor-
sional stiffness in shells with Z < 400 (see [13.37]).

13.1.5 Boundary Effects


In the smeared stiffener analysis outlined in the previous sections, classical simple supports
or clamped ends were assumed. Since the effect of the in-plane boundary conditions on
buckling loads in unstiffened cylindrical shells had been found to be very significant (see
for example [9.65]), their effect was also investigated for stiffened cylindrical shells.
Earlier studies of orthotropic and stiffened shells (for example [1.25], [3.11], [9.235] and
[13.27]) had considered some of the different in-plane boundary conditions, but more com-
plete parametric studies were carried out only in the early seventies (see [13.28]-[13.30]).
These parametric studies of the influence of the in-plane boundary conditions employed the
displacement method, which is essentially an extension of the linear Donnell type theory
outlined in Section 13.1.3, which was originally derived for conical shells ([9.296]).
For ring-stiffened shells under axial compression ([13.28]), the effects of the in-plane
boundary conditions are similar to those observed in isotropic shells, which are well known.
The "weak in shear" boundary conditions SSl and SS2 (with N"" = 0 instead of v = 0),
which rarely occur in practice, reduce the buckling load to about half the classical value,
whereas axial restraint (u = 0 instead of N, = 0) has a negligible effect, yielding practically
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

identical results for SS3 and SS4.


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
962 Stiffened Shells

For stringer-stiffened cylindrical shells under axial compression, however, the effects of
in-plane boundary conditions differ appreciably from those in isotropic shells. The effects
are different for external and internal stringers, and depend also on the stringer geometry. In
general, the axial restraint (u = 0 instead of N, = 0) is the predominant factor, whereas the
"weak in shear" boundary conditions have only a minor influence. Figure 13.5 (reproduced
from [13.29]) shows the variation with the shell geometry parameter Z of the ratios of
buckling loads for typical heavily stringer-stiffened cylindrical shells under axial compres-
sion. Heavy stringers were chosen in this example to emphasize the rather different effects
of in-plane boundary conditions for stringer-stiffened shells.
With internal stringers, axial restraint (u = 0, SS4 or SS2) raises the buckling loads by
about 45 percent and considerably above as the shells become longer and thinner, irrespective
of the tangential boundary conditions. On the other hand, (SS 1I SS3) is approximately unity,
confirming the negligibility of the v condition. With external stringers, axial restraint is less
effective and even ineffective in a certain important range of shell geometries, but the weak-
ening N, 4, = 0 is still practically absent, except for a relatively minor decrease in a limited
range of Z.
For weaker stringers, the importance of the axial restraint diminishes, and as the isotropic
shell is approached, the tangential constraint obviously gains in importance. However, even
for relatively weak stringers, with (A,Ibh) ~ 0.1 and (e 1 /h) ~ ±0.5, the effects are far from
the ones observed in isotropic shells. These pronounced effects of axial constraints appear
for simple supports (with M, = 0). Axial :restraints also affect the buckling loads for clamped
ends (with w,, = 0), but the effects differ slightly. An example of the effect of axial or
rotational restraints on buckling for typical stringer-stiffened cylindrical shells of two lengths
and varying area ratios is presented in Figure 13.6 (reproduced from [13.31]). Comparison
between the fully clamped C4 and SS4 curves in the figure shows that whereas rotational
restraint is effective only for short shells, axial restraint is even more effective for long ones.
The predominant influence of the boundary conditions, and in particular the axial restraint,
on the buckling of stringer-stiffened cylindrical shells, lead to many further studies (see for
example L13.30]-[13.32]). Similar effects were also observed on the lower natural frequen-
cies of vibrations whose mode shapes resembled the buckling modes (see for example
[13.39], [13.40]). This, and the rather complicated effect of the boundary conditions, moti-
vated extensive correlation studies between vibration and buckling at the Technion for two
decades (see for example [13.31], [13.32], [13.40] and [13.41]), aiming at a nondestructive
experimental determination of the actual boundary conditions. The resulting vibration cor-
relation technique is discussed in Chapter 15 and is also summarized in [13.411.
The influence of in-plane boundary conditions was also studied for ring-stiffened shells
under hydrostatic pressure. For weak rings, the primary effect is axial restraint, as in the
case of unstiffened shells (see [9.234]). As the rings become stronger, the effect of the "weak

P/Pss3 I A,J bhl~ 1.44


2 .2 3
(111/bh J~84.84
...------.
/
(SS4/SS3)in ~ /~

- p-'1 SS2iSS31; 0 /;
1,4 -._ :::..:, /'/ISS 2/SS 31 0 ut
~-,,
,,....
I SS4/SS31 0 ut -';./
...........,....
I 551/5531 in''~---- -----~:::::· .-
-----....-.-::::....·-------=--~~::=:--"------

( 551/55 3)0 ·,;\'---·/


0,61------_j__ _ _ _ ___L_ _ _ _ _ _i _ __ _ __ _ J
4
10
1 10 2 10 3 10
z
--`,`,`````,`,````,,``,``,,`-

Figure 13.5 Buckling load ratios for heavily stringer-stiffened cylindrical shells with different in-plane
boundary conditions-variation with shell geometry (from [13.30])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Global and Local Buckling of Stiffened Shells 963

R/h=480,LIR=1 ,Z=458
R/h=480,LIR= 2, Z=1832

~
55 3

c 4
/ 55 3
/
/
/
/ 554
/ /--·----- 553
/ /
/ /

554
553
1. 2

Figure 13.6 Effect of axial and rotational restraints on buckling load ratios (Pj P,, 3 ) of stringer-
stiffened cylindrical shells of two lengths (L/ R = 1 and 2) with varying area ratio (from
[13.31])

in shear" boundary conditions becomes dominant, since then the shell has a relatively smaller
resistance to the axial component of the hydrostatic pressure. For medium stiffening, say
(A 2 / ah) = 0.5 and (e 2 / h) = ± 0.5, the ring-stiffened shell under hydrostatic pressure is
affected by the in-plane boundary condition as an axially compressed isotropic shell (yielding
for SS I and SS2 buckling loads about half the classical ones), except for shells with Z >
1000.
Nonlinear prebuckling deformations represent another boundary effect in stiffened shells
that was studied in [13.1], [13.31]-[13.33] and [13.42], with the conclusion that for the cases
investigated they arc apparently not a major factor in the determination of the buckling load
of stiffened cylindrical shells, except for short stringer-stiffened cylindrical shells. Figure
13.7 (reproduced from [13.32]) shows an example of the variation of the influence of pre-
buckling deformations on the buckling load ratio (PPRE = PPRE/ PMEMB• where PPRE is the
buckling load calculated with consideration of nonlinear prebuckling deformations and
PMEMB is that computed by linear theory), for two types of stringer-stiffened shells, which

illustrates this conclusion. However, whereas the linear theory is in most cases adequate for
parametric and optimization studies, the buckling load of the final design should always be
verified by a more rigorous buckling load calculation including the nonlinear prebuckling
deformations caused by the edge constraints.
Load eccentricity is also a very important boundary effect for stringer-stiffened shells and
has been actively investigated (see for example [13.1], [13.42] and [13.43]). The load ec-
centricity effect can be summarized qualitatively by pointing out that moments which tend
to bend the stringer-stiffened cylindrical shell into a barrel shape (and thus give rise to tensile
prebuckling hoop stresses) increase the buckling load. Hence for external stringers, for ex-
--`,`,`````,`,````,,``,

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
964 Stiffened Shells

1.4
STIFFENED

1.3 - - - - A1lb 1h =0.25.·-!-----\---\-!-----\--\----+- 5 HELLS


--A,Ib 1h =0.50
1.2

lLLJ

g:
Q.r
1.1
1• 0

0.9

0.8

0.7101 2 5 8 :10 2 2 5 8 10 3
z----
Figure 13.7 Effect of nonlinear prebucklin;~ deformations on the buckling load ratio of stringer-
stiffened cylindrical shells-variation with shell geometry (from [13.32])

ample, loading through the mid-surface of the skin will yield higher buckling loads than
application of the load at the stringer centroids, whereas for internal stringers the opposite
occurs. One may add that for medium lengths shells of (LI R) = 1 the effect may amount
to up to 20-40 percent in practical shells.

13.1.6 Adequacy and Bounds of Validity of Linear Smeared Stiffener


Theory
The conclusions about the influence of eccentricity, boundaries and stiffener and shell geo-
metric parameters on the buckling loads of stiffened shells arrived at the previous sections
are meaningful only if the linear smeared stiffener theory predicts the buckling load ade-
quately and if the discreteness of stiffeners has no noticeable effect. As pointed out in Section
13.1.4, the discreteness effects were shown by analysis usually to be negligible, but the
adequacy of linear theory can be conclusively settled only by tests or by extensive nonlinear
calculations (not feasible 30 years ago). Extensive experimental studies were therefore carried
out at the Technion Aerospace Structures Laboratory and other research centers on the buck-
ling of closely spaced stiffened cylindrical and conical shells in the sixties and seventies (see
for example [1.25], [9.212], [9.222], [9.228], [9.149], [9.278], [9.279], [13..4], [13.31],
[13.32], [13.44]-[13.47]).
The applicability of linear theory is conveniently expressed by the ratio of the experimental
buckling load Pcxp to that predicted by linear theory Per• sometimes called "linearity" p =
(Pc,/ Pc,) in stiffened shells, in preference to the term knock-down factor used in unstiffened
shells (because of the large reductions in predicted loads necessary there to correlate them
with the observed ones), since in closely stiffened shells p is usually closer to unity.
For ring-stiffened shells, the ring area ratio (A 2 1ah) was found to be the predominant ring
and shell geometry parameter, and therefore their "linearity" (knock-down factor) is plotted
in Figure 13.8 as a function of (A 2 Iah). Shells of different materials, tested by different
investigators [1.25], [9.212], [13.44], [13.48] and [13.49], are plotted in the figure. Since for
ring-stiffened shells the boundary effects are relatively small, the results from different test
setups fit within a reasonable scatter band. --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Global and Local Buckling of Stiffened Shells 965

1.3[
1.2

~!.I

·.... .·.
.
,· '
.
* • ·-
:•:.,
.-

. .. •
---------·-~---- -·- _._-----+------~--


..
. ' .. . . . 4 •

• STEEL IR•7"J [Ref. 9.212]


• •
A
CALTECH (CLAMPED 606!-T6i[Ref.'.25]
ALMROTH I ALUMINUM ALLOY) [Ref.13.48]
+
••
STEELIR•7"1 [Ref. 13.44]
t STEELIR•S"i[Ref. 13.44]
e 7075-T6(R-5")-AR SHELLS [Ref. 13 44]
• 7075- T 6(Rol20mm)-RO 5HELLS[Ref.13.49]

Figure 13.8 "Linearity" or (knock-down factor) of ring-stiffened cylindrical shells under axial com-
pression as function of the ring area parameter (from [13.31])

For (A;,Iah) > 0.3 there is ± 10~15 percent scatter about a mean p = 0.95 (except one
point that is 17 percent below 0.95), or one can state roughly that p = 0.8. As the ring area
ratio decreases below 0.3, p decreases, at first slowly, but below 0.15 rather rapidly. Hence
applicability of linear theory appears to be bounded here by (A 2 / ah) 2 0.2.
For stringer-stiffened shells, the stringer area ratio (A J bh) appeared to be the predominant
stringer and shell geometry parameter determining their "linearity" (knock-down factor). p
is therefore plotted in Figure 13.9a for simply supported stringer-stiffened cylindrical shells
(p553 ) and in Figure 13.9b for clamped ends (pc 4 ), both as a function of (A,Ibh). Again
shells of different materials and tested by different investigators [1.25], [9.222], [9.228],
[13.21], [13.45], [13.47] and [13.49] are plotted in the figures. The scatter for the stringer-
stiffened shells in Figures 13.9a and 13.9b is larger than that for the ring-stitlened ones in
Figure 13.8, but the trend of the primary dependence on the stiffener area ratio is still evident,
though less clearly defined.
The "linearity" (knock-down factor) of all the shells considered simply supported (Figure
13.9a), except those of [13.47], is above 0.65 even for weak stringers, and for (A,Ibh) >
0.5 all shells have p > 0.72. In some of the shells tested by Katz [13.47], which are the
exceptions here, early local buckling was clearly observed and local buckling may have
occurred also in some of the others of this series. The scatter in Figure 13.9a is usually
about ± 20 percent, and even ± 30 percent in regions where many tests accumulate, about
an average p that rises from p ~ 0.7 for weak stringers to p ~ 0.9 for heavier ones. This
scatter is partly due to incomplete definition of the boundary conditions and can be signif-
icantly reduced by a better determination of the actual boundary conditions with the aid of
the vibration correlation technique, as shown in Chapter 15.
For clamped shells (Figure 13.9b), lower values of "linearity" (knock-down factor), down
to p ~ 0.6, are observed. The relatively low p values in the figure for some of the RO shells
[13.49] with fairly heavy stitlening, (A 1 I bh) ~ 0.5~0.7, are conspicuous. (Bifurcation anal-
ysis with nonlinear prebuckling could probably have achieved better correlation here.)
Though again more realistic boundary conditions, obtained by vibration correlation, would
correct the deviations and improve the "linearity" (see Chapter 15), the values are low, which
may be partly due to additional imperfections introduced by the clamping.
The influence of shell geometry and stiffener spacing in stringer-stitlened cylindrical shells
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,

was also studied in [13.1]. [9.222] and [9.228] but found to be less significant than the
stringer area ratio in determining the bounds of validity of linear smeared stitlener theory.
Experimental results for other loading cases have provided additional support for the
adequacy
Copyright Wiley
of linear theory as a first estimate of the buckling of stitlened cylindrical shells or
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
966 Stiffened Shells

• •
.
0

...
........
.. .. +

+ t'0?"'-TG,(n~l20m<n)(WITHOlJT SHELLS OF e>o.6t.. l [t<ef. 9.228]


~ Sllt.l.. (R=I75mrnl 9.?22]
0 KATZ [Ref.
0 LE:"JKO [Ref.
o 707'J-T6(R-,\20'nrnl-PO SHELLS [Ref 13.49]

(a)
------~·~----~~~L~- __ _! _ _ __I
.3 •4 .s .5 .7 .8
AJbh

~~~-----~~--~-~--- -------------

CJCARD&JONES [Ref.13.Z1]
• CA.UECH (Ref \.25]
o 7075-Tt (R:\20mmJ-RO SHELLS [Rei. 13.49]

(b)

Figure 13.9 "Linearity'' or (knock-down factor) of stringer-stiffened cylindrical shells under axial
compression as function of the stringer area parameter (from [13.31]): (a) simply sup-
ported shells, (b) clamped shells

for their parametric and optimization studies. For example, in some bending and hydrostatic
pressure tests, discussed in [13.1], p > 0.9 for all cases. Tests on ring-stiffened conical shells
[9.149] also support the conclusions arrived at for cylindrical shells.
The adequacy (or good approximation) of linear theory for prediction of the buckling
loads of closely stiffened shells stems from the similarity of the actual buckling modes to
those predicted. Often this similarity holds even under axial compression, contrary to isotro-
pic cylindrical shells, where the experimentally observed buckling modes differ substantially
from the ones predicted by linear theory.
In this section only the results of the buckling tests were presented, as experimental evi-
dence of the adequacy of linear smeared stiffener theory. Some details of their test techniques
are discussed below in Sections 13.2 and 13.3.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

13.2 Model Fabrication for Stiffened Shells


13.2.1 Machined Shells
As pointed out in Chapter 9, Subsection 9.3.5, machining from thicker stock and special
forgings, though also used for isotropic shells of revolution, is primarily employed for stiff-
ened shells. Models of integrally stiffened cylindrical shells are usually machined from thick-
walled tubes, but for thin stiffened shells, representing for example the stringer-stiffened
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Model Fabrication for Stiffened Shells 967

stages of the Saturn V launcher of the Apollo program with (R I h) of the order of 500, an
internal support mandrel is necessary for accurate machining.
At the Technion Aerospace Structures Laboratory in Haifa, extensive tests on thin inte-
grally machined cylindrical and conical shells were performed in the sixties, seventies, eight-
ies and nineties (see for example [1.17], [1.24], [9.149], [9.212], [9.222], [9.228], [9.229],
[9.274], [9.278], [9.279], [13.31], [13.32] and [13.44]), and their fabrication processes are
good examples of the employment of internal mandrels and of a thermal shrink fit by either
cooling or heating.
The earlier integrally ring-stiffened cylindrical shells, with an average (Rih) of 660 (see
[9.206] and [9.212]), were machined from 14-in. diameter AISI 4130 steel alloy drawn tubes
with a 11 4-in. wall thickness. The test series included 18 shells. The mechanical properties
of the material, and in particular E, were measured on 8 specimens cut out from the tubes
before machining (both in the longitudinal and circumferential directions), as well as on 16

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
specimens cut out from the shells after buckling failure. The average values found were those
that appeared in the literature, the average E = 19.4 X 104 kNimm 2 (or 28.5 X 10' ksi)
and the yield stress a, :==: 490 kN I mm 2 (or 71 ksi), which was considerably above the
predicted buckling stresses.
The dimensions of a typical specimen of the test series (MZ-2) are shown in Figure 13.10.
Note that to prevent end moment effects the end rings of the specimens have ridges of width
h (marked with arrows in the figure) that represent a continuation of the shell wall.
In the interest of precision, the machining process was divided into stages. In the first
stage the internal and external surfaces of the tubes were roughly machined. Then the internal
surface was precisely turned to the dimension of the cooled mandrel, on which it was
mounted later for machining of the stiffeners. The dimension of the inside diameter was
chosen to give a medium press fit between the cooled mandrel and the mounted blank. The
blank was then mounted on the special cooled mandrel (see Figure 13.11) made of cast
aluminum with a high silicon content and having the shape of a reservoir with many fins
around its inner surface. The mandrel was fitted with a special pivot. Liquid air poured into
the reservoir of the mandrel cooled it appreciably and as a result its diameter contracted 0.4
mm, enabling the shell to slide onto the mandrel. After returning to room temperature, the
shell sat well on the mandrel and permitted accurate machining. After completion the shell
was removed from the mandrel by another liquid air cooling and a second shell was im-
mediately mounted.
This technique, combined with extreme care in the machining and continuous measure-
ments, resulted in precise specimens in which the deviation of thickness (the most sensitive

.I
~ ~c "' l5'00S
'----·
~l ht
I I t
I

r---- ··- r-n-·- It- ¥


I
II
'I
I.
'I
I!
'i
~L
_ ___t

Figure 13.10 Technion tests on integrally ring-stiffened steel cylindrical shells under axial compres-
Copyright Wiley
sion-dimensions of typical specimen (MZ-2) (from [9.206])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
968 Stiffened Shells

Figure 13.11 Technion lc,ls 011 inlcgrolly s1iffened cylindrical 'hclls- Lhe cooled in ~emal ;uppo11 man-
drel for ; lccl 'hell < (from [9.206])

dimension) of the shell did nol exceed 5 percem of 1hc average in the woN ca~>e and was
U!>Ually within 2.5 percent The accurdcy of the heigh! of 1he rings d was similar 10 that of
lhc thickness. whereas larger deviations occurred in !he wid1h of the ring> c. Differences of
up to 16 pcrccm from lhc nominal width were mea~ured. but these could, a1 worM. lead to
an error of less than 3 pcrcenl in the computed buckli ng load. The dimension!> of !he spec-
imens were carefull y measured befo re each 1es1. the wall thickness m about 300 poi nts for
each shell and the sl iiTcncl' dimensio n at about 200 locations fo r some of the speci mens.
Typical specimens arc shown in Figures 13. 12 ~1ntl 13.13.
In the continuation of lhc lest program [9.2221. ~•ringer-stiffened cyl indrical she lls (sec
Figure 13. 14) were manufactured by a similar process. The firM phase of the >~ringer-stiffened
shell tests employed ~tee! 'pccimens. 50 of 14 in. diameter and 34 of 10 in . diameter. The
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(a)

Figure 13.12 Technion ICSI< on irncgrally s1iffened cylindrical shells-lypical AISI 4110 ''eel ring-
sliffened shell, MZ 12. ICSLed under axial comprc"iun (from [9.206]): (a) before loading,
(b) af1er fni lu•·c. (c) aflcr failure. viewed from 1hc ulhcr side
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Model Fabrication for Stiffened Shells 969

Figure 13.13 Technion tests on integrally stiffened cylindrical shells-buckling of typical AlS1 4130
steel ring-st iffened shell. MZ 18. tested under axial compression with a distance tube.
arresting the axial displacemems when it was three ti mes the linear prebuckling one
(from [9.206])

larger shells were made from 25 CD-4F steel alloy drawn tubes with mechanical properti es
similar to Al$ 14130 (and a slightly higher yield stress av = 549 kN /mm2 • or 78 ksi). whereas
the smaller specimens were made from another alloy steel in a softer condition (with the
same average measured E as the larger shells but a smaller yield stress (a,.= 421 kN/nun 2,
or 60 ksi).
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

~ .,..>, Y
c.
SEC liON 8·8
• ENLARGED

Figure 13.14 Technion tests on integrally stringer-stiffened shells under axial compression-specimen
supports and end conditions (from [9.222])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
970 Stiffened Shells

The machining process for the steel stringer-stiffened shells was again in stages, employing
similar cooled mandrels. When the blank sat again well on the mandrel, it was set between
the centers of a lathe and the external surface was turned to the designed outside diameter
of the specimen (the difference between the radii of the internal and external surfaces being
equal to the height of stringer, (h + d) ± 0.010 mm of Figure 13.14, measured with reference
to the surface of the mandrel). Then the tube was ready for milling of the stringers.
The mandrel was centered on a milling machine in a manner that assured that the ovality
of the mandrel together with the eccentricity of the centers did not exceed 0.005 mm. Milling
was started only after such precise centering was achieved. Special form cutters with a curved
cutting profile that fit the space between the stringers were ordered for the milling process.
Two types of cutters were used, one of 5 mm width and the other of 10 mm width. These

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
spacings between stringers were one of the manufacturing parameters for obtaining different
values of Koiter's measure of total curvature 8, which determines the initial postbuckling
behavior of the panels between stringers, and where ([13.50])
(13.8)
One of the centers on which the mandrel was mounted fitted into a division head. Using
different division discs, different stringer distributions were obtained with the same cutters
yielding different 8. Variation of stringer distribution and cutters also changed the area ratio
of the shell (A 1 I bh).
During machining it was found that the most precise and even distribution of stiffeners
was obtained if opposite spaces were cUI: one after another. Cutting of adjacent spaces was
also tried and then avoided since it caused uneven stringer distribution as well as variations
in skin thickness of the shell. This was the result of local "relief" of the blank from the
mandrel due to high local stresses, which influenced the fit between the blank and mandrel
and hence caused a deeper cut of the cutter. The best results were obtained when the stiffeners
were cut in as symmetric a manner as possible.
The depth of cutting, or rather the skin thickness, was carefully controlled during the
cutting process by a dial gage with a (I /1000) mm division, which followed the cutter and
measured the thickness relative to the mandrel surface. In spite of this careful control, the
precision of skin thickness was not as good as expected, and thickness variations up to 10
percent of the smaller value were obtained. These variations resulted from accumulating
errors of manufacturing such as local "relief" of blank from mandrel under the cutter during
the cut and deformation of the frame of the milling machine (which was observed to be of
the same magnitude as the allowed tolerances of skin thickness, 3 percent of nominal).
Since the aim of this and other similar Technion test programs was to study the effect of
shell and stiffener geometry on the "linearity" (knock-down factor), the shell parameters
(RI h), (L/ R) and Z, as well as the stiffener parameters (eJ h), (A 1 I bh), (/ 1 J bh 3 ) and Koiter's
8, had to be varied. Many shell configurations were calculated prior to manufacturing a
specimen, checking also the expected stress levels. To ensure elastic buckling of the speci-
mens, care was taken that at buckling, stresses should not exceed half the yield strength of
the material.
In order to study the length effect on imperfection sensitivity predicted in [3.7] and [13.51],
short shells were also manufactured. These shells were machined simultaneously with cor-
responding long shells from one blank. Hence the imperfection sensitivity could be studied
by comparison of the "linearity" (knock-down factor) obtained for the short shell with that
obtained for its twin long shell of practically identical dimensions and very similar manu-
facturing imperfections (see Figure 13.15).
The 34 smaller 10-in. diameter shells of this phase were manufactured in an identical
manner on another similar cooled mandrel of an appropriate smaller diameter.
The second phase of the test program on stringer-stiffened cylindrical shells included 34
integrally stiffened aluminum alloy shells (see [9.228] or [13.52]). The specimens were
machined from 7075-T6 aluminum alloy tubes, 10 in. in diameter and 1/2 in. wall thickness.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Model Fabrication for Stiffened Shells 971

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
(b)

Figure 13.15 Technion tests on integrally stiffened cylindrical shells-typical 7075-T6 suingcr-
Stiffcned shells (from [9.228])

SPECIMEN

CC»·UCA l

BAIH

HCAttNG
--- ELEMENT

Figure 13.16 Technion tests on imegrally stiiTcncd cylindrical shells- the heated internal support man-
drel for aluminum alloy shells and their releasing setup. with the shell half released from
Copyright Wiley the mandrel (from [9.2281)
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
972 Stiffened Shells

The mechanical properties were obta ined from s pecimens c ut from the tubes in the axia l and
circumferential directions, yielding average values E = 7.35 x 10" kN/mm 2 (or 10.6 x 10-'
ksi) and CJ',. = 530 kN/ mm' (or 76.7 ksi).
Fo r calc ul ation of inelastic effects, the s tress-strain curves were approx imated by a
Ramberg-Osgood relat ion [2.77] for both steel and alu minum a lloy s hel ls.
The machining process for the 7075-T6 s pecimens was similar to that descri bed above,
except for the mounting of the b lank on the mandre l and removi ng o f the fi ni shed sti ffened
shell from it. Here, ins tead of using the cooled alum inum mandrel employed for s teel shells.
the b la nk was care fully heated in a temperature -co ntrolled oi l bath (Fig ure 13. 16) to about
IOO"C and then mounted on a steel mandrel very si mi lar in its details to the earl ier cooled
mandrels (see Figure 13.16). After the culling process was completed, the mandrel was
located on a special platform within the oi l bath and the oi l was heated again to IOO"C by
a n e lectrical heating element wi thi n a tube to which the platform was au ached. The heated
s pecime n was then released from the m andrel wi th the a id of a s pecial cylinder tha t pushed
it into the lower part of the oi l contai ner. The bottom of the oi l reservoir was padded with
rubber to avoid damage of the specimen when it was relea~ed from the mandrel. The mandre l
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

was then taken o ut from the bath a nd the s hell was picked up from the bottom of the
container. T hi s process is still used today in the fabrication of si milar s tringer-sti ffened s hells .
The precision of the 7075-T6 specimens did no t differ from that obtained for the earlier
steel o nes, though they were manu factured from a softer material than the s teel alloys used
in (9.206) and (9.222(. The machining procedure of the present s peci mens involved the same
methods o f c ulling a nd measuring and hence s imi lar accumulated errors were also introduced
in the aluminum a lloy s hells. For the 7075-T6 s pecimens the worst deviation in s he ll thick-
ness for a few s he lls was up to 10 percent o f the minimum ski n thickness. The usual devi-
ation, however, was wit hin 5 percent of the minimum th ickness.
In orde r to obtain ditlerent values o f s tringer dis tribution (b//r). stringer area ratio (A, I
bit) and Koiter's c urvature parameter IJ of Eq. ( 13.8). special form cullers were again o rdered.
Fo ur types of cullers were used. with a c utting width of 5 nun, 8 nun, I0 111111 and 12 tnm.
By use o f these dill'e rent c ulle rs together with various division discs, the above mentioned
parameters were separately or s imu lta neously a ltered. Postbuckli ng pallems of typical 7075
s tringer-stiffened s hells are shown in Figure 13.17.

Figure 13.17 Technion tests <m integrally Stiffened cylindrical shciiS- t)•pical postbuckling panems of
Copyright Wiley axially compressed 7075-T6 shells (from [ 13.951. 113.96])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Model Fabrication for Stiffened Shells 973

As in the case of the steel shells, many shell configurations were calculated prior to the
manufacturing of a specimen to predict the stress levels. To ensure elastic buckling the
specimens were again designed to fail at stresses less than half of the yield strength of the
material. As before, twin and triplet shells were manufactured in order to assess the predicted
length effect on imperfection sensitivity [3.7] and [13.51], as can be seen for example in
Figure 13.15.
As discussed in [13.52], [9.222] and [9.228], the tests on both steel and aluminum alloy
stringer-stiffened shells did not verify the strong imperfection sensitivity length effect pre-
dicted in [3.7].
As pointed out in Chapter 9, Subsection 9.3.5, the investigators at the U.S. Navy David
Taylor Model Basin excelled in the quality of their steel and aluminum alloy, unstiffened
and stiiiened machined shells. Their stiffened shells were all ring-stiffened and usually rel-
atively thick (R I h ~ 40-180, mostly below I 00), which resulted in their collapse being
elasto-plastic or entirely plastic (see for example [ 13.53]-[13.57]), and in many cases per-
mitted accurate machining without internal supporting mandrels. To minimize initial imper-
fections and residual stresses, most shells were machined from thick seamless, stress-relieved
tubes, and high-yield steels or aluminum alloys were employed to prevent premature yield-
ing. Many of the models were fairly small (8 in. diameter and less) and in the thinner ones
mandrels were employed for the final accurate machining stages.
The DTMB tests included not only many ring-stiffened circular cylindrical shells but also
stiffened oval cylindrical shells (see for example [ 13.58]), stiffened conical shells (see for
example [9.325]) and stiffened spherical shells (see for example [9.312]). The measured
initial out-of-roundness (magnitude of initial imperfection) and thickness variations of these
models were usually not more than a few percent of the shell thickness. For example, 2-3
percent of h in the stiffened spherical caps of [9.312], 6-9 percent in the stiffened cylindrical
shells of [13.56], but approaching 25 percent in the welded cone-cylinder junctures of
[9.325]. Besides machining, also rolling and welding, being more representative of full-scale
fabrication methods, were used for the manufacture of stiffened shell models (see for example
[9.325]).
Machined models of stitiened shells were also tested by many other experimental groups
in the sixties, seventies and eighties-for example, aluminum alloy internal integral ring-
stiffened cylinders tested under axial compression and external pressure at the Avco Cor-
poration, Wilmington, Massachusetts [13.59]; steel ring-reinforced cylinders under external
pressure at the Portsmouth Polytechnic, England [13.60] and [13.61]; steel ring-stiffened
cylindrical shells tested under axial compression at the Norwegian Institute of Technology,
Trondheim [13.62], where a comparison with similar larger cold rolled and welded shells
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

was also made; steel ring-stiffened cylindrical shells used at Tokyo Denki University [13.63],
where again a comparison was made with similar welded shells; and other groups mentioned
in Babcock's survey [9.58].
Some comments are warranted with respect to these examples of machined shells: Midgley
and Johnson at Avco [13.59] undertook the difficult task of machining 16-in. diameter models
with internal reinforcing rings, and indeed their shells had a maximum out-of-roundness
varying between 10-55 percent of the wall thickness. Ross et al. at Portsmouth [13.60]
machined their shells from solid 6-in. diameter steel bars, using an additional mandrel for
machining of the external surface, and obtained accurate models with a maximum out-of-
roundness of 1.6-3.3 percent of the shell thickness, similar to the precision achieved in the
larger Technion shells discussed in the beginning of this section. Odland at Trondheim
[13.62] machined his 400-mm diameter models from 12-mm thick seamless tubes. For the
final external machining, the tubes were mounted on a loosely fitting mandrel (or plug) and
the space between plug and tube was filled with a low melting point alloy Cerrobend, which
could then be easily remelted for removal of the finished specimen. This method of support
resulted, however, in some eccentric machining, which produced thickness variations of up
to 15 percent of the shell thickness. Yamamoto and his associates at Tokyo University [13.63]
Copyright Wiley
employed
Provided by modern
IHS Markit under numerically controlled machine
license with WILEY tools to Inc
Licensee=McDermott obtain accurate
- India/8215328006, specimens
User=G, Boopathi and
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
974 Stiffened Shells

machined them slowly to reduce residual stresses. They thus obtained better specimens than
in their earlier experiments [13.64]. Their more recent 300-mm diameter shells had maximum
thickness variations of about 4 percent of the wall thickness, similar to those of the Technion
shells. However, they also succeeded in keeping the height and width of the ring-stiffeners
within nearly the same tolerances (4-6 percent of h), whereas the width of the rings of the
Technion shells exhibited larger deviations. Indeed the buckling pressures of the Tokyo ma-
chined shells were within 0-5 percent of the theoretical values, supporting the assertion of
[13.63] that "buckling of machined models is governed by the theory for perfect cylinders."
It can therefore be concluded from this section that with sufficient care, highly accurate
specimens even with R I h = 600, nearly perfect ring- or stringer-stiffened shells, can be
readily machined today. However, though machining includes stress relieving at intermediate
stages, some residual circumferential stresses may still remain, as was found at Technion
when some shells were cut up for use in stiffened panel tests.
Furthermore, the comparisons between cold-rolled and welded shells and their equivalent
smaller machined models carried out in [13.62] and [13.63] indicated a relative reduction in
load-carrying capacity of the welded shells of 20-35 percent.

13.2.2 Small-Scale Welded Specimens


The frequent use of stiffened cylindrical shells as components of offshore structures and
other marine structures has motivated considerable experimental work. The claim that be-
cause of the difference in the fabrication of offshore shells, which are welded from high-
strength steel plates, the results obtained in tests of high-precision integrally machined spec-
imens are not reliable has led to many tests on large shells, to be discussed separately, but
also to the development of special model fabrication techniques.
Walker and his associates at University College, London, developed a manufacturing and
welding technique of small-scale specimens with Rlt = 200-360 (see Figure 13.18) that
represented typical offshore elements on a 1120 scale [13.8] and [13.65]. The cylindrical
shells were built up segmentally by rolling and machining thin steel sheet, closely similar
in material characteristics to that employed in offshore structures, into curved panels and
then welding these together with the stiffeners.
The welding technique consisted essentially of clamping two curved panels and a stiffener
plate between heavy, copper-lined heat sinks that had been accurately machined to the ap-
propriate inside and outside radii and transversing the junction under a tungsten inert gas
(TIG) welder electrode to form a fusion weld (see Figure 13.19 and [13.8] and [13.65]).
Both the quality and heat input of the weld could be controlled by altering the speed of
trolley that carried the welding torch and the amperage of the arc. The high quality of the
weld achieved can be seen in Figure 13.18. Measurement of imperfections and residual
stresses showed that their relative level on the small models was similar to that obtained on
large shells.
The work was continued by Walker and his co-workers at the University of Suney [13.66],
and small-scale ring-stiffened shells with Rlt = 150 were made by a similar process (see
Figure 13.20). For the ring-stiffened models, the rings and shell bays were machined to
precise dimensions and mounted in jigs set on a circular table, as shown schematically in
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 13.20. "The bays were held at the top and bottom between copper lined circular
plates at the inside and rings on the outside. The welding was carried out by rotating the
table at a controlled speed relative to the electrode. The vertical joint on each bay was butt-
welded using an arrangement similar to that used in the stringer-stiffened specimens."
As part of the same U.K. offshore element experimental program, small-scale models
similar to those of University College, London, were developed at Imperial College by
Dowling and his associates, using a different fabrication technique. Their models were fab-
ricated by first TIG welding the longitudinal stiffeners to a fiat plate, the length of which
was equal to the circumference of the cylinder plus a small allowance for weld shrinkage.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Model Fabrication for Stiffened Shells 975

(a)

(b)

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Figure 13.18 University College. London. small-scak wcklcd ;tccl stringer-stiffened shell~ that failed
by overall buckling >hell UC 9. with 40 winger. (courtesy of Professor A.C. Walker):
(a) postbuck.lcd ; hapc under axial compres~ion, (h) '!ringer tripping failures that arc pan
of the ovemll huck ling collapse

After welding. this stiffener sheet was wrapped around a copper forming mandrel and the
closing T-butt weld was completed. The cylinder was then enclosed in an outer copper
mandrel. and in this tightly re\trained condition it was ~tress-relieved in an inen atmosphere
heat treatment oven (sec Figure 13.21 and [ 13.9). ( 13.671 and 113.68)). Then the ends were
turned square and parallel and bound in Araldite and sand resin between hea\'Y steel collars
to maintain rigid end conditions (see Figure 13.22). The stiffened shell protruded a shon
distance fro m the end clamps, so that the load was applied to the cylinder itself.
Imperfection and residua l stress measurements o n the 30 University College London and
6 Imperia l College models yielded simi lar relati ve levels to those obtained o n large shells.

Fiuurc 13. 19 University College London ;mall -sca le welded steel stringer-stifTened shells welding
of shell panel and <tiffcncr\ u~ing copper-lined heat sinks (from [ l3.8j. Crown copyright
is reproduced with the pcmli<~ion of the Controller of Her Majesty's Stationery Office)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
976 Stiffened Shells

Figure 13.20 University of Surrey s mall-scale welded steel ring-stiffened shells-welding of ring stiff-
eners with cylindrical shell bays using copper li ned heat sinks (from [ I3.8). Crown
copyright is reproduced with the permission of the Controller of Her Majesty's Stationery
Office)

0 5 10 15 20 25.
(a)

(b)

Figure 13.21 Imperial College small-scale welded steel stringer-stiffened shells- a typical 40-stringer
model (courtCS)' of Prok~sor P.J. Dowling): (a) the shell after welding and heat treatment,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(b) detail of fusion fi llet welds of the 0.84-mm thick plates


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Model Fabrication for Stiffened Shells 977

Fi~:ure 13.22 Imperial College ;rnall·scale welded steel stringer·\Uffcned shells-completed 40->tringer
model (courtc'y or Prorc>sor PJ. Dowling): (a) completed model. (b) end nng detail. (c)
rings used in end ca\ling process. Nme thm in (b) the shell protrudes s lightly rrom the
resin to take the load direct ly

The results of the small -scale 1es1s could therefore be taken as a reliable complemenl to the
largc-~cale tests performed in the same program at Imperial College and the Universily of
Glasgow (see I13.69) and [1 3.7)). An additional test program of 12 similar single-bay
slri nger-stiffened shells wilh Rlt ;, 190 was al~o carried out a1 Imperial College for Del
norske Veri las (DnV). the shell~ being tes1ed under combined axial compression and external
pressure (see [ 13.68)). The wa1erproofing requiremenh of the pressure loading in lhe com-
bined loading 1ests required different end clamping. leadi ng to the end ring detail discussed
taler in Subsection 13.4. 1.
T he fabrication of the Imperial College small-scale s1iffcned cyli ndrical shells was sum-
marized and discussed in dewil in a 1987 paper [ 13.70]. well worth reading, parts of which
arc now briefly recapitulaled.
The model shells were joimly developed for some years by a small firm of model makers,
RME Ltd .. Reading. U. K.. and the Civil Engineering Department of Imperial College. AI
the oubet the aim was defined as "to provide accurate modelling at a small. and more
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

economic scale, of models which ~atisfied the tolerance requirements of the pro1o1ype struc-
ture. and which could be used for both elastic and uhimate load testing to provide behaviour
wh ich would resemble in nearly all respects the behaviour of the prototype" 113.701.
" h was accepted thai lhe welding process used could not precisely model lhe process
used in the pr0101ype, thai lhc heat input would he proportionally much greater, and 1hat
ex tensive jiggi ng and heat conduction capability would be required to mainwin model dis-
ton ions within acceptable bounds. h was also recogni'"ed at an early stage that 1hc model s
would need to be stress relieved after fabrication. while maintaining strict control of toler-
ances. to remove the unreali~tically high levels of weld induced residual strc'>se~ which
resuhed from the very small ~calc fabrication 1echniquc."
Slringer-stiffened shells were the first type of models 10 be developed. partly because of
1he research requirements, and partly because all stiffener welding was on nat plate prior to
rolling and therefore the ac1ual fabrical ion process was ralher ~irnpler.
"AI a taler stage ring sliffcned models were also fabricaled and then OJthogonally stiffened
models. incorporating bolh rings and stringers were made. using extremely soph isticated
jigging:·
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
978 Stiffened Shells

As outlined earlier, the first main operation in the fabrication of stringer-stiffened shells
was the fillet welding of the stringers, while the second one, described now in more detail,
was the wrapping of the shell around a segmental mandrel (see Figure 13.23).
[The inner part of this] mandrel is made from 40 mm thick steel tube. The thickness eliminates
any distortional problems during heat treatment. The segments are made from a similar tube. Both
tubes are stress relieved prior to finishing to ensure correct shape. The mandrel has recessed ends
to accept a tooling disc so that its outer face can be accurately machined to the required diameter
for the segments. Three lines of tapped holes, one at each end and one at the centre of the mandrel,
allow the segments to be held in position whilst their outer diameter is turned to size. The segments
are then held in position while their sides are gang milled to the theoretical stiffener separation
with a 0.05 mm clearance allowed each side of each stiffener. The depth of the segment caters for
the depth of the largest stiffener to be used.
To assemble, about six segments are slid between adjacent stiffeners and one segment is then
attached at each end using cap screws so that the segments are aligned with the mandrel axis. The
remainder of the segments are then progressively slid into place, with cap screws where possible,
and the wrap around completed using two segments with copper inserts on the outer corners where
the butt welding of the can is to take place.

When the shell had been closed with a T-butt weld, the weld was "dressed down to the
shell surface and a 0.25 mm shim is wrapped around the model so that the outer split clamp
can be bolted into position without damage to the model."
After stress relieving, the outer clamps were removed, then the segments were slid out
and the inner part of the mandrel was withdrawn, and finally the model was mounted in the
end rings (see Figure 13.22).
The ring-stiffened models were also made in two stages, but in reverse order. Here the
first stage was making complete unstitiened cylindrical shells of length equal to the spacing
between rings, while the second stage was T-butt welding of these shells with their internal
rings from the outside. The ring stiffeners were then made as complete annuli, either by
trepanning out of plate in the case of flar stiffeners or by turning from the tube in the case
of T-stiffeners. Only for ring stiffeners have other than flat bar stiffeners been used in the
various series of models.

outer clamp
/
str!nger----
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

stiftened
plate

single
segme-nt
ottached
I to mandrel
segments

Figure 13.23 Imperial College small-scale welded steel stringer-stiffened shells-exploded view of
complete fabrication assembly with segmental mandrels (from [13.70])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Model Fabrication for Stiffened Shells 979

The welding and assembly jig used, shown in Figure 13.24, consisted of a central solid
core (solid apart from lightening holes), stiffener retaining and welding jigs with copper
inserts (acting as heat sinks) and locking screws. The configuration shown in the figure is
for T-rings, whereas for fiat ring stiffeners copper inserts with a plain rectangular cross-
section replaced the shaped ones.
An interesting point relating to the heat treatment is that for the ring-stiffened cylinders
this
is accomplished with the model shape retained using sand rather than a clamp. The models are
appreciably stiffer than the equivalent stringer stiffened cylinders because the rings maintain cir-
cularity. A tooling ring is positioned in both ends of the cylinder and a fine foundry sand packed
inside the model and around the outside to a thickness of about 75 mm, in a drum container.
After the treatment, end flanges are attached in a similar manner as for stringer stiffened cylin-
ders.
Orthogonally stiffened shells, with internal stringers and rings, were also fabricated by
similar techniques, but with a more complex stiffener cage welding jig, described in detail
in [13.70].
The cost of the models was also discussed in [13.70], having been the primary motivation
for testing at such a small scale. It was quoted for 1986 as on the order of $3,500 for a 320-
mm diameter steel shell stiffened in one direction and about $15,000 for a similar orthog-
onally stiffened model. Though these small shells were not inexpensive, their cost was still
only on the order of one-third to one-fifth of a larger-scale (say 4-5 times larger) model
fabricated by the usual rolling and welding process. However, the cost savings of actual
small-scale testing are probably considerably higher because of the significantly smaller test
equipment required.
By comparison, the 1986 cost of one of the 250-360-mm diameter accurately machined
Technion ring- or stringer-stiffened cylindrical shells (like those shown in Figures 13.12 and
13.15) was about $1,000. Hence for extensive behavior tests or parametric studies small-
scale tests appear essential, even if they do not entirely replace large-scale tests in the final
stages of a project.

13.2.3 Realistically Fabricated Small-Scale Stiffened Shells


The "realistic specimens approach," discussed in Chapter 9, Subsection 9.3.7, has also been
employed for stiffened shells. Miller [9.169] fabricated his small ring-stiffened shells (with
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

./"
Shell plate

Solid cenlral core

Figure 13.24 Imperial College small-scale welded steel ring-stiffened shells-welding and assembly
jig for ring-stitTened cylinders (from [13.70])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
980 Stiffened Shells

Rl t = 250-500) by rolling and welding, similarly to his large-scale shells of Rl t = 500,


and showed that the rings have a beneficial effect on increasing the elastic buckling strength
even when their spacing is considerably larger than the theoretical "safe" spacing of
L72vRt, and therefore the collapse is by local buckling of the unstiffened subshells between
the rings. The reason for this increase is the reduction of imperfection sensitivity due to the
ring stiffeners. A similar conclusion was reached in a series of tests carried out at Det norske
Veritas [3.22] and [13.71] on larger ring·stiffened shells (with Rlt = 240) fabricated from
aluminum alloy by rolling and welding.
Another example is the 20 hydrostatic external tests by Miller and Kinra [13.72] on 16-
48-in. outside diameter ring-stiffened shells made from commonly used platform steels by
routine platform fabrication procedures. The shells, with (R It) ~ 16-63, buckled partly in
the elastic and partly in the inelastic range. The geometric initial imperfections for most
specimens were found to be larger than the allowable out-of-roundness tolerances. The spec-
imens of many of Miller's tests then became even more "realistic" by being larger, and are
therefore discussed in a following subsection.
The cold-rolled and welded ring-stiffened shells used by Odland [13.62] and Yamamoto
[13.63] for comparison with their integrally stiffened shells, are also examples ofrealistically
fabricated ones. In both cases the welded shells were three times as large as the machined
ones (and with R = 600 and 490 mm, respectively, could already be considered large shells).
Of the 12 Trondheim shells, with R It = 200 (which were fabricated in an industrial
workshop, experienced in the manufacture of cylindrical tanks-[13.62]), six were formed
to cylindrical shape by rolling of one single plate, and then closed by a vertical butt weld
welded from both sides. For the remaining six shells, two plates were first joined by a butt
weld, and only then were they formed to cylindrical shape by rolling and finally closed by
a vertical butt weld as in the first batch. The additional welded joint in the second batch
was introduced to create intentionally a zone of moderate distortions for comparison with
the first batch of shells without these distortions. The internal end rings as well as the external
reinforcing rings were also formed to circular shape by rolling and then joined to the shell
by intermittent staggered fillet welds. The edges of the stiffened shell were finally machined
plane and parallel to high accuracy in a vertical mill. However, though the load-carrying
capacity of the rolled and welded shells was considerably below that of the machined ones,
as pointed out earlier, for closely spaced rings the buckling loads were found to be practically
unaffected by both the additional circular butt welds and the vertical ones.
In the case of the more recent Yamamoto eta!. tests [13.63], the two welded ring-stiffened
shells were fabricated by conventional techniques. Two 7.5-mm thick plates were bent to a
semicircular shape and then joined by welding. The rings were machined from 15-mm thick
ring-shaped plates (which had been fabricated by welding) and were joined to the shell by
full-penetration welding. The postbuckling collapse patterns of the welded model W4 and
its corresponding 1/3 scale machined model M4 are shown in Figure 13.25. Both shells
buckled in similar patterns with the same number of circumferential waves n = 3, but
whereas the machined model carried 99 percent of the predicted load, the welded one carried
only 82 percent of the predicted load.
Further examples of "realistically" fabricated stiffened shells are the welded aluminum
alloy cylinders (R/t = 400) with large rectangular cutouts tested in compression by Pal-
chevskii and Polyakov [13.73] or the welded steel shells (Rit = 260, 320, 400), reinforced
by internal pressure, an elastic core or externally wound glass filaments, tested in bending
by Grinenko et a!. [13.74].
Still other examples of "realistically" fabricated stiffened cylindrical shells are the ring-
and stringer-stiffened model shells of the joint CBI Industries, Plainfield, Illinois, and Glas-
gow University test program [13.75] and [13.76]. This extensive test program, jointly spon-
sored by Conoco Inc., the American Bureau of Shipping, Standard Oil Co. of California, the
U.K. Department of Energy and Mitsubishi Heavy Industries Inc., included 66 stiffened steel
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

shells of (Rit) = 170-530, 22 ring-stiffened ones (Series 1) and 44 ring- and stringer-
Copyright Wiley
stiffened ones (Series 2). All the rings and stringers
Provided by IHS Markit under license with WILEY
in these shells were internal ones (see
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Model Fabrication for Stiffened Shells 981

(a) (b)

Figure 13.25 External pressure tests by Yamamoto et al. on ring-st iiTened welded and corresponding
one-third scale machined steel cylindrical shells-<ollapse patterns for shells with
(R it) = 65 and (LI R ) = 2. having 11 = 3 (from [13.63), courtesy of Professor Y. Yama-
moto): (a) welded shell W4. (b) machined shell M4

for example Figure 13.26, showi ng one of the models tested by CBI). Most of the models
except the smallest ones, though scale models of much larger prOtOtypes, cou ld probably be
classified as large shells, having radii of 360- 957 mm (1.42-37.7 in.) and will therefore
appear again in Subsection 13.2.5.
The fabrication processes used for all the models were representative of those employed
in offshore applications, such as shearing, name cutti ng, rolling and welding (though in the
models this was limited to metal inert gas, [MIG) and tungsten inert gas. (TIG]. welding).
The required tolerances were also based on the requirements for the corresponding full-size
struct ures. Though the fabrication techniques were essentially similar (for example, shell
plates were welded together prior to rolling) for the 18 ring-stiffened GU shells and the 48
CBI models. there were some differences, as set out below.
For the Glasgow University ring-stiffened models,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 13.26 CB I, Plainfield Ill .. collapse tests of fabricated sti iTencd steel sheiiS- t)•pical stringer-
and ri ng-stiffened model, 2-7F. buckled under hydrostatic extemal pressure (from [13.75),
Copyright Wiley courtesy of Dr. C. D. Miller!)
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
982 Stiffened Shells

the shell plate was rolled into a cylinder and the longitudinal joint was welded. The completed
cylinder was cut into the three bay lengths and the center bay was rotated about 15° to stagger the
seam welds. The shell bays were separated to allow the rings to be inserted; the outer edge being
positioned just beyond the inside surface of the shell. Custom built collapsible clamps were placed
internally and left in place during heat treatment. Copper strips were placed in the clamps near the
weld joints to act as heat sinks. Heavy rings were also fitted to the ends of the models during
fabrication to help maintain shape.

These models were actually three bay unstiffened shells, with the rings serving only as
edge rings for the bays, the central one being the test section proper.
For the CBI ring-stiffened and ring- and stringer-stiffened shells,

the shell plate was rolled into a cylinder and the longitudinal joint was welded. The rings were
burned with a plasma arc into four 90o segments and welded together. The radial slots for the
stringers were made with a punch press. The ring stiffeners were fitted inside the cylinder and tack
welded. The stringers were inserted through the slots in the rings and tack welded. The end rings
were fitted and tack welded. All fillet welds were then completed. All welds were made with the
TIG process except the butt welds in the shell that were made with an automatic seam welder
using the MIG process. A water bath was used as a heat sink for controlling distortion.

Before leaving this section, it may be of interest to briefly mention two examples of
realistically fabricated small-scale sparsely stiffened aerospace shells. The first is the stringer-
and ring-stiffened shells tested by Harris et al. at North American Aviation, Downey, Cali-
fornia, in the late fifties as part of their well-known series of experiments discussed in
Chapter 9 [9.187]. Their exploratory tests were initiated to investigate the feasibility of
fabricating small-scale stiffened shells economically by adhesive-bonding techniques.
The 8.75-in. (222-mm) radius cylindrical shells were made of very thin, 0.0032 and 0.0087
in. (0.08 and 0.22 mm), half-hard 18-8 stainless-steel foil, typical of the Rl t = 1000-2700
shells considered then for pressurized launchers. The external hat-section stringers were also
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

formed from the same foil, whereas the internal channel-section rings were made from 2024-
T3 aluminum alloy. "The fabrication sequence of the shells was as follows: (l) the longi-
tudinal seam of the shell was bonded, (2) the stringers were bonded to the outside surface
of the shell and (3) the aluminum alloy rings, which were split to allow a reduction in
diameter for insertion inside the cylinder, were then bonded in place." All bonding was with
Epon VI adhesive, under temperature and pressure.
In the final collapse tests, however, most of the specimens failed prematurely by stringer
debonding, indicating that at the time metal-to-metal bonding still needed further improve-
ment. Today the process would prevent premature debonding failures.
The second example are the stringer-stiffened cylinders tested under axial compression at
the Aeronautical Research Institute of Sweden (FFA) in the late seventies [13.77]. The 250-
275-mm radius shells were made of 0.7-mm aluminum sheets and the stringer from extruded
aluminum L-profiles. Their Rl t = 360-390 was therefore in the range of aircraft construction
practice. The stringer spacing varied from 4SO for the heavier stringers to 30° and 18° for
the lighter ones. In 7 of the 13 specimens the stringers were located on the outside, and in
the remainder on the inside.
The shells were fabricated by rolling the sheets, after the stringer locations were marked
with a soft pen, and joining them with a bonded lap joint. The stringers were cut shorter
than the shell length to prevent local point loading on one or more of the stringers. The
stringers were bonded and widely spaced riveted to the shell. Araldite AW-106 adhesive was
used and the bonded surfaces were roughened up with emery cloth prior to bonding. The
rivets were used to provide pressure on the bonds during curing.
The shells were designed for local panel buckling. For general instability to predominate,
many more stringers would have been required, about 55 compared to the maximum of 20
used. Beyond panel buckling, however, the deformation and final collapse also involved the
stringers, with no premature debonding (see for example Figure 13.27, showing the collapse
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Model Fabrication for Stiffened Shells 983
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 13.27 FFA (Sweden) tests on ax1all) compressed aluminum alloy stringer-Miffcned shells-
collapse pattern of a typical ;pecimen with 20 >!ringers. no. 20-2 (from 113.77])

of a typical specimen with 20 stringers). Though loc~ J buckl ing dominated in these tests,
the stringers were found to influence the buckling loads, in particular for the heavier stringers.

13.2.4 Plastic Models


The fabrication techniques for plastic models developed for unst iffened shells and discussed
in Subsections 9.3.2- 9.3.4 of Chapter 9 were also employed for fabricat ion of accurate
stiffened shells, and new techniques have been developed for them.
The advantage~ of Mylar and similar polyester films. which due to their high (uJ £)values
permit repeated deep postbuckling deformations. motivated their use as model~ for stiffened
shells (see for example 113.78) and ( 13.791). Figure 13.28 <,hows typical postbuckling pat-
tern> of one of the DFVLR (now DLR) Bn1unschweig ~tnnger- and ring-~tiffened Mylar
shells [13.78) under axial compression. The stiffeners were Mylar angle sections. or double

Fi~u rc 13.28 DFVLR (now DLR) w inger- and ring-stiffened Mylar cylindrical <hell' subjected to
axial compression-postbuckling pauems (from 113.781): (a) under ~mall axial shon-
cning and (b) under large ax1al >hol1ening
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
984 Stiffened Shells

angle sections, of the same thickness as the skin of the shell and bonded to it (see Figure
13.29). There was no stiffener debonding, even at very large postbuckling deformations.
These DFVLR shells, of R = 100 mm, R It = 518, Ll R = 3.3, were not closely stiffened,
but the rings and stringers were rather \\<idely spaced, i.e. notably discretely stiffened. The
experimental load-shortening curves under axial compression presented in Figure 13.29 show
their buckling and postbuckling behavior distinctly. The panel buckling is practically unaf-
fected by the stiffeners, whereas the postbuckling behavior is. However, only stiffening by
both ring and stringers raised the collapse load beyond the buckling load of the corresponding
unstiffened shell. Though the stiffener weight was only about 20 percent of that of the shell
and their spacing was rather wide, the final collapse load was about 62 percent above that
of the unstiffened shell, or would still have been about 13 percent higher had the corre-
sponding unstitiened shell been thickened to account for the additional material of the stiff-
eners.
Another example of stitlened Mylar shells is the 15-in. (381-mm) diameter ring-stiffened
cylinder with reinforced openings tested at CBI as part of a test program on shells with
openings ([13.79], or see Figure 3 of [1.17]).
Among the techniques for manufacture of accurate plastic models employed for unstif-
fened shells that have been developed for stiffened shells, one may mention the spin casting
of liquid epoxy plastic applied also to fabrication of stringer-stiffened cylindrical shells of
Rlt = 100-150 at UTIAS, Toronto [1.15] and [13.80].
Or one may note the casting technique for epoxy ring-stiffened 60° truncated conical shells
developed by Williams and Davies at NASA Langley (see Figure 13.30 and [13.81]). This
yielded models of maximum and minimum radii of 223 and 142 mm, respectively, with
mean thickness of 1.02-1.04 mm and a maximum thickness variation of ± 2.5-4.3 percent.
The specimens had very uniform material properties and very small geometrical imperfec-
tions, except near the end rings, where meridional imperfections of up to 0.8 t were measured,
which, however, are still fairly small.
The casting process, using a precisely machined two-part aluminum mold, involved plac-
ing the mixture of the component materials in a vacuum chamber for 10 minutes (which
drew off volatile gases that could cause voids), then filling the ring cavities of the male mold
with resin and finally lowering the female mold component. The epoxy resin was then fed
into the mold cavity from two reservoirs above the mold, with a vacuum pump simultane-
ously developing reduced pressure inside the cavity (which again ensured the absence of
voids). Then the vacuum line was disconnected and the mold transferred to an autoclave for
curing under pressure at room temperature.
The casting technique developed produced repeatable high-quality epoxy specimens of
relatively complex shells, which could be repeatably tested without failure and could provide
useful strain and displacement data.

8 stringers ouls;de
2 rings ;ns;de

.., 700 ------TSiringii;:~a;;d.


ring stiffened
~
1

I
kg
·~
"" 50

0 o~--------------~os----------------~~o·mm
--`,`,`````,`,````,,``,``,,`-`-`,,

Axial shortenmg

Figure 13.29 Experimental load-shortening curves of DFVLR (now DLR) stiffened and unstiffcned
Copyright Wiley Mylar cylindrical shells under axial compression (from [ 13.78])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Model Fabrication for Stiffened Shells 985

Figure 13.30 NASA Langley ring·, Liffcned cast-epoxy conical ' hclls-buclded shell cxhihiting three
full circumferential wavc5 (from [13.81]1

For MiJTened hyperboloidal >hells. Mungan at Ruhr Univeroity Bochum. Gennany [ 13.82[.
extended the essentially similar technique of casting hi~ models from epoxy resin between
two molds that he had used for unstiffened she lls (sec [9.135J). The rectangulareross-scction
ring!. o r stringers (meridional ribs) were glued onto the shell o nly after the ba~ic unstiffened
she ll was first tested for buck ling behavior. Furthermore. the glui ng technique allowed re-
peated tests o n the same shel l with different stiffenct· arrangements.
Gluing the rib reinforcement to the shell was also employed hy Tillman at the University
of Manchester [ 13.83 ] for hi s '> pherical dome buckling tests. The specimens there were
fabricated from hot PVC sheets. being pressed by air agai n'> t a spherical metal mold. For
accuracy. the ribs were made from a second dome. fonned oimultaneou.sly. which after suit-
able sectioning was glued onto the well-fining mated shell. The different rib arrangements
were therefore each mated to its shell and no repeated buckling of the same shell was
attempted.
As the last examples of plastic shells. o ne may mention the development-type experi ments
for improved methods of stiffening, catTiecl o ut in nerospncc industries in differenl countries
in the six ties and seventies. 'ly pical of these are the Lcxnn (a type of perspex) isogrid panels
and cylindrical shells employed by McDonneii-Dougl a~ Astronautics 10 model Delta inter-
~tagc sections and Space Shuttle tank cylinders (sec for example [ 13.84] and [ 13.85 ]). lsogrid
~tructures are efficient grid-member and skin structure' in which the grid members are ar-
ranged in an isosceles triangular pauem. Aluminum alloy flanged isogrid stiffened cylindrical
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

and conical shells were u\ed in the Thor/Delta vehicle. :md Lexan models were extensively
employed in their preliminary buckl ing tests.

13.2.5 Large Stiffened Shells


It wa~ pointed o ut in Sect ion 13.2.3 that many of the "reali stically" fabricated models could
be cla~sified as large shell s. Indeed. many of the models were made fairly large to incorporate
the detailed construction of the full-scale prototype!. and therefore were practically full scale.
The :.late of the an of fuii ·'>Cale and large model con.,truction methods in the aerospace
indu~try m the earl) seventie'> wa.~ reviewed b) Babel et al. [ 13.84]. Earlier large model
buciJing tests of cylindrical shells at NASA. a' well a., those carried o ut in the seventies,
were di:.cussed by Peterson and other researchers at NASA Langley [9.249] -[9.252[ and
[3.23 [.
The NASA models also included two large machi ned alum inum alloy (22 19-T8 1) cylin-
ders of 1.96 m (77 in.) diameter and 1.22 rn (48 in.) leng th. integrally stiffened by rings and
stringer<; of rectangu lar cross-secti on (see Fig ure 13.31 and [9.251 ]). The cy linders were
Copyright closely
Wiley :.tiffened to prevent local buckling and en~urc dominance of general instability. Each
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
986 Stiffened Shells

(a) (b)

Fi~:u re 13.31 Large 1.97-m diameter machined aluminum alloy C) lindrical shells. integral!) stiffened
by rings and stringers of rectangular cros>-'«IIOn tested at NASA Langle) Re~arch
Cemer ([rom [9.251J): (a) \hell I with closet) 'paced rings. (b) shell 2 "llh wider ring
spacing

cylinder was fabricated from 10 equal-width panels of the same length as the cyl inder. The
panels were miUed from llm 12.7-mm (0.5-in.) thick plate. "The first operation consisted of
removing approximately 1.5 mm (0.06 in.) of material from each surface of the plate. Then
the pockets between stringers and rings were formed by mechanical mi lli ng to a depth. width
and length which left approximately 0.38 mm (0.015 in.) of material to be removed later by
chemical milling. The panels were then rolled into cyl indrical panels." For the rolling op-
eration. the pockets fonned by the milling were filled with a low-melting-temperature alloy
in order to prevent ring buckling and llat areas between stringers. After rolling. the filler
alloy was removed by melting and the excess aluminum alloy was removed by chemical
milling. The panels were fabricated initially from 22 19·T3 1 alloy and were then aged to the
T81 cond ition in the time interval between the rolling and chemical milling.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
The panels were welded together along generators of the test cylinder with tungsten inet1
gas fusion welding (TIG). Thoug h steps were taken to mi nimize the geometric imperfections
d ue to the welds, signi ficant general bowing wa~ found (in the measurements carried out
with a special scanner) at most of the weld lines. Variations of ±I 0-20 percent in skin
thickness were also measured. with similar variation~ in stringer and ring wall thicknesses.
The cylinders were subjected to bending in the NASA Langley test rig (see Figure 9.75b).
In ~pite of the significant thickness variations and bowing imperfections along the weld lines,
the di~crepancies between experimental and predicted perfect cylinder buckling moments
were smal l. experimental va lues being 2-8 percent below the predicted ones. Apparent ly the
types of geometric imperfect ion present were not simi lar to the shapes of the buckling modes
(which are definite in the case o r closely stiffened she lls) and there fore no t very detrimental.
Other typical aerospace examples are the 3.05-m (10- ft) diameter 2024-T851 ulum inum
alloy isogrid-cylinder test speci mens. mentioned earlier. which represented Skylab wall pan-
el; or propellant tanks [1 3.85] and [13.86]. These isogrid cylinders (wi th integral stiffening
member~ that formed a network of equilateral triangles) were assembled from three curved
panels. which were also initially machined from Hat plates and then roll·fonned to the proper
curvature. This fabrication method was essentially the same as that used on ,; milar-sized
integ rally stringer-stiffened shells discussed above. Funhcr aerospace examples arc the riv·
ctccl 1.8-m (6-ft) diameter stiffened cyli ndrical shells. which were closer to aircraft fabrica-
tion practice, tested by Horton f 13.87] at Georgia Tech in the seventies.
T he ocean engineering industry in the last two dec;~ucs has motivated extensive buckling
tes t programs on large models of unstiffened and still'ened shell structures. ,;nee in order to
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Model Fabrication for Stiffened Shells 987

pro' ide reliable data for welded marine and offshore structures. it was thought advisable to
fabricate rather large specimens that would repre~ent the practical geometric imperfections
and residual stresses accurately.
These elTorts yielded many hi gh·quality stiffened shell speci mens in the seventies and
eight ies. such as for example the Det norske Yeritas 2.5-m diameter aluminum alloy welded
closely spaced stringer-stiiTcncd shells (Figure 13.32) or simi lar ring-stiffened shells [3.22]
and [ 13.711. The DnV shell~. like that shown in the figure. had (Rit) = 360-420, stringer
area mtios = 0. 17-0.20. and the number of stringers was large enough to ensure failure by
general instability (N = 60-85. \tringer spacing. bit - 31-44). The stilTeners were welded
to the shell with staggered fillet "elds of 100 mm length. In order to limit the radial welding
distortion~. the shells were Mtpported during welding by two to three heavy rings. After
weldi ng. the models were kept eight hours at I50"C to reduce the residual welding stresses.
One may notice in Figure 13.32 the heavy deep !-section steel loadi ng rings of the test rig.
whose purpose was to transfer the concentrated jack loads into evenly distributed load at the
upper and lower edge of the stiiTened cylindrica l she lls.
As part of the U.K. Department of Energy Stiffened Shell Study for Offshore Structures.
large welded steel shells were tested to complement the small scale tests discussed earlier.
Figure 13.33 shows such a 1.14-m C3.7+ft) diameter ~tringer stiffened shell GU I. one of
the ;hells tested at Glasgow University [ 13.7f. This shell (with Rlt = 275) \\3S an 1/8 scale
model of the actual otTshore component and wa.~ compared with the 1/20 scale ·smaller
shells tested at University College and Imperial College. The Glasgow Uni versi ty shell; were
fabri cated from plate section~. rolled to the requ ired rad ius. tack welded and then re-rolled
to improve circularity. Then the st ri ngers were welded on. Residual strains and geometrical
in itial imperfections were measured. The comparisons with the small Un iversity College,
London, tesLs [13.88] showed some similarity in imperfect ion patterns. residual stress levels
and buckling stresse~.
This Glasgow Univer;ity program included three large >!ringer-stiffened shells. one of
which was a three-bay shell. in an attempt to obtain more realistic boundary conditions for

Fi~:urc 13.32 Large Del nor<ke Vcriws 2.5-m diameter aluminum alloy stringer-stiffened cylindrical
--`,`,`````,`,````,,``,``,

shell and axial comprc"ion test rig (from ( 13.7 1J)


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
988 Stfffened Shells

(a)

(b)

Figure 13.33 Large Glasgow Universit)' 1.14·m (3.74·11) diameter steel stringer·stiffened cylindrical
shell GU I (courte~y of Professor 0. Faulkner): (n) shell prior to testing. (b) details of
stringers and end ring

the middle bay. A few years later. testing of large ring-Miffcned shells continued at Glasgow
University, as part of a joint program with CB I of Plainsficld, ntinois, mentioned in Section
13.2.3 [ 13.75] and ( 13.76(. and further tests on stilfencd shells have been carried out.
Large stiffened shells were also tested at Imperial Co llege. London, as part of the U. K.
Department of Energy Offshore Structures Study ( 13.9(, ( 13.68], and [13.69). complementing
their experiments on small-scale models (13.9]. [ 13.68], [ 13.69). [ 13.89] and [ 13.90]. Three
of the IC large shells were ring-stiffened (two of them three-bay models) and one (IC4) was
a three-bay shell with 20 !>~ringers. The three bay; were again intended to result in more
realistic boundary conditions for the center bay. which was the actual test specimen.
The fabrication details of this 1.2-m (3.9+-ft) diameter stringer- and ring-stiffened shell.
IC4. arc shown in Figure 13.34. It had an (Rit) 170 at the center bay, an (Rit) = 150 at
the slight ly thicker outer bays, whi le (LI R ) = I. II for all three bays. ' 'The first stage in the
fabrication process was the butt welding of the fl at panels. Then the combined plate was
rolled, bull welded along the longitudinal seam and re-rollcd to enhance its singularity. Next
the end flanges were auachcd. The ri ng-stiffeners were made up in sections of one quarter
the circu mference and anachcd to the cylinder using intenn inem fi llet welds." The process
for the ring-stiffened shells was identical up to this part. For IC4 tbe stringers were welded
on at this stage, using again rclathely light intenninent welds. The central bay stringers
were anached first. and then those in tbe outer bay;. In each case diametrically opposite
stringers were welded on alternatively. The MI G welding process was used in all the welding.
Residual strain readings were taken after each major stage of fabrication, usi ng an elec-
trical extensometer of the type developed at Cambridge University [8.68 ]. The measurements
were taken across I 00-mm gage lengths, defined by small punch marks. A large increase in
circumferential strain readi ngs was observed during roll ing. This was primarily due to the
reduction in the linear distance hetween gage marks when the Hat plate was bent to an arc.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Model Fabrication for Stiffened Shells 989

48x4·0mm stringer

''[\,50~ ~
1200
"'--lonQ~d !
./ 4·0mm plate
-.---~i~l~,~~~~r=~~~~~~~~r=l~l=nl=
(Grode 50)
IIi ! I i i I I II
666
I I I I I II ~ b~tt weld

I i i I I II 60 x 3·5
mm web 8
'ill\. (Grade 43) ,~~ 13·5 mm plate
I I ~ 1 I I l
\!_v / (Grade 43)

I
I I

i
I

I
I

I I
I I I

I I
3 mm intermittent
fillet weld 70weld 80mi r ~

I
I I
666
I i I I I I 48 x 3 5 m'm stringer

lI r I i I I I~
(Grade 43)

+-- I I I
I
I
I
I
I

I
I
I

I I
I
I I

I I
I

656
i\ I I I I I :
1 I I i I I r I I
I I / f?~lliOrnm ri~
__j_____ I I l I I
1'.360

Figure 13.34 Large Imperial College 1.2-m (3.94-ft) diameter steel stringer-stiffened cylindrical shell
TC4-fabrication details (from [13.68])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

and reached the order of I 000 f.Le. The remaining fabrication produced further average
changes in the hoop strain of about 80 f.Le, with some individual readings of up to 350 f.Le.
The longitudinal strain readings were only little affected by the rolling, reached an average
of about 85 f.Le, but larger ones were accumulated during the attachment of the stringers.
Finally, at completion of the model, the longitudinal strain was fairly uniform throughout
the shell, with an average value of about 350 f.Le, which was 21 percent of the yield strain.
Initial imperfection readings of the shell surface were obtained by rotating the cylinder
against a vertical bank of transducers. The readings were analyzed and referred to the best-
fit cylinder in the manner discussed in Chapter 10.
In the United States extensive test programs of large stiffened shells have been initiated
by the American Petroleum Institute (API), the American Bureau of Shipping (ABS) and
the offshore industry. Most of these programs have been completed (see for example [13.72],
[13.75], [13.76] and [13.91]), and more are in progress. For example, a major program
sponsored by Conoco and ABS, discussed already in Section 13.2.3, involved 66 ring- and
ring- and stringer-stiffened cylindrical shells of different geometries, tested under combined
axial compression and external pressure.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
990 Stiffened Shells

An example of large fabricated stiffened cylindrical shells for other civil engineering uses
is the longitudinally stiffened steel cylinders (for conveyor galleries) tested in bending at the
University of Alberta, Edmonton, Canada [13.92]. Four 1.27-m diameter shells, of (R/t) =
133-206 (see Figure 13.35), were tested in bending in a test rig discussed in Section 13.4.2.
All test specimens were fabricated cylinders, longitudinally stiffened over part of the cir-
cumference with rectangular hollow structural sections (HHS). The number of stiffeners
varied from three to eight, and the corresponding stiffener spacing s = 180-140 mm. The
dimensions of the stiffeners were h = b 25.4 mm and t, 2.35 mm, except for the shell=
with three stiffeners, where h = 50.7 mm, b = 25.3 mm and t, = 3.04 mm. Each specimen
consisted of a 1300-mm-long central test portion and two shorter and thicker end segments.
The purpose of the end segments was to preclude local buckling near the end boundary due
to nonuniform stress distribution. All three segments had the same outside diameter and were
joined together by circumferential groove welding.
As described in [13.92], the cylinders were fabricated from hot-rolled steel plates by a
local fabricator employing its standard manufacturing techniques. The steel plate was cold-
rolled several times until the desired curvature was reached. The two edges of the curved
plate were then welded together by complete penetration groove welds. No springback re-
straints were used, to eliminate the so-called "locked-in" stresses that could otherwise arise.
After each specimen was delivered to the laboratory, light interior fixture frames were in-
stalled into both ends of the shell to maintain the circular cross-section.
The rectangular HSS section stiffeners were then attached to the cylinders with 5-mm
intermittent fillet welds, each 50 mm long and separated by I 00-mm intervals. The location
of the intermittent fillet welds on a given stiffener alternated from side to side.
Considerable attention was paid to initial imperfections, and they were carefully measured
on two of the test specimens. The imperfection measuring device employed consisted es-
sentially of a radial rod, with an LVDT at its end, attached at the other end to a trolley that
could move axially on an aluminum bar, whose ends were supported at the centers of the
interior end frames of the shell in a manner permitting it to rotate about its axis. Thus, the
LVDT output provided the difference between the inside shell wall and the reference circles
drawn by the rotating LVDT at the succes,ive axial positions of the trolley. The imperfections
were measured after the stiffeners had been welded on but before the end plates were welded
to the shell. The test configuration was, however, simulated by suspending the specimen
horizontally from a crane for the measurements. The imperfection data were analyzed in a
manner similar to that outlined in Chapter 10, but it seems that the total of 192 measurements
(16 circumferential points at 12 axial locations) taken for each of the two shells did not
provide a fine enough mesh.
The longitudinal residual stresses, which are the most important ones here, were measured
by Demec gages (with gage lengths of 5; mm) on the outer surface, as well as with strain
gages attached to the inner surface of the test cylinder. Again measurements were limited to
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

two specimens, the thicker one and one of the thinner ones. In the thicker shell (SB2) average

Thick end Thick end


Test segment
se(lment segment

I· 1270 0.0. ~-68_ _._:_._ _ _:=o~""~~"------·1_·__7_68_ _ j


Figure 13.35 University of Alberta bending tests on large longitudinally stiffened cylindrical steel
shells~schematic of test specimens (from [13.92])

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Sti ffened Cylindrical Shells Subject to Axial Compression 991

compressive residual mains of 35 percent of the yield strain were observed in the stiffener
area and about 15 percent of r., in the remainder of the • hell. In the thinner shell (SB3a)
higher average compressive residual strains were found, about 55 percent of e, in the stiffener
area and about 30 percent of e, in the remai nder of the shell. The residual stresses resulting
from the welding of the st iffeners were found to be substantial. These compressive residual
stresses have a signi ficant ciTcct on initial yielding and shell buckling, an effect that may be
augmented when combined with that of the initial imperfections.

13.3 Experiments on Stiffened Cylindrical Shells Subject to


Axial Compression
In some of the earlier d iscussions on shell buckling. it was poi nted out that the test rigs and
test procedures for unst ilTened shells can serve equ:tlly wel l for experiments on stiffened
shells. except that for the Jailer more sti ffness is usuall y requi red. Hence there is no definite
dividing line between test systems for stiffened and unstiffened shells, and therefore the
discussion also spreads over Chapters 9 and 13. The emphasis in the following sect ions will
therefore primarily be on 'orne of the special features of the test setups developed for stiff-
ened ~hells. But referrals back to Chapter 9 will often be necessary.

13.3. 1 Technion Axial Compression Tests


A~ set o ut in Chapter 9 and in Section 13.2, very extensive experimental programs o n the
buck ling of ring- and stri nger-integra lly stiffened cyli ndrical sheUs have been ca rried out at
the Technion Aerospace Structures Laborato ry in the la~l decades. The Techni on experiments
focused on general instabili ty. local buckl ing being considered only a limi ting case. T he
specimens were therefore designed with stiffener spacings that ranged from very close to
just wide enough to make local buckling remotely po~~ible.
The test equipment and techniques \\Crc already briefl) touched upon in Chapter 9. The
earlier test setups, like that prescmed in Figure 13.36. were fairly simple and conventional.
except for the efforts to record the onset of buckling. di~cussed in Section 9.5 of Chapter 9.
T he later experimental setups (sec for example Figure 15.21 ). which included measurement
systems for vibration correlat ion (to determine no ndeslructively the actual boundary condi-
tions). that are discussed in Chapter 15 were however more sophisticated.

n .:ure 13.36 Technion experiments on buckling of intcgmlly stiffened cylindrical ~hell s subject to
axial comprc"ion early test setup (frum )9.2 12)). Note the large size of the multi-
channel strain plollcr typical of the late 'ixtic'
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
992 Stiffened Shells

The basic test setup (shown in Figure 13.36) consisted of a rigid load frame and a 23-ton
hydraulic jack (or two jacks with a connecting beam in some cases), controlled from an
Amsler universal testing machine. The load was transferred to a central shaft with a thrust
bearing, on which the lower supporting elise fitted. The upper supporting disc of the shell
reacted against a calibrated axial load cell.
A system of guide pin and mating sleeve was used in the earlier tests on stiffened cylin-
drical and conical shells, but it was discarded in the later tests because of possible friction
load sharing with the shell. For most of the tests the stiffened shell itself was therefore the
only means of alignment and preservation of load concentricity, which, however, introduced
all the axial alignment errors accumulated in the fabrication process of the specimens. Half

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
of the many strain gages bonded to each of the shells were therefore directed axially, and,
further to confirming elastic behavior up to buckling, they assisted in adjustments for an
even distribution for the applied load and in compensations for possible misalignment.
As may be noted in Figure 13.14, which shows the specimen support and end conditions
for the stringer-stiffened shells, the specimens were not clamped to the supporting disks.
They were just put on the lower disks, which had a low central location platform with a
clearance of about h to 2h (once or twice llhe thickness of the shell), and the similar top disc
was put on the top of the specimen. The boundary conditions were therefore not far from
classical simple supports, somewhere between SS3 and SS4, with very small rotational re-
straint.
About 48 strain gages were bonded to each specimen, except for the short shells, some
of them as close as possible to the edges of the shell to study edge effects. As pointed out
in Subsection 9.5.4 of Chapter 9, the strain gages provided indications of the onset of buck-
ling. The gages became "lively" at many locations simultaneously, lending support to the
hypothesis of a complete periodic pattern of initial buckling that could differ significantly
from the stable postbuckling one. It was also emphasized there that with the convenience of
modern data reduction methods extensive strain gaging is well worth considering for shell
buckling tests.
It may be of interest to examine the behavior of one of the axially compressed stringer-
stiffened shells tested in the seventies, Shell AS27M1 [13.52]. The test specimen was a short
shell made of 7075-T6 aluminum alloy, with (L/ R) = 0.648, (RI h) = 618 and relatively
heavy rectangular section stringers with (A 1 I bh) = 0.760, (e 1 I h) = - 3.86, (/ 1 J bh') = 2.86
and significant torsional stiffness T/, 1 = 23.5 ( T/, 1 being defined as G/, 1 I bD, where /, 1 is the
torsion constant of the stringer cross-section).
Koiter's measure of total curvature of Eq. ( 13.8), which determines the initial postbuckling
behavior of the panels between the stringers, was for this shell e = 0.807. As shown by
Koiter [13.50] for perfect panels with simple supports, or stringers with zero torsional stiff-
ness, the change from stable plate-type behavior to unstable cylindrical shell-type behavior
occurs when e has increased to 0.64, and after the transition the postbuckling tangent changes
its direction rapidly. In practice, however, the limiting value of e = 0.64 is a conservative
value that will be raised by the torsional stiffness of the stringers (see [ 13.50], [9.212],
[13.37] and [13.93]). But even for stringers with large torsional stiffness, analysis showed
that the panels have to comply with e < 0.8 for linear theory to apply and for stable plate-
type behavior to occur (see [13.37] and [13.93]).
Three stringer-stiffened Technion shells AS27Ml, M2, M3 [13.52] were therefore designed
to investigate experimentally this limit of e. The three specimens were machined from one
blank and thus were a triplet of short shells, exhibiting only small (though still significant)
variations in the measured geometric parameters. The behavior under axial compression of
one of these shells, AS27Ml, is examined in Figure 13.37, which shows the shell at suc-
cessive loads. The side of the shell facing the camera was kept nearly clear of strain gages,
and strong lighting was employed to emphasize the formation of local buckling waves. At
axial load of 1600 kg no buckling can yet be detected, except for some kind of edge crushing,
partly due to the shell protruding beyond the stringers (see detail A in Figure 13.14). At
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Stiffened Cylindrical Shells Subject to Axial Compression 993

(a) 1600 kg

(c) 2800kg

(d) 3200 kg

(e) 3400 kg

(f) 3500 kg

(g)3600 kg

Fit:ure 13.37 Technio n experiments o n buckling of integrally >tringer-stiffeoed cylindrical shells subject
to axial comprc\\ion ~ buckling sequence' of a \ he ll AS27Ml with wider panels. attbe
limiting 8 for <table plate-type behavior

2-100 l..g local plate buckling of some of the paneh io, already clearly visible. As the load
increases. this periodic plate buckl ing becomes more distinct. while simultaneously ncar the
edges. the buck les (which arc sti ll part of the periodic pauern) become more pronounced.
At 3400 and 3500 kg. the periodic plate panern covers the whole shell , with the edge buckles
dominating in most of the panels. Shortly after that. at a load of 3600 kg, sudden collapse
occurred. with a completely different general instability pallern. Hence at the limiting () ~
0.8. stable plate-type buckling still occurred in the panch•.
Similar behavior was observed in the other two ~hell ~ AS27M2 and M3, with () - 0.789
and 0.793 respectively. The three shells also exhibited high values of "linearity" (knock-
down factor). P<.." = 0.859. 0.784 and 0.925. respectively. as expected for shells with large
~tringer area ratios, (A,Ibh) = 0.760. 0.717 and 0.729. respectively.
Mo;.t of the stri nger-stiffened cylindrical shell > tc>ted at the Technion. however. were
designed to definitely fail by general instabili ty. lienee values of Koiter's measure of total
curvature 0 of their panels were relati vely low. For example. in the six axial ly compressed
AB •hells of (13.94], shown in Figure 13.17 (whi ch were tested with end condi tions simu-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

lati ng typical missile joi nt ~ and employed vi bration correlation techniques for detennination
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
994 Stiffened Shells

of the boundary conditions), the panels had IJ = 0.461 - 0.-167. They therefore failed indeed
by general instability with well-defined overall buckling modes, which could be closely
predicted by linear theory. Tile clean postbuckJing patterns of all the six shells in Figure
13. 17 (reproduced from [1 3.95) and (13.96]) are typic:~ I of such closely and fairly heavily
stringer-stiffened shells with (A,/bh = 0.59-0.70). Once their boundary condit ions had been
nondcstructively determined. as explained in Chapter 15. their buckJing loads could also be
fairly well assessed.
Relatively heavy and clo~ely spaced rings also significantly influence the postbuckJing
pattern of cylindrical shell s. This can be seen, for example. in Figure 13.38. pre~cnting the

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
postbuckling mode of an axial compressed ring-stiffened shell R0-24 with CA 1 /alr) = 0.441
(from [ 13.49)). whose two-tier diamond panem wa:. noticeably elongated circumfercntially.
indicating the pronounced onhotropic behavior caused by the ring-stiffening.
Before closing this section, one may recall that the Technion experimenL~ on st iffened
cyli ndrical shel ls were not li mi ted to ax ial compression loading, but extensive tests were
also carried out for external pressure and combined londing. These are considered in detail
in Section 13.4.

13.3.2 Axial Compression Tests on Small Welded Shells


Whereas the fabrication of small welded stringer- and ring-stiffened shell models. described
in Section 13.2.2, required considerable skill and development efforts in order to achieve the
necessary quality. their testing presented no serious problems. The shells could be tested in
a standard servo-hydrau lic m:1chine programmed to apply uniform increments of end short-
ening (as measured by three linear displacement transducers placed round the shell ), as was
done. for example, at University Col lege. London. in the early eighties (sec (13 .8]). The
displacement control ensured essentially rigid loading. which permitted study of the post-
failure characteristics. An extremely slow rate of loading was employed. The specimen was
held between a rigid stationary cross-head at the top and a movable actuator at the bottom.
This bottom planen rested on a ball seating to eliminate the effect of any slight nonparal-
lcli sm between the top and bottom ends of the shell . Since the testing machine was much
~tiffer than the thin model shells. its rigid pans were indeed rigid, as required.
During preliminary testing at University College London, it was found that the direct-
read ing displacement transducers did not provide a su fficientl y accurate end shorteni ng rec-
ord and therefore special bow transducers were developed for direct measurement of end
shortening between points on the specimen itself. These transducers consisted of bent strips
of high-strength spring steel to which strain gages were honded (see Figure 13.39). The bent
~trips were each fitted between V grooves cut on two horizontal Hat bars, one ncar the top

Fl~:urc 13.38 Technion cxpcrimcnls on buckling of intcgmlly >t iiTened cylindrical shells subject to
axial comprcs~ion-postbuckling mode of a 7075-T6 nluminum alloy ring-stiffe ned shell
R0-24 with (A 1 /a/r) = 0.441 (from [13.491)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Stiffened Cylindrical Shells Subject to Axial Compression 995

5211
bearmg
p1vot

Strain
g<Juges

Shell

Knrfe edge
p1vot ---........... . .

Figure 13.39 University College, London, small-scale welded stringer-stiffened shell tests-schematic
of bow transducer for measurement of end shortening (from [13.8]. Crown copyright is
reproduced with the permission of the Controller of Her Majesty's Stationery Office)

of the shell and the other near its bottom. Each of the bars was pivoted at one end and in
contact at the other with a hardened steel ball bonded to the shell surface. As the shell was
compressed, the flexing of the steel strip resulted in a signal from the strain gauges which
could be related to the values of end shortening by a previous calibration. Note that the
values of lateral displacements of the spring steel strut arc an amplification of its end short-
ening and that the amplification-factor decreases with increasing curvature. Thus by suitably
adjusting the initial curvature of the strip, it was possible to achieve any desired degree of
accuracy. The device had the particular merit that its output was little affected by the local
rotation of the shell surface in the vicinity of the balls. Three of these "bow transducers"
were arranged round the specimen at locations mid-way between the direct-reading trans-
ducers round the shell. Thus, in all, there were six recordings of end shortening.
The lateral deformations of the shell under load were measured with standard dial gages.
Geometric initial imperfections were measured with the shell, mounted in its end supports,
being rotated relative to a bank of linear displacement transducers. The residual stresses were
measured between finely drilled gage points 50 mm apart, using a standard Demec gage on
which the dial gage had been replaced by a proximity transducer, and a digital voltmeter
was employed to record the strains. The instrumentation could be considered typical of shell
tests in the early eighties, as was seen in Chapters 9 and 10.
The somewhat similar Imperial College small-scale welded stiffened shell models (of
[13.9], [13.68] and [13.70]) were tested in a special circular test frame 2.5 m high and 1.0
m in diameter (Figure 13.40). The frame consisted of four 125-mm thick platens acting as
a multiple sandwich for the shell model, load cells and screw jack. The 100-ton screw jack,
which was driven by an electric motor, was located between the two bottom platens and the
load was applied by reaction against the upper platen through six 50-mm diameter tension
rods. The load was transferred directly through three spherically capped load cells, which
alone supported the third thick platen, which carried the model and the transducer frame.
The load was then applied to the test shell by two smaller circular end plates, which were
supported between the two upper platens by two thick-walled tubes. Columns on races
ensured that the platens were kept parallel during loading.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
996 Stiffened Shells

threaded
tension rods

r lateral support
columns

~~
Figure 13.40 Imperial College small-scale welded stringer-stiffened shell tests-test rig (from [13.68])

Although self-reacting, the test rig was also anchored to the laboratory strong floor by
means of a heavy cross-beam. This increased the stiffness of the test rig and facilitated better
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

control of the post-collapse path. All the ::ests were carried out under displacement control.
A circular rotating framework, holding four vertical banks of transducers, was used for
efficient measurement of the deflections of the shell under load, as well as of the initial
geometrical imperfections. The remaining instrumentation was again not different from that
used in unstiffened shells and also included extensive strain gaging.
Tests in other laboratories on small-scale welded stiffened shells, subject to axial com-
pression, were usually carried out in the same test setups as the unstitiened circular shells
discussed in Chapter 9 (see for example [9.169]), except when the models became larger
and the investigators had to resort to special test rigs (see for example [13.71] and [13.97]).

13.3.3 Testing of Large Shells Subject to Axial Compression


Some of the earlier axial compression tests of large stiffened cylindrical aerospace shells
were carried out in large-capacity testing machines, such as the NASA Langley 1,200,000-
lb (~545-ton) capacity machine (see [13.98] and [13.99]). In order to obtain the load-
shortening characteristics of the shells more accurately, this testing machine was especially
adapted in the late fifties with a 300,000-lb (~136-ton) capacity hydraulic jack, which op-
posed and thus controlled the motion of the cross-head, thereby overcoming its inertial lag.
However, later NASA tests on large stiffened shells subject to compression were mostly
bending tests, discussed in Subsection 13.4.2 (see also [9.247], [9.249]-[9.252], and
[13.100]).
As an example of a special otishore industry-related test rig for large stiffened cylindrical
shells subject to axial compression, the Imperial College setup developed for its series of
large shells IC1-IC4 [13.9] and [13.68] i;: now described.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Stiffened Cylindrical Shells Subject to Axial Compression 997

Tn the test rig (shown in Figure 13.41), the vertical loading was provided by 24 jacks,
each of 20-ton rating, that were equally spaced beneath the cylinder's circumference and
were all fed from a conunon pressure supply to ensure a uniform concentric loading. Because
the jacks were directly beneath the cylindrical shell, a relatively light bouom platen was
sufficient to distribute the load. At the top, however. the reaction against the heavy overhead
cross-frame was through a central spherical bearing and the top platen had to be substan-
tial to minimize bending distortions. The cylinders were bedded onto the platens using an
Araldi te-and-sand mixture to ensure a uni form load appl ication.
A rotating transducer frame was employed to measure the out-of-plane deflections during
the test. The frame, which carried two vertical banks of transducers, rotated on an accurately
mach ined steel ring bedded onto the top platen and could be located at 20 predetermi ned
positions using a spring-loaded plunger. With this method, a repeatability of readings to
withi n 0. I mm was achieved.
A large number of strain gages were used in the tests to investigate the distribution of
strains in the shell, stringers and rings . Residual strai n and initial imperfection measurements
of shell IC4 were outlined in Subsection 13.2.5.
It may be of interest to briefl y describe the test procedure on this stringer-stiffened shell.
The autograph record of the testing machine for IC4- a continuous plot of the load against

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
the deflection of the loading platen-is shown in Figure 13.42. After an initial '' bedding
in," the load-deflection behavior was linear.
Because the model had already been loaded to 160 tons in an elastic test and some plastic straining
had occurred. on reloading a small change in slope was observed when this load was exceeded.
'TI1e first part of the test. up to 320 tons, was carried out under load control. After each load
increment, a small amount of creep occurred, which is rep,·esented by the short horizontal sections
of the graph. AI 320 tons the machine was changed to displacement control, and from then on the
model relaxed at the end of each increment: the ve11ical sections of the graph represent this shedding
of load.
Visoal inspection of the model first revealed the imminence of failure at a load of 41 0 tons: a
small buckle had appeared near Lhc top of a panel almost diametrically opposite the longitudinal
weld. A slightly larger load fall-off can be seen on the autographic record. However. the model
remained stable and testing continued until the load reached 420 tons or 0.87 times the squash

Figure 13.41 Imperial College tests on large axially compressed stiffened cylindrical shells-test rig
Copyright Wiley fo r shells IC I-IC4 with IC4 in it (from [13.68], courtesy of Professor J.E. Harding)
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
998 Stiffened Shells

.... ....,
p

... ,_

-- ,.. /
I/--: ~

, .. / 1....-1--v-
~
,.. I r r-.....
-
... I " '"'
.,. I : I

I I

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
/
"" 1/
00 / f-
I / - r---
.. _/ v
"'
- v . ..
/

(a) 0
•• .. • .. .. "
,.A,_

(b)

Figure 13.42 Imperial College tem on large axiall} comprc'scd stiffened cylindrical shells: (a) load
deOection plot for ~hell IC.I (from 113.68]). (b) buckled shape of shell IC4 (coune'y of
Professor J.E. Harding)

load. New buckles grndually formed about 90• around fmm the first one and then linked up to
form a ''chequerboatd'' pancm of local panel buckle;, extended around approximately one third of
the circumference. The load d1·oppecl to 340 ton;. The unloading part of Lhe curve was obtained
by applying further deOection increments. which incrca;cd both the magnitude and cxtcm of the
buckles.
One should recall that the failure here was local panel buckling and not general instability
(as was to be expected for the relatively widely >paced 'tiiTcning. which resulted in Koiter's
9 = 4.11 >> 0.8). Hence the postbuckling behavior in Figure 13.42 was indeed not violent.
One may also note that these local buckles initiated ncar the top of a panel. but then
moved to the cemer. The final buckli ng mode could therefore diller signi ficanrly from that
at onset of buckling.
Many other similar tests o n large sti ffened cylindrical ~hells subjecL to axia l compression
were carried out in and for the offshore and marine indu~try, some of them also under
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Cylindrical Shells under External Pressure 999

combined loading. One pronunent test setup. developed for ring-stiffened cylindrical shells.
subject to axial compression and external pressure. at the Chicago Bridge and Iron Company
(CB I) in the late seventie~ and early eighties. was discussed in Chapter 9. Sub~cction 9.8.2
(sec Figures 9.89 and 9.90). Another test setup, developed at DnV (Det norske Veritas) in
Norway in the late sevent ies (see Figure 13.32), was hricfl y cons idered in Suhsection 13.2.5
of this chapter; while si milar experimental progra111s carried out in Japan in the eigh ties (see
for example Figure 13.43 and [ 13. 1011 and (13.102]) were discussed in [ 13.103] and com-
pared there with American and European tests. Note that in the MiL~ubishi Heavy Industries,
Japan. test on ring-stiffened cylindrical shells. ~hown in Figure 13.43, the loading was not
axial compression but a characteri,tic earthquake loading (a combination of tranwe~ ~hear
and bending). The extensive four-year test program r 13.1 02(. however, included also ring-
stiffened shells subjected to ax ial compressive loading.

13.4 Experiments on Stiffened Cylindrical Shells under


External Pressure Bending and Torsion
13.4.1 Ex1emal Pressure Tests
A~ pointed out in Chapter 9. external pressure is one of the prime causes of in~tability of
cylindrical shells. and indeed buckling under external pres~ure has been the motivation for
the early development of strengtheni ng by ring-stiffeni ng. External pressure loading is usu-
ally hydrostatic pressure that includes in addition to the lateral (or radial) pressure also an
axial cmnponent, and therefore represents a type of combined loading.

(a)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(b)

Figure 13.43 Mitsubishi Heavy Industries. Japan, tc>ts " " ring-"iiTened cylindrical shell~ subject to a
typical cao1hquakc loading. a combination of trnn.verse >hear and bending (from [ 13.10 I]
and courtc>y of Dr. T. Yuhara of MHI): (a) tc>t ;ctup, (b) finally buckled specimen,
buckling having occuoTed between ring-stiffeners
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1000 Stiffened Shells

Many of the "classical" experiments of the twenties and thirties were actually on ring-
stiffened shells, such as those at the U.S. Navy David Taylor Model Basin (DTMB) (sec
[9.34] and [13.54]) or in particular those of Tokugawa in Japan [9.37], which probably
represent the first comprehensive study of buckling of ring-stiffened shells.
As a result of their primary interest i11 submarines, the many buckling experiments on
stiffened shells carried out at the DTMB in the fifties and early sixties focused on external
pressure loading. A typical DTMB test serup of the fifties is shown in Figure 10.7 of Chapter
10. Though built over 40 years ago, this test rig was advanced in concept and can be
considered a forerunner of similar later test setups (such as those of [13.60], [13.64], [13.104]
and [13.110]).
The 37-in. diameter pressure tank, in which the tests on a series of welded ring-stiffened
cylindrical shells and on one of machined shells were conducted, is shown in Figure 10.7.
The shell is suspended in the pressure chamber from a top mounting ring, while its lower
end is fixed in a stiff end bulkhead. Hence complete visual inspection of the shell during
the test was possible, as well as the insertion of a deflection measurement device, the de-
flectometer. This automated recording deflectometer, described in detail in Section 10.2 of
Chapter 10 (see also [13.53]), was essentially a circularity measuring device mounted on a
spindle fixed along the axis of the model, with the radial displacements at the probe being
transmitted electrically and amplified to actuate a pen in proportional motion. The probe and
its axial movement and position indicating system can be seen in Figure 10.7, and a typical
plot of the radial displacements obtained with the recording deflectometer is shown in Figure
10.8.
The concept of methodically measuring the growth of initial imperfections, and even
recording them automatically, was rather advanced for the time, the early fifties (see Chapter
10), though today it has become quite routine and the data acquisition and reduction more
sophisticated. The principle of a rotating and axially advancing probe is, however, still the
essence of most modern shell imperfection or deflection measurement devices. The conclu-
sion of Wenk et a!. [13.54] "that the initial shape of externally loaded shells significantly
affects pre-buckling behavior; it may be supposed of equally vital importance when com-
puting buckling strength" was very up to date.
It also led them to another series of tests on carefully machined models, whose initial
imperfections were expected to be minimal and were indeed very small, the initial departure
from circularity being less than 5 percent of the thickness of the shell. However, for these
near perfect specimens incipient, non-axisymmetric, lobes appeared prior to buckling (at
about two-thirds the buckling pressure). These initiated with such small amplitudes as almost
to be obscured by the experimental errors and were actually easier to identify in the con-
current strain measurements, but they too grew rapidly, nonlinearly, as the buckling pressures
approached.
As is apparent from the discussions in Chapters 3 and 10, Wenk and his colleagues [13.54]
were only at the beginning of the right road. Today, if one has measured the shape of the
initial imperfections, one can compute (with available codes) the actual buckling loads or
pressures of an imperfect stiffened cylindrical shell.
In the mid-sixties two series of external pressure tests on aluminum alloy ring-stiffened
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

shells that failed at the low values of pressure of interest in aerospace structures were carried
out at NASA Langley Research Center.
First (see [13.105]) 10 small 14-in. (== 356-mm) radius and 17-in. (==432-mm) long 2024-
T3 shells, with different skin thicknesses (Rit = 264-421), and varying ring spacings (4-7
rings) were tested inside a heavy steel tank subject to a uniform external air pressure. At-
mospheric pressure was maintained inside the test cylinders throughout the test by means of
a vent tube. In one specimen, however, the larger differential pressure needed to fail the shell
was obtained by connecting the vent tube to an air ejector, which lowered the pressure inside
the test shell to less than 30 percent of an atmosphere. The tank had a large window for
visual observation and photography of the buckling process. The shells were instrumented
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Cylindrical Shells under External Pressure 1001

with strain gages, and in some specimens the axial shortening between the end rings was
also measured. The differential pressure was measured during loading by an electric differ-
ential pressure transducing cell mounted on the test cyli nder end plate. All the data were
recorded on a central di gital recording system. However, no imperfection or deflection scans
were taken.
The test cylinders were designed such that the local buckling between rings would sig-
nificantly precede collapse, and indeed it occurred at 48- 85 percent of the collapse pressure.
With increase in pressure beyond buckling, the cylinders still had substantial load-carrying ·
capabi lity and the buckling deformation grew. In that process many cylinders experienced
one or more changes in the number of buckles (a decrease in the number. as observed also
in other stiffened shell buckling experiments). Just prior to collapse of the shells. the buckles
were extremely deep (see for example Figure 13.44).
Then, a year later, a large 10-ft (ar3-m) ring-sti ffened 2024-T3 aluminum alloy cylindrical
shell (with Rlt = 1489 and LIR = 1.58) was tested under external hydrostatic pressure (see
(13.106]). The rings in this shell were closely spaced to ensure that general instabili ty should
occur before or at the local buckling load. a configuration having been shown to be struc-
turally more efficient. The test shell had eight longitudinal wall splices. consisting of lap
joints held together by double lines of seam welds. and its 36 reinforcing hat-section rings
were spot-welded to the skin of the cylinder.
The initial imperfections (deviations from circulari ty) were already measured with a spe-
cial scanner. which autographically plotted them as a function of axial location on the cyl-
inder as the scanner moved along one of its generators. The input to the two axes of the
plotter was obtained from an LVDT probe and a potentiometer indicating the vertical loca-
tion. Plots were taken around a quarter of the ci rcumference of the shell, at 6-in. ( 152-mm)
intervals, showing average imperfection ampli tudes o f about one ski n thickness, but reaching
more than three skin thickness as at some locations. These imperfection measurements (in-
itiated more than a decade after the DTMB ones mentioned earlier in this section) were
systematic, though still far off the modern automatic schemes discussed in Chapter 10.
The shell was loaded by slowly evacuating air from inside the cylinder, using a mechanical
vacuum pump. The specimen failed catastrophi cally by general instabili ty at 76 percent of
the predicted pressure, a rather low (and not fully explained) result. However, all the strai n
data indicated considerable nonlinearity at loads approaching the ultimate one, similar to the
overall " becoming lively" of strain gages, mentioned in Chapters 9 and 13 for the axially
compressed Technion stiffened shells.
The Technion external pressure experiments on stiffened cylindrical shells [13.107] actu-
ally focused on combined loading of external pressure and axial compression, the aim being
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 13.44 NASA Langley buckling and postbuckling tests on ring-stiffened aluminum alloy cylin-
drical shells subject to unifonn extemal pressure-deep postbuckling dcfonnations of
shell no. 10 just prior to collapse (from [13.105]. courtesy of NASA Langley Research
Center)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1002 Stiffen ed Shells

improved interaction curves. They also employed the vibration correlation technique (VCT),
discussed in Chapter 15, for nondestructive defi nition of the actual boundary conditions. T he
experimental test setup. shown in Figures 13.45 and 13.46. therefore also incl udes a system
for detection of the natural frequencies of the shell and for scanning of the circumferential
and longitudinal modes of vibration. required by vcr.
The rest setup was designed to provide axial loading and external pres~ure separately or
as combined loading. The test shell is clamped at both ends to rigid rings (sec Figure 13.46b).
The upper ring is connected to the lid of the pressure vessel. The lower ring is an integral
part of the bottom plate, which is connected via a shaft . a uni versal joint (8 in Figure 13.45)
and a load cell (no. I in Figure 13.46a) to the loading bridge (A in Figure 13.45). T he bridge
is supported by two screw jacks (no. 3 in Figure 13.46a). Ax ial loading is applied to the
shell by lifting the loading bridge with the screw jacks. since this raises the bottom plate.
which presses the shell against the upper end plate. The universal joint in the shaft ensures
pure axial loading without bending. The connection between the loading bridge and the
screw jacks also assists pure axial lo:tdi ng. The externa l pressure is applied as air pressure,
supplied via a control valve and measured on a manometer (C in Figure 13.45).
For vibration correlation the shell is excited by direct contact. Though the vibration ex-
ci tation, detection and scanning system is described in detail for axial compressed sti ffened
shells in Chapter 15. it is briefly outlined here for the case of external pressure. for com-
pleteness. An electromagnetic shaker. Pye Ling V.G. type 47/30 (no. 6 in Figure 13.46a),
is glued to a small rubber pad, which in tum is glued to the inner wall of the shell. The
shaker is suspended on three springs connected by a rubber base. which rest on the bortom
plate (sec Figure 13.46b). The purpose of this floating support is to cancel the influence of
the small mass udcleclto the shell when the shaker is attached to it.
Resonance is identified by Lissajous figures on the oscilloscope (Di n Figure 13.45). The
circumferential and axial vibration modes arc scanned by moving the microphone (no. 19
in Figure 13.46a) circumferentially and axially and recording the output on an X- YY recorder
(E in Figure 13.45 and no. 12 in Figure 13.46a).
The bending and compression stresses were measured by five pairs of strain gages anached
to the inside and outside of the shell at about one-third of the height from the bollom plate.
The test procedure was as follows. The shell wa.~ loaded with an axial load or external
pressure or a combination of both. The strains were read and recorded. Then the shell was
vibrated, the natural frequencies were noted and the modes were scanned and recorded. The
procedure was repeated for each loading step.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 13.45 Technion tests on integrally stringer-stiffened cylindrical shells under external pressure
and combined loading-experimemal setup thlll also includes the excitation. detection
and scanning instrumentation for vibration corrcl:llion (from [13.1071>
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Cylindrical Shells under External Pressure 1003

-----------GJ
1. LOAO CELL
2. BRIDGE CIRCUIT
3. SCREW JACKS
4. BRIDGE CIRCUIT
5. STRAIN GAGES
6. ELECTROMAGNET! C SHAKER
7. AMPLIFIER
B. PRESSURE VESSEL
9. CONTROL VALVE
IO. OSCILLATOR
I 1. FREQUENCY COUNTER
11. X - YY RECORDER
13. AXIAL POTENTIOfiETEn
14. CIRCUMFERENTIAL POTENTIOMETER
15. OSCILLOSCOPE
16. FILTER
17. AMPLIFIER
lB. PREA~trLIFIER

19. fHCROPHONE·
20. MAI~OMETER

21. SHELL
22. ENIJ "LATE

(a)

--SHELL SIMPLY SUPPORTED

-SHAKER

--- SUPPORT SPRING

CLAMPED

PLATE

(b)

Figure 13.46 Technion tests on integrally stringer-stiffened cylindrical shells under external pressure
and combined loading-schematic diagram of experimental setup including vibration
correlation system (from [9.273]): (a) diagram of setup, (b) detail of boundary conditions
of shell and shaker attachment

Two groups of stringer-stiffened shells were tested in the first test series (of [13.107]): one
of the four steel specimens and one of six aluminum alloy ones. Additional test series
involving 17 aluminum alloy stringer-stiffened shells followed after a few years [9.273] and
[9.274]. The same manufacturing technique employed for the axially compressed Technion
shells and detailed in Section 13.2.1 was used here.
The specimens were either nominally clamped or simply supported to the loading rings
in circular grooves fill with Cerrobend (as shown in Figure 13.46b). The actual boundary
conditions were then determined by application of the vibration correlation technique and
VCT was also used to obtained interaction curves, as explained in Chapter 15.
The final experimental interaction curves were obtained for all the aluminum alloy shells
by the repeated buckling approach, described for unstitiened shells in Section 9.8.3. As
pointed out there, this approach is usually successful if elastic buckling is assured, which is
often the case for thin shells. For thin closely stiffened aluminum alloy shells, elastic buckling
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1004 Stiffened Shells

still dominates in the initial phase of buckling. Thus. when care was taken to detect the
onset of buckling. the repeated buckl ing approach also gave good results here, though some
minute inelastic deformations occurred, as was observed by repeated imperfection scans (see
(9.274]).
The buckling studies of Walker and hi~ associates at the University of Surrey in the eighties
focused on stiffened shells under external pressure and simultaneous pressure and axial load-
ing (see [13.108)). A special hyperbaric test rig (see Figures 13.47 and 13.48) was built for
the experimental pan of the research.
The facility combines a self-reacting frame with a very heavy flange to whose upper
surface a hyperbaric chamber is fitted (sec Figure 13.48a). Axial load is applied by an Jnstron
500 kN servohydraulic actuator, situated below the chamber. on which the ram displacement
is controlled by a signal from a 1-mm LVDT mounted on the frame of the test rig. The
sensitivity of this transducer is bener than 0.12 percent. and hence the end displacement can
be controlled to 10 ' rnm. which for a 50-mm long steel model represenL~ only I percent
of its yield strain.
In the tests the shells were located on a machined mounting ring fitted to the heavy flange
(sec Figure 13.48a). The axial load was transmitted to the specimen from the actuator by
means of a spherical seating. which could acconunodate uny lack of parallelity of the end
plates (see Figure 13.48a and b). The spherical seating and nut can be inverted to facilitate
either axial compression or tension loading. The latter made tests under pure lateral pressure
possible by incrementally offsetting the downthrust from the hydrostatic pressure during the
pressurization phase with tension applied by the actuator.
The hyperbaric chamber is served by its own variable-speed positive injection pump and
pressure monitoring equipment with digital read-out. The head of the chamber has pons with
special seals for the strain gage lead wires and for the shaft of the submersible rotating radial
displacement measurement system (sec Figure 13.48c).
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 13.47 Univctliity of Surrey hyperbaric test rig for cylindrical shells (courtc>y of Professor A.C.
Copyright Wiley Walker)
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Cylindrical Shells under External Pressure 1005

'o· rmg seals

D!splactmcrnt loading
mcrosura:mgo"lt plate
dctVICI2 housing
Trcrtsduc~rr spherical
nut

dr:ver-anC:~~~Indpxc
Hydroul1c
octuotor glonC ~--~
_.--,~otaAy

ScriT r•octrng
fromll

(a)
~-

~500 -k.N. instron


(b) toed celt (c)

Figure 13.48 University of Surrey hyperbaric test rig for cylindrical shells-schematic diagram (from
[13.108] and courtesy of Professor A.C. Walker): (a) general view of test rig, (b) detail
of test section, (c) detail of submersible displacement measurement system

The Surrey hyperbaric test rig was used for external pressure (including pure lateral pres-
sure) and combined loading (external pressure and axial compression) tests on small ring-
and stringer-stiffened welded steel shells. Figure 13.49 presents a typical postbuckling pattern
of one of those stringer-stiffened shells (see [13.109]). The type of boundary conditions
needed for waterproofing that was used in the Surrey tests is shown in Figure 13.50. Note
that the ends of the cylinder were fitted against a small, accurately machined step in the
bottom of the groove in the end rings, and a sand-Araldite mixture was again employed to
cast the shell in position.
In connection with the Surrey facility, one should recall here the larger Liverpool hyper-
baric chamber, in which torispherical heads and other shells were tested, described in Chapter
9, Subsection 9.11.2 (Figures 9.132 and 9.133).
Before we close this section, a more recent study at the University of Waterloo, Ontario,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Canada [13.110], may be of interest. It focused on the change in buckling behavior as the
ring-stiffening on otherwise identical shells varied from many equally spaced weak rings to
a few heavy ones and thus investigated experimentally the transition from the general insta-
bility mode to the panel (or shell) instability mode.
Ten 254-mm internal diameter 6061 aluminum alloy shells, of (R!t) = 63, with rings of
widely varying cross-section area and spacing but a nearly constant area ratio (A 2 / at) =
0.21-0.26, were tested. The specimens with 2-mm wall thickness were machined from l-in.
(25.4-mm) thick tubes to reduce initial geometric imperfections. The maximum initial out-
of-roundness measured was about 0.90 mm, less than half the shell thickness, while the
variation of the wall thickness t was within ± 0.0125 mm (0.625 percent oft). The test rig
was fairly conventional, with oil as pressuring fluid. The initial deviations from circularity,
as well as buckling deformations, were measured on the inside surface of the shells with a
number of direct current displacement transducers attached at selected heights to a smoothly
rotating central spindle. The circumferential position of the spindle was monitored by a
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1006 Stiffened Shells

Figure 13.49 University of Surrey external pressure tests on small welded steel stringer·stiffened
shells-typical postbuckling pattern (counesy of Professor A.C. Walker)

potentiometer and thus X-Y plots and more accurate circularity contours could be readily
obtained.
The Southwell method was used to estimate the buckling pressure of the corresponding
perfect shell since it was found to be consistent and convenient to apply, as was also con-
cluded in Chapter 9, Subsections 9.6.4 and 9.6.5. S ince the shells were relatively th ick, the
maximum recorded strains at collapse were fairly close to the yield strain of the material,
60-86 percent of it, and in most specimens nearer the higher value. But the o bserved buck-
ling behavior was still elastic and the Southwell method provided consistent results. In gen-
eral, the agreement with predictions by Kendrick's analysis (13.111) was fairly good for
bo th general instability and shell instability modes, the predictions bei ng co nsistently 12- 33
percent below the Southwell estimates.

13.4.2 Bending Tests


As mentioned in Subsection 13.3.3, extensive experimental studies on the bendi ng of large
sti ffened shells were carried out at NASA Langley in the fifties and sixties (see for example
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
[9.247], (9.249]- (9.252) and (13.100)). The Langley rigid bending test rigs were discussed
in Chapter 9, Subsectio n 9.7.2 (see Figure 9.75) here only some particular aspects relating
to general instability are considered.
One should recall that general instability may precede possible local buckling, as is usually
the case in closely stiffened shells. or it may occur after local buckling had taken place and
at much higher loads than the local buckling loads. as in the case of widely spaced stiffening.
The latter behavior was observed, for example, in the small DFVLR Braunschweig Mylar

Figure 13.50 University of Surrey external pressure tests on small welded steel stringer-stiffened
shells-type of boundary c<:mdi1ions (from [ I3. 108])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Cylindrical Shells under External Pressure 1007

shells with widely spaced ring and stringers, discussed in Section 13.2.4 (see Figure 13.28),
whereas the former was, for example, typical of the Technion integrally stiffened shell s (see
Figures 13.17 and 13.38).
One of the earlier NASA Langley bending tests (13 .1 00] examined seven such thin dis-
cretely ring- and stringer-stiffened cylindrical shells, for which local buckling occurred very
early in the loading history (at 10- 15 percent of the collapse moment). The 7075-T6 alu-
minum alloy specimens (of Rl t E;!! 1940- 1970) continued to carry load till general instability
collapse occurred, for all but one shell that failed in a panel instability mode. The failure of
a typical shell (see Figure 13.51) clearly shows the overall instability mode, as well as the
deep local buckling deformations, which had grown with the increase of bending moment.
Later NASA Langley bending tests studied mostly large closely stiffened or corrugated
cylindrical shells. A representative example of these is the 1971 buckli ng experiments on
two integrally stiffened cylinders subjected to bending (see Figure 13.3 1 and [9.251 ]). T he
test cylinders were 1.96 m (77 in.) in diameter and stiffened by rings and stringers and were
loaded in bending through a loading frame, usi ng a hydraulic testing machine (see Figure
9.75b). Two tests were made on each cylinder, where for the second test the shell was rotated
1800 in its fixture, so that the generator subjected to maximum tensile stress in the first test
was the generator of maximum compressive stress in the second one. T he presence of stray
loads in the test cylinders was minimized by employing rollers between moving surfaces
and counterbalancing the test fixtures near their centers of gravity. Rollers were used between
the loading frame and the floor supports, as well as between the loading frame and the
testing machine. The rollers allowed the cylinders to shorten during application of load and
restricted the loads at the roller locations to no1mal loads. T he rollers and the surfaces against
which they reacted were case-hardened.
The test cyl inders were instrumented with many strain gages on stringers, rings and skins,
the strains being recorded during the tests at the Langley central digital recording facility.
One of the results was again, as poi nted out in Chapter 9. that the load intensity carried by
the cylindrical shells in bending exceeded that cruTied in uniform axial compression, here
by about 12 percent. This can be explained, as in Section 9.7.2, by the smaller imperfection

Figure 13.51 NASA Langley bending tesls on large l'ing- and slringer-stiffened 7075-T6 aluminum
alloy cylindrical shells (from [ 13.100], counesy of NASA Langley Research Center): (a)
general instability failure of tesl cylinder no. I, group I. (b) dimensions of specimens
(in mm): R = 980. L = 1829, <l = 152. 229, 305, and b = 63 (group 1). 108 (group II)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1008 Stiffened Shells

sensitivity of the shells in bending, resu lting from the fact that in bending only the imper-
fections in the narrow regions of the greatest compressive stress affect the buckling loads.
The experimental buckling loads approached the predicted ones and the observed buckli ng
patterns were close to the calculated ones.
The primary interest in bending instability of shells for civi l engineering applications has
focused on inelastic buckling, which is discussed in Chapter I 6. But in many thi n shells
elastic buckling is also prominent in bending, though failure usually also involves inelastic
effects. A modem example of bending tests on stiffened cylindrical shells related to civil
engineering is those carried out at the University of Alberta on longitudinally stiffened large-
diameter fabricated shells of the type used in conveyor galleries. stacks or offshore structures
(see [ 13.92], [13.112] and [13.113]).
A diagram of one of the four 1.27-m diameter shel ls tested is shown in Figure 13.35. The
test setup used is presented by an overview in Figure 13.52 and a two-view diagram in
Figure 13.53. It was similar to the test rig used in eartier bending experiments on unstiffened
shells at the University of Alberta [13.114]. It consisted of two major parts: a large composite
beam and a loading frame. T he composite beam consisted of two end trusses and a test
specimen between them. The test region was subjected to pure bending under a two-point
loading system. The ends of the specimen were assumed to remain practically plane because
they were welded to the heavily stiffened base plates of the end trusses (see Figure 13.53).
Each end truss was bui lt up from two parallel truss elements connected by short !-sections.
The chords of the plane ttuss elements were welded wide-Aange shapes, and the web mem-
bers bolted to them were channe l sections.
The composite beam was simply supponed at both ends, with horizontal translation per-
milled at one end. Each plane truss had its own supports, cons isting of a knife edge and a

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
set of rollers. The load was applied by means of two 2200 kN (- 225 ton) hydraulic jacks,
which were controlled by an air pump so that the two jacks could maintain the same load.
1be jacks were simply supported between the loading frame and the composite beam. Hor-
izontal rollers above the jacks permitted them to move with the loading points. A spherical
device below the jack al lowed for possible rotations.
Accurate measurement of loads and reactions was obtained by six load cells, two 2200
kN ones under the loading jacks and two 1300 kN ones and two 900 kN ones at the four
supports under the composite beam. The measured discrepancy between the loads and the
reactions during testing was within 3 percent. The applied moment was calculated from the
average of the load cell readings, after the moment caused by the gravity loads (the weight
of the end trusses and the test specimen) was added.

Figure 13.52 University of Alberta bending tests on large longitudinally stiffened cylindrical steel
shells- test setup (courtesy of Professor G.L. Kulak)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Cylindrical Shells under External Pressure 1009

(a)

Thick end segment

12 440

(b)

Figure 13.53 University of Alberta bending tests on large longitudinally stiffened cylindrical steel
shells-test setup elevation and plan (from [13.92]): (a) elevation view. (b) plan view

Displacements at various locations were measured by LVDT's, cable transducers and dial
gages. Fifty strain gages measured the strains on the specimen. All the data were recorded
on a data acquisition system driven by a microcomputer.
Eight dial gages were used to check that the base plates between the specimen and the
end trusses remained plane during the bending deformation. Small out-of-plane displace-
ments, of less than 1.6 mm, were measured, which were further reduced to less than 0.4
mm by adding four additional bolts to the connection between the trusses and the base
plates-a reminder of the difficulty of obtaining the assumed planarity in a loaded test rig!
Several other dial gages on the loading frame monitored the lateral movement of the end
trusses, which was found to be negligible.
Two shells, SB2 and SB3a (the test of which was arrested before complete failure to allow
addition of two stiffeners and retesting as SB3b ), failed by shell buckling outside the stiffened
area. The remaining two specimens, SB3b and SB4, failed by general instability buckling,
in which the stiffeners buckled gradually with the shell. The moment capacity of the latter
two shells, which failed by general instability, was significantly higher than that of the two
former ones, for only a very small increase in weight due to additional stiffeners. These
results reiterated earlier conclusions about the significantly higher structural efficiency of
closely stiffened shells.

13.4.3 Torsion Tests and Combined Loading


Though pure torsional loading seldom arises in aerospace applications, it is an essential
element of the frequently occurring combined loading of bending and torsion. Torsional
buckling of cylindrical shells was therefore extensively studied in the thirties and forties, as
presented in Subsection 9.7.3, though only very few experimental investigations were carried
out. The experimental studies of torsional instability of stiffened cylindrical shells were even
rarer.
The first systematic experimental investigation of stiffened cylindrical shells subjected to
torsion was apparently that carried out by Dunn at Caltech in the mid-forties [ 13 .115] as
part of a series of NACA-sponsored studies on the general instability of stiffened metal
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

cylinders.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1010 Stiffened Shells

The test were conducted in a 500,000 lb-in. (~56,500 N-m) capacity combined bending
and torsion testing machine, built at GALCIT (Caltech), which was a larger development of
the earlier Donnell test rig, shown schematically in Figure 9.82. In the pure torsion loading,
the bending anns and one torsion arm were locked in place, the load being applied to the
second torsion arm.
Two series of specimens were tested, one series of 22 shells having a diameter of 64 in.
(~1626 mm) and one of 15 shells having a diameter of 20 in. (~508 mm). The shells of
both series were rather thin-walled, with (Rit) = 1067-1600 for the first series and (R/t) =
500-1000 for the second one. The shells in both groups were of medium length, with (L/
R) = 4, and were made of 24 ST dural sheet. The longitudinal reinforcing stringers were
round 24 ST dural tubing drawn to an elliptical shape, while the frames consisted of 24 ST
Alclad rectangular bars.
The aim of the test program was to provide the designer with information enabling him
to estimate the allowable general instability stress of a stiffened metal cylindrical shell. Hence
the emphasis was on measurement of the applied torsional moment at failure as a function
of the geometrical variables b, the spacing of the longitudinal stiffeners, and d, the frame
spacing for a constant shell radius. This yielded an empirical parameter for predicting general
instability in torsion. Some attempts were also made to measure local buckling loads and
wave form of the panels bounded by two frames and two longitudinal stiffeners, but very
large variations were found. The local buckling that preceded the general instability was
therefore practically ignored and only the maximum torsional moment at general instability
was considered. A rather extensive and careful parametric experimental study was undertaken
to make the conclusions reliable and therefore of value to the designer. In the forties and
fifties such extensive parametric experimental studies were not uncommon at NACA to sub-
stantiate their design recommendations.
Another similar experimental study was carried out at NASA (then NACA) Langley in
the same period [9.258]. A decade later, the well-known North American Aviation experi-
mental study, discussed in Subsection 9.4.1 of Chapter 9, included a few tests on stiffened
cylindrical shells under torsion (see [9.187]).
In the early sixties Milligan and Gerard in the United States studied the general instability
of orthotropically stiffened cylindrical shells both analytically and experimentally (see for
example [13.4]), and their investigations included the case of torsion [13.116]. They tested
10 small 8-in. diameter integrally machined 6061-T6 aluminum ring- and stringer-stiffened
cylindrical shells under torsion. The specimens were proportioned for general instability
failure by designing the shear buckling stress of the skin between stiffeners to exceed the
global buckling stress and ensuring stress levels well below the proportional limit in shear.
The shells were therefore closely stiffened, with (R It) = 400, and in order to minimize the
stiffener eccentricity effects [13.18], the stiffeners were very shallow, of only about 2 t depth.
Half the shells were ring-stiffened and the remainder stringer-stiffened, mostly with external
stiffeners, except one internally stiffened specimen in each group.
The torsion test rig was designed so that an axial load P to the rig by the pneumatic
testing machine, on which it was situated, resulted in torsion loading in the test specimen
(see Figure 13.54). This simple setup achieved the torque by applying half of the axial load
to the top of one end ring while applying half of it to the bottom of the other end ring. The
reactions were applied similarly to the opposite ends while the shell was supported along its
axis.
Very good correlation was obtained between the torsion test results and linear general
instability theory. For the shallow stiffeners of these test specimens, their location was indeed
immaterial, but with deeper stiffeners the eccentricity effects would be significant and would
have to be taken account of.
Whereas unstiffened circular cylindrical shells under torsion were again extensively stud-
ied and tested more recently (see for example [9.263]), hardly any new experiments on the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

buckling and postbuckling of stiffened cylindrical shells subjected to torsion (or combined
Copyright Wiley
torsion and
Provided by IHS Markit under other
license loadings) have been reportedLicensee=McDermott
with WILEY in the literature. The fewUser=G,
Inc - India/8215328006, exceptions
Boopathi [13.117]
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Cylindrical Shells under External Pressure 1011

Figure 13.54 Gerard's torsional buckling tests on stiffened circular cylindrical shells-schematic dia-
gram of torsion test rig (from information of [13.116])

and [13.118] relate primarily to stiffened curved composite panels and could perhaps be
relegated to Section 13.6, dealing with curved stiffened panels. However, in view of their
method of testing as a triangular cylindrical shell, their discussion is more appropriate here.
The Israel Aircraft Industries and Technion experimental study in the early nineties on the
postbuckling behavior of stiffened composite panels loaded in cyclic compression and shear
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

consisted of two major phases: the first included static and cyclic loading in axial compres-
sion of flat and curved stiffened panels and of individual stiffeners, while the second dealt
with curved stiffened panels loaded by combinations of axial compression and torsion (which
is of interest here).
The flat panel experiments, carried out at Technion, are discussed in Chapter 8, Subsection
8.4.4 (or see [8.122]-[8.125]). The second phase, dealing with curved panels, included cy-
lindrical panels (0.7 m long and 0.7 m wide) stiffened both longitudinally and transversely
and is discussed here.
For the curved panels, a special test rig was designed by Technion and IAI for simulta-
neous axial and torsional loading. The test rig (see Figure 13.55) had two arms with jacks,
attached to a heavy metal end plate that was fastened to the top side of the torsion box
(consisting of three curved panels), through which the torsion moment was applied. An axial
loading jack was fixed below the torsion box, which could apply axial compression simul-
taneously with the torque. When the torsion box was subjected to pure torsion, evenly dis-
tributed shear loads were applied to the stiffened curved panels. The static loading capacity
of the test rig was 70 tons axial load and 35 ton-meter torque. The corresponding cyclic
loading capacity was 50 tons axial and 25 ton-meter in torsion. The dimensions of the test
rig were 2 m wide, 4.7 m long and 4 m high, to accommodate specimens up to 1.3 m long
and 1.0 m in diameter.
The actual test specimen was a cylindrical torsion box, composed of three curved graphite
epoxy panels, each stitiened by four longitudinal J-section stiffeners and two heavy trans-
verse !-section stitieners. The experiments were performed in three stages. In the first stage
a metal triangular cylindrical torsion box was tested in compression and shear for evaluation
of the test rig and calibration. In the second stage a torsion box consisting of one curved
stiffened composite panel and two stiffened metal panels was tested to overcome problems
of uneven strain distribution. The third stage was the test on the actual triangular composite
test specimen.
Extensive instrumentation was used. Over 100 strain gages were attached, moire grids
were placed all around, audio effects were recorded for tracing exact times of buckling and
Copyright video
Wiley filming and still shots were widely employed.
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1012 Stiffened Shells

Figure 13.55 Israel Aircraft Industries and Technion tests on stiffened composite curved panels sub-
jected to cyclic shear and compression-test setup for axial and torsional loading (from
[13.118]): (a) side view, (b) top view

The actual test torsion box was first loaded statically in pure compression, then in pure
torsion and then in a combination of both. It was repeatedly loaded to 40,000 cycles of
combinations of compression and torsion. After ultrasonic inspection, the panels were im-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

pacted at several locations, a 50-mm hole was cut in one panel, and the damaged torsion
box was retested cyclically to another 40,000 loading cycles. Measurements taken during
the cyclic loading tests showed no change in the behavior of the panels. Also, ultrasonic
inspection revealed no damage formation after 40,000 cycles and no damage growth in the
"damaged" panels during the additional load cycle block. Unfortunately, lack of funding
limited the tests to one triangular torsion box, thus diminishing the value of the test results,
which should have included additional test specimens. Some indications of repeatability
were, however, obtained from the fact that the tests included three practically identical curved
stiffened panels.
As in practice, torsional buckling experiments often include combined axial compression
and torsion loading. In general, the problems of combined loading tests are similar for
unstiffened and stiffened shells, and therefore the discussion in Chapter 9, Subsection 9.8.2,
also applies to stiffened shells. Indeed, one of the examples there, the large test setup built
at CBI (Figure 9.89), served primarily for combined loading of stiffened shells.
The repeated buckling approach for combined loading tests providing data for interaction
curves, discussed in Subsection 9.8.3, can also be applied to stiffened shells. As mentioned
in Subsection 13.4.1, the Technion external pressure experiments on stiffened cylindrical
shells [13.107] focused on combined loading of external pressure and axial compression.
The test setup (Figures 13.45 and 13.46) therefore had the capability of applying external
pressure and axial compression separately or as combined loading. As pointed out in Sub-
section 13.4.1, the repeated buckling approach for combined loading gave good results for
two of the stiffened aluminum alloy shells of the earlier test series [13.107], for six of a
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Stiffened Conical and Spherical Shells 1013

later series [9.273] and for the six shells comprising another more recent test series [9.274].
These were all the shells in those series on which the approach was tried, and all of them
yielded good results. Hence one can state that the repeated buckling approach can be applied
successfully to obtain interaction curves also for closely stiffened shells, provided care is
taken in the tests to ensure early detection of the initiation of buckling.
It is also of interest to note that a careful study of the effect of sequence of loading in
the two more recent combined loading test series [9.273] and [9.274] indicated that this
effect is negligible, in spite of the nonlinearity of the interaction.

13.5 Stiffened Conical and Spherical Shells


13.5.1 Stiffened Conical Shells
The discussion on conicity effects, definitions and boundary conditions of conical shells
presented in Subsection 9.9.1 of Chapter 9 obviously applies also to stiffened conical shells.
Ring-stiffening, which is the common type of stiffening for conical shells, may, however,
considerably influence the buckling pattern, and even counteract conicity effects, if its spac-
ing varies. For example, under hydrostatic pressure an optimum stiffened conical shell would
require non-uniform spacing of rings to ensure the necessary uniform local buckling strength
(see [13.119]).
Not many experimental studies on the instability of stiilened conical shells have been
reported in the literature, and they sometimes used the same test setups as employed earlier
for unstiffened conical shells.

13.5.2 Technion Tests on Stiffened Conical Shells


Probably the earliest systematic experimental investigation on stiffened conical shells was
the tests on ring-stiffened conical shells under hydrostatic pressure and under torsion carried
out at the Technion Aircraft Structures Laboratory in the early sixties [9.278], [9.279] and
[13.119]. The main purpose of these experiments was the verification of the smeared stiffener
theory for stiffened conical shells developed at Technion at the time [13.14], [13.25] and
[13.119].
The hydrostatic pressure tests included seven integrally ring-stiffened conical shells, three
machined from AG5-X516 aluminum alloy and four from SAE 1015 mild steel, all designed
to fail by general instability. The specimens were formed from oversized sheets, the seam
was welded and then they underwent a thermal stress-relief process. Following that, the inner
face of the still relatively thick shells was machined. Then each shell was mounted on a
conical mandrel, the taper of which corresponded to the cone angle, and the outer face of
the specimen with its integral ring-stiffeners was machined. Typical aluminum alloy shells
are shown in Figure 13.56. The seven shells had cone angles of 20°-22°, taper ratios 1/J =
0.50-0.678, average radii of curvature Pav = 98-113 mm and thickness ratios (p") h) = 145-
186.
The hydrostatic pressure tests were carried out in the test setup shown in Figure 13.57,
essentially similar to the one used for the larger unstiffened shells discussed in Subsection
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

9.9.3 except that the pressure vessel was smaller and the shells were clamped to end rings
whose mating faces matched their cone angle. This near-clamping was preferred over the
nearly simple supports used in the unstiilened conical shells, in view of the higher buckling
pressures. Agreement between experimental results and theory was good.
Additional tests were carried out on five similar ring-stiffened conical shells subjected to
torsion in the load frame employed earlier for unstiffened conical shells (see [9.279] and
[13.120]), yielding fairly good agreement between experiment and theory.
A few years later another experimental study was carried out at Technion on ring- stiffened
conical shells under axial compression [9.149] and [9.229], which was already mentioned
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1014 Stiffened Shells

(a) (b)

}' igure 13.56 Technion experiments on ring-stiffened conical shells under hydrostatic pre;sure- typical
AG5-X516 aluminum alloy specimens (from [9.279]}: (a} shell ALR-3 before test, (b)
shell ALR· I aficr test

in Chapter 9. In order to obtain shells without a scam or welded joint5, conical blanks were
fabricated by hydrospinning. as detailed in Subsection 9.3.5. To ensure elastic buckling of
the thin shells, high-yield-strength 17-7 PH al loy steel was used for the shear-spun blanks.
T he hydrospinn ing also served as a cold-draw ing process that transformed the original semi -
austeniti c microstructure into a martensitic stage. After appropriate stress-relief heat treat·
ment, a fairly high-yield-strength steel, with u, about 120 kg/mm2 (= 170.000 psi) and a
stress relief of about 80 percent, was obtained. The machining of the stiffened conical shells
from the blanks is also detailed in Subsection 9.3.5. One may, however, reitemtc that after
extensive trials. low-cutting-speed turning was found to be more accurate than grinding for
the final machining of the shells, yielding specimens with thickness variations of less than
4-5 percent of the mean.
The test setup for the ax ial compression tests was a pair of mounting plates placed in a
30-ton Amsler universal testing machine. The specimen with its end fitti ngs (sec Figure
13.58) was located between the moving compression table of the machine and its central
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

compression rod. The small end fitting was placed upon the moving table, while a 20,000

Figure 13.57 Technion experiments on ring-stiffened conical shells under hydrostatic pre>surc-test
setup (from 19.2781)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Stiffened Conical and Spherical Shells 1015

Figure 13.58 Technion tests on ring-stiffened 17-7 PH steel conical shells under axial compression-
end fittings for specimens (from [9.229])

lb (~89 k:N) calibrated BLH load cell was located between the large end fitting and the
central rod. Calibrated linear variable differential transformers (LVDT's) were attached to
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

the large mounting plate, while their cores touched the small mounting plate to measure the
end shortening of the shell.
Twenty-four strain gages, bonded to the shell and on the stiffeners, measured the strains
that appeared on the surface of the shell and on the stiffeners during loading. The strain
measurements (recorded automatically on a B & F plotter) assisted in the detection of buck-
ling and checked the development of any inelastic strains prior to buckling. Six gages were
bonded to the inner surface of the shell, very close to its small end, equally spaced circum-
ferentially and in the direction of the generators of the shell, to measure longitudinal strains.
Their main purpose was to ensure uniform axisymmetric distribution of the axial loading.
The shells were loaded and reloaded until a nearly uniform distribution of these six strains
was achieved.
The remaining gages were all attached circumferentially, six opposite the longitudinal
ones, six along the average radius of the shell (paJ, and six near the large end of the shell
(R2).
In some of the shells with heavy ring stiffeners, six gages were bonded to the surface of
the stiffeners in the ¢ direction and another six to the inner surface of the shell opposite to
those on the stiffeners. The efficiency of the rings in carrying the applied load could thus
be investigated.
During the tests, the points where the load versus strain plot became noticeably nonlinear
were an indication of local onset of buckling. The state when almost all the gage plots, or
all the gages at the same height of the shell, ceased to behave linearly was considered to
represent buckling.
In these tests the shells were approximately clamped at their ends. The clamping procedure
was as follows. The shell was clamped at its small end (R 1) in the small mounting plate and
clamping plate (see Figure 13.58) with the aid of a special aligning jig. The large mounting
plate was then attached to the loading frame used for torsion tests (see [13.120] and
[13.121]), and there the shell was finally fixed to the large plate with the large clamping
ring. Uniform fastening of this ring was achieved by tightening of the bolts with a torque-
meter. There was almost no rotation of the generators of the shell at the end fittings, as well
as zero displacement in the w direction and average zero displacements in the u and v
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1016 Stiffened Shells

direction of the shell. Thus experimental boundary conditions were approximately full clamp-
ing (C4).
Before each test. the out-of-roundness of the shell was measured, after it had been clamped
at it$ large end (R!), at three heights of the shell.
In this test program 29 integrally ring-stiffened steel conical shells were tested. Though
only 17 shells were made and all the specimens were fabricated with the same taper ratio
o/J = 0.679, some of them. which buckled close to one of their ends. could be shonened and
retested. thus significantly increasing the total number of shells examined. Agreement be·
tween the experimental results and linear smeared stiffener theory was found to be affected
primarily by the area ratio (A l /ar/1), and for (A 2/a,/l) > 0.2 to be fairly good, as was already
pointed out in Subsection 13.1.6. Beyond an area ratio of 0.5 the improvement in ( P,., / P,.)
was insignificant. similar to the trend observed in Figure 13.8 for ring-stiffened cylindrical
shell s.
One may note in Figure 13.5S the guide rod and ma ting sleeve designed to ensure con-
centricity of top and bottom end plates. In later evaluations of the test ri g and results. as
well as of the corresponding setup for stiffened cylindrical shells, the possibi lity of friction
load sharing by the guide rod and sleeve was con~idercd. Since such friction load sharing
would vary fonuitously for each test and could overstate the measured axial load, it was
decided in later tests on cylindrical shells to discard the guide rod and sleeve. as was men-
tioned in Subsection 13.3. 1. In retrospect. it might have been advisable to devote some effon
to measuring the suspected load sharing (for example hy strain gages on the gu ide rod, or
insertion of an additiona l sensiti ve load cell). Unfonunatcly. this was not done at the time.
As pointed out in Subsection 9.6.4. an extension of the Southwell plot was applied to the
stiffened conical shells, both the intercept and the slope method. In the ca.~e of hydrostatic
pressure (see [9.278]) very good results were obtai ned for all shells. whereas in the case of
axial compression (sec [9.149)) the Southwell method was successful for most speci mens.
but could not be applied for some because of their almost linear strain plots.
Two typical 17-7 PH steel ring-stiffened conical shells are shown in Figure 13.59. Typical
postbuckling patterns in axial compression for a long and a shon shell are shown in Figure
13.60. Only 3 specimens out of the 29 exhibited an axisymmetric, ring-shaped postbuckling
pattern (see Figure 13.61), as predicted by [13. 131 and 19.229]. These three shells. however,
were the most heavily stiiTcncd ones of the whole test series. with ring cross-sections that
were much too large to be structurally efficient.
Before we close th is subsection. it may be of interest tO reiterate that the Technion ex-
perimental program on ring-stiffened conical shells included torsion tests and combined
loading of torsion and axial compression (sec [ 13. 1201 and [ 13.122]). The torsional buckling

Figure 13.59 Technion tests on ring-stiffened 17-7 PH .tccl conical shells under axial compression-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

typical specimen,, t/J = 0.679 (from [9.149])


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Stiffened Conical and Spherical Shells 1017

(a)

(b)

figure 13.60 Technion tests on ring-stiffened 17-7 PH >tccl conical shells under axial comprcssion-
t)•pical posthnckling patterns (from (9.149)): (a) long shell M3-7, of!= 0.679. (b) shon
shell M3-6C. of1 = 0.374

tests involved also 17-7PH hydrospun specimens with non -unifom1 stiffener spacing (see
Figure 13.62). The torsional buckling waves that were close to the small end in the long
shell with unifonnly spaced ri ngs (Figure 13.62a). due to the higher stresses there. moved
to the center of the conical shell when the spacing of the rings was made narrower ncar the
small end (Figure 13.62b).

13.5.3 More Recent Tests on Stiffened Conical Shells


Only a few buckling tests of stiffened conical shells have been carried out in the last two
decades. One interesting example from the mid-seventies is the cast epoxy ring-stiffened
conical shells. representative of the then current spacecraft reentry structures (Viking aero-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

shell ). tested at NASA Langley Research Center (sec Figure 13.30 and [ 13.81 ]). mentioned
in Subsection 13.2.4. Five of these epoxy shells were cast and machined and then tested
under external pressure after being mounted on a fiberglass-epoxy-covered solid wooden test
fixture. The top specimen support ring sat on a rubber 0-ring, the base ring was scaled with
a flexible membrane seal and then the shel l was pressurized by reducing the pressure in the
cavity between the specimen and the test fixtu re (see Figure 13.63a), as is often done in
spherical cap tests. The specimens were loaded until global buckling occurred , and the tests
were repeated at least once more (feasible because of the high strength-to-stiffness ratio of
the epoxy) in order to determine their repeatability, which was found to be excellent.
Displacements normal to the shell surface were measured with LVDT's. Eight LVDT's
were positioned circumferentially at a constant meridional coordinate (as can be seen in
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1018 Stiffened Shells

figu re 13.61 Technion tests on ring-stiffened 17-7 PH steel conical shells under axial compression-
axisymmctrical post buckling pauem of shell MJ-13. 1/1 = 0.679 (from [9. 1491>

Figure 13.63b) to provide the displacements needed to define the circwnferential buckling
mode shape. After this mode shape had been detennincd. six LVDT's were placed along a
line of maximum inward displacement ampli tude to yield the meridional buck ling mode
shape. The displacements recorded by the LVDT's were accurate to within ±25 JLI11. All
di splacements. as well as strains from strain gages and the measured applied pressures, were
automatical ly recorded on magnetic tape.

(a) (b)

Figure 13.62 Technion tests on ring-stiffened 17-7PH steel conical shells under torsion- typical post-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

buckling pallcms (from [ 13.122)): (a) shell MJ-18 with uniform ring spacing. (b) shell
Copyright Wiley
MJ- 17 with non-uniform ring spacing.
Provided by IHS Markit under license with WILEY
Both were long shells with o/1 • 0.679
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Stiffened Conical and Spherical Shells 1019

(a)

(b)

Figure 13.63 NASA Langley ting·sl iffcncd ca>I-epoxy conical ;hell 1es1s -1es1 sclup (from ( 13.8 1]
and cout1esy of Dr. R. Davies): (a) cross-seclion of conical she ll and IC>I fixiUre, (b)
view of shell in 1es1 selup

Good agreement was obtained between the buckling pressures and number of circumfer-
ential waves predicted (by BOSOR 2. an early version of the widely used BOSOR 4 code.
discussed in Chapter 2. S ubsectio n 2.2.3a and [2.92]) a nd the observed inilimi on of asym-
metric buckling. With increasing pressure the basic characteristic mode shape grew to large
amplit udes (see Figure 13.30), but with the shell demo nstrating increa.~i ng postbuckling
strength till it collapsed at a pressure about 20-40 percent higher than the onset of buckling,
or bifurcation.
The tests on the cast-epoxy model~ served as a parametric study (together with the ac-
companying BOSOR 2 calculations) to guide the design of the Viki ng aeroshell . The re-
sulting prototype was finally tested fu ll scale (actual ly two ring-sti ffened alumi num alloy
ring-sti ffened shells of revolution were tested at NASA Langley-see [ 13. 123)). Two similar
large 4.5-m diameter magnesium alloy ring-stiffened conical shells were also tested at NASA
Langely in the same period (see [ 13. 124 J).
A more recent example is the buckling and vibration tests on ring-stiffened mild steel
conical shells carried out at the Depanment of Mechanical Engineering of the University of
Ponsmouth in the late eighties and early nineties [13.125] and [1 3.126]. The study focused
on the vibrations and correlatio n between vibratio n and buckling, discussed in more detail
in Chapter 15, and o n the infl uence of ex ternal pressure o n the resonant frequencies of the
vibration modes thai arc similar to the buckling modes.
The test specimens were machined from a solid block of mild steel. The internal surface
was machined first. and then. after snugly fining the shell onto a conical mandrel. its external
surface with the reinforcing rings was machined. The first series [13.125) included three
shells with R, = 19.05 mm, R2 = 50.8 mm. (p, Jt) = 55.6. a length of 2 11 mm with six or
seven evenly spaced rings, and (A 2 / lt 111) =
0.0539-0. 123, i.e. relatively thi ck-walled and
lig htly stiffened shell s. The second series [ 13.126) consisted of another three shells with
similar R,, R1 , length, and ring-sti ffeni ng but thicker skin. These later conical shells with
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

( p.Jt ) = 34.7 and (A 2 /a0t ) = 0.0337-0.0770 were therefore even more thick-walled and
Copyright Wiley
Provided lightly stiffened
by IHS Markit than
under license with the first series. Indeed. inelastic
WILEY buckling
Licensee=McDermott appeared User=G,
Inc - India/8215328006, already in one or
Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1020 Stiffened Shells

two of the shells of the first series, whereas the second test series was deliberately designed
for inelastic buckling.
Plots of the initial out-of-circularity for shells 1 and 2 at mid-height (between the third
and fourth rings) and for shell 3 at 0.56 the height from the large end were carried out
before and after the vibration tests. The vibration tests were found to cause no permanent
deformations and hence did not influence the initial imperfections of the shells. The mag-
nitude of the worst out-of-circularity was therefore the same before and after these vibration
tests, and rather small, ± 2 percent of the shell thickness t for shell 1, ± 0.4 percent for shell
2 and ± 1 percent for shell 3. The specimens could therefore be classified as very accurate,
with similar imperfection amplitudes to those of the earlier Portsmouth or Technion ring-
stiffened cylindrical shells discussed in Subsection 13.2.1 or the Technion ring-stiffened
conical shells treated in Subsection 13.5.2.
In the vibration tests the conical shells were excited by two electromagnetic shakers and

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
with the aid of a frequency response analyzer and other electronic equipment, as used in
earlier Portsmouth vibration experiments on hemi-ellipsoidal domes (see for example
[13.127]). Even lobar modes were obtained by keeping the shakers in phase, while for odd
lobar modes they were kept at 180° out-of-phase. For vibrations in air a microphone was
employed to scan the mode shapes, whereas for underwater vibrations the microphone was
replaced by a hydrophone. The underwater experiments were carried out in a relatively large
tank of 1.5 m diameter, large enough to prevent any significant influence of the walls of the
tank.
The external pressure buckling tests that followed the vibration experiments were carried
out in the test tank shown in Figure 13.M, which somewhat resembled the earlier Technion
test rig for stiffened conical shells (Figure 13.57) and Ross's test fixture for ring-stiffened
cylindrical shells under external pressure [13.104]. The large end of each conical shell was
fixed to the top of the tank with four bolts, as was the small end to its blanking plug. The
large end was therefore effectively fixed, while the small end was clamped but free to move
axially. Due to the relatively weak stiffening, the shells buckled by general instability (see
Figure 13.65).
Eight or 10 strain gages, placed circumJerentially, were bonded to the internal surface of
each shell between the first and second or second and third rings, where the largest deflection
of the buckles was expected. Plots of pressure versus strain provided indications of the
buckling behavior. For example, for shell 1 just before buckling the strains suddenly became
nonlinear, indicating onset of buckling; but since these critical strains were well below the
yield point of the material, this verified that the specimen failed elastically. For shells 2 and
3, on the other hand, the strains near buckling were near the yield point, indicating the
possibility or likelihood of inelastic buckling. The three more recent conical shells [13.126]
buckled inelastically and hence showed significant inelastic knock-down factors when com-
pared to predictions for elastic instability.
A few additional tests on ring-stiffened conical shells would have greatly enhanced the
value of the results, as was also pointed out by the authors in [13.126]. Additional tests
would also have emphasized that the relative simplicity of the experiments and their instru-
mentation appear not to have precluded good and reliable results.
In passing, one may mention the extensive submarine oriented tests on stiffened cylindrical
shells under hydrostatic pressure, which also included junctions with truncated conical shells,
carried out by the French naval establishment at Cherbourg (DCAN) in the mid-seventies
[13.128].

13.5.4 Stiffened Spherical Shells


Reports on buckling tests on stiffened spherical shells are even scarcer in the literature than
those on conical ones. A good example, though from the sixties, is the tests of Krenzke and
Kiernan at the U.S. Navy DTMB on stiffened and unstiffened machined spherical shells
Copyright Wiley
[9.312],
Provided by IHS mentioned
Markit under in Chapter 9, Subsection 9.1
license with WILEY 0.1.
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Stiffened Conical and Spherical Shells 1021

Alr w nt

·~ 1-i-- W~ter

- - - Pressure
vessel
'0' Rin9

End bf ~nkincJ
plug

(a)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(b)

Fi ~:ure 13.64 University of Portsmouth bucldiog tests on conical shells under external pressure-test
tank with shell (from [13.125). counesy of Professor C.T.F. Ross): (a) cross-section of
conical shell and prcs~ure •·esse!. (b) view of ; hell and test tank before final assembly

The stiffened shells (series SSS in the OTMB tests) consisted of 17 externall y stiffened
hemispheres. machined from solid 7075-T6 aluminum alloy. The first nine models had cir-
cumferential stiffeners only, whereas models SSS- 10 through SSS- 13 had meridional stiff.
encrs only and models SSS - 14 through SSS-17 had both circumferential and meridional
stiffeners covering a central portion of the hemispheres considered adequate to represent a
complete sphere (see Figure 13.66).
The interior contours of each model were machined with a form tool. In models SSS-1
to 9 the exterior contours were machined using a lathe with a ball-turning anachment. as
employed for the accurate larger unstiffencd hemispheres of [9.323] (except that here no
mating support mandrel was needed since the shells were relati vely thicker). ' 'The ex teri or
contours of models SSS- 10 through 17 were obtained by first turning the contour of the
outside radius of the stiffeners and the thick-walled parts of the shell. Then the material
between the stiffeners was removed individually by indexi ng the model in relation to an
electrode in an electric di scharge machine." In other words. the stiffeners were obtained by
electrochemical machining of the contoured thick-walled hemi sphere, and thus the stiffeners
Wiley ~hell fonned an integral un it.
and
Copyright
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1022 Stiffened Shells

•' lgure 13.65 University of Ponsrnouth buckling tests on st iffened conical shells under extcmal pres·
sure-postbuckling pattern of ring-stiffened couical shells (reprinted from C.T.F. Ross
and H. Hamer. '' Inelastic Buckling of Ring-Stiffened Conical Shells under External Pres-
sure:· in: Advtmus i11 Srntcwral Et~gi11eeri11!l Comp111i11g. B.H.V. Topping and M . Pa-
padrakakis. eds•. copyright 1994. with pennission from Civii-Comp Ltd. and counesy of
Professor C.T.F. Ross)

The wall thickness of each SSS specimen was measured by using a small ball -support
and a dial gage, but the ini tial out-of-roundness. or out-of-spherity. was not measured in this
test series. All the SSS models were tested in the th ick-wal led 2-in. diameter tank shown in
the test setup of Figure 13.67. No strain data or internal volume measurements were recorded
during this test series. For shells SSS- 14 through 17, however, visual inspection of the inside
contour of each model was conducted at all pressure increments prior to collapse. The
inspections indicated that excessive local deformation of the shell between stiffeners occurred
during the test~ of these four models.
Krcnzke and Kiernan compared the experimental collapse pressures of the 17 stiffened
spherical shells with the estimated collapse strength of equivalent unstiffened shells of the
same weight. They found that in no case was the collapse strength of the stiffened shell as
great as that expected for a machined unstiffened shell of the same weight. Analyzing their
results, they concluded that un less it was closely spaced. stiffening could not be effective
and that stiffening in both circumferential and meridional directions seemed most promising.
They also affirmed that considerable theoretical and experimental study was req uired before
rational designs of efficient spherical shells could be achieved.
The relatively incomplete measurements in the SSS test series can be explained by their
being only an exploratory study of the effect of different stiffening systems on the collapse
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

strength of spherical shell s. The DTMB researchers probably left the more comprehensive
measurements to the proposed continuation of tests on shells with the most promising sti ff-
ening configurations.

13.6 Experiments on Stiffened Curved Panels


13.6. 1 Curved Panel Experiments
Stiffened curved panels. mostly cylindrical panels. arc frequently employed as structural
elemenL~ in aerospace vehicles. as well as in marine and civi l engineering. Hence knowledge
of their buckling and postbuckling behavior is important. Funhennore. stiffened cylindrical
panels also serve as possible practical " test vehi cles" for experimental evaluation of the
buck ling and postbuckling behavior of complete large stiffened cyli ndrical shells, too large
for full-scale testing of the complete cylinder (as for example the 10-m diameter liquid
hydrogen propellant tank of the Saturn V launcher). The feasibility of this more economical
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Curved Panels 1023

__ }121N

I f.-· - 1,62 5 IN .lo.- -1 I -!


1- ---·-2.,50 IN.oo~- -1 I-Q020

(a) 1-- Q250 IN.DiA,

L 0.149 iN.DiA.
j, O.Q4 IN

0.0385

1'------C-~~~~~
!__ __ 1.625 IN.IJ- ~ ,.._ -~ - 1.625 :N.DIA. ---

(b) (c) (d)


Figure 13.66 U.S. Navy DTMB experiments on stiffened hemispheres-main dimensions of specimens
(from [9.312]):

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
(a) Models SSSl-4: wall thickness t = 0.0074-0.0105 in.
stiffener height d = 0.030-0.040 in.
number of stiffeners N = 3-6
(b) Models SSSS-9 t = 0.0077-0.0163 in.
d = 0.032-0.042 in.
N = 6-9
(c) Models SSSl0-13 t = 0.0046-0.0088 in.
d = 0.040-0.044 in.
N = 8-16.
(d) Models SSS14-17 t = 0.0070-0.0096 in.
d = 0.029-0.032 in.
N = 16 meridional
2 circumferential

Figure 13.67 U.S. Navy DTMB experiments on stiffened hemi-


spheres-test setup (from [9.312])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1024 Stiffened Shells

simulation of the complete stiffened cylinder by a stiffened cylindrical panel depends on


sufficient width of the panel and on appropriate boundary conditions along the straight edges.
These prerequisites create boundary condition problems for stiffened curved panel tests,
which will be discussed in detail in the next section. Here it should only be pointed out that
whereas in a complete cylindrical shell the axisymmetry assists in the definition of the
boundary and loading conditions, in a stiffened cylindrical panel both the definition of the
boundary conditions and the uniformity and centroidal position of the loading present dif-
ficulties.
As in stiffened plates and shells, the local buckling of some sub-panels in a stiffened
curved panel does not necessarily indicate the limit of its load carrying capability. If the
local sub-panels are prevented from coHapsing, additional load can be supported by the
structure. The postbuckling strength of stiffened panels can usually be most efficiently util-
ized if the structure can be allowed to buckle locally at relatively low load levels.
General instability of closely stiffened curved panels can be analyzed by methods that are
extensions of the smeared stiffener theory developed for closely stiffened shells (see for
example [13.129] and [13.130]). On the other hand, the postbuckling analyses for stiffened
curved panels are still incompletely developed, in particular for composite panels, though
some computer codes such as NASTRAN or STAGS can handle them. Hence experiments
also have to provide essential design information for stiffened curved panels under different
loadings (see for example [13.131]).
The experiments on the buckling and postbuckling behavior of hat-stiffened, composite
cylindrical panels under axial compression carried out by Agarwal at the Northrop Corpo-
ration in California in the late seventies and early eighties [13.131] are a typical example of
tests on relatively small stiffened curved panels. The specimens were 30° four-layer graphite
epoxy panels 630 mm wide and 508 mm high, with (Rit) ~ 2240 and with four cocured
graphite epoxy trapezoidal hat-section stiffeners on the concave side (see Figure 13.68).
After cure, the panel edges were trimmed. Since the skin of the panels was very thin, the

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
panels were clamped between two curved templates while their ends were potted in alumi-
num-filled epoxy. Then the potted ends were machined flat, care being taken that both ends
were parallel. The panels were then tested flat between the loading heads of a 50,000 lb (~
220 kN) universal test machine. Strain gages on both end stiffeners served to monitor the
uniformity of the load across the panel, while additional strain gages determined the stress
distribution within the panel, all gages being mounted back to back. In some cases thin fillers
had to be inserted between the loading head and the end plates for more uniform load
distribution.
The panels were designed for early initial local buckling, at about 25 percent of limit
load, and indeed reached ultimate loads in the tests of about five times their bifurcation
buckling loads. Two of the four panels were tested under repeated loading. Since curved

JJ
I
r,·
Figure 13.68
45.0~
~
!

Northrop postbuckling tests on hat-stiffened composite cylindrical panels under axial


compression-dimensions of test specimens (from [13.131])
-1
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Curved Panels 1025

panels have a tendency to slip between the test heads during repeated (fatigue) axial loading,
this slipping was prevented by providing guide blocks along the potted ends. These guide
blocks were fixed to the loading head. As usual in composite flat and curved panels, these
two panels sustained two blocks of 100,000 cycles of repeated buckling with practically no
effect on the residual strength.

13.6.2 Early Lockheed-UC Berkeley Tests


An example of larger stiffened panel tests is those carried out by the Lockheed Missiles and
Space Company in California in the mid-sixties as part of a comprehensive experimental
program related to the Saturn V launcher [13.132] and [13.133]. This NASA-sponsored
program included tests on four full-scale integrally stiffened aluminum alloy panels, phase
I; and on 29 one-quarter scale (of width 710, 1420 and 2130 mm) integrally stiffened panels
as well as 6 one-quarter scale (2515 mm diameter) complete stitiened cylindrical shells, all
made of aluminum alloy, phases II and III. The tests were conducted at the University of
California, Berkeley, Structural Engineering Laboratory, with Lockheed-designed and fabri-
cated test fixtures. Extreme care was taken in duplicating the test conditions of the full-scale
tests for the one-fourth scale experiments, which meant close tolerances and more precise
alignment techniques for the one-quarter scale panel test fixtures.
Two universal testing machines were employed for the one-fourth scale tests: (1) A
400,000 lb (=1780 kN) capacity Baldwin Universal Test Machine; and (2) a 4,000,000 lb
(= 17,800 kN) capacity Southwark Emery Universal Test Machine for the complete stiffened
cylindrical shells and the widest (2130-mm wide) stiffened panels.
In the smaller Baldwin machine, load was applied through the hydraulically actuated base
table and reacted by the top cross-head, incorporating a swivel, spherical joint and a 10-in.
(=254-mm) diameter upper load platen. Two sets of two wide-flange load beams, of 30 in.
(=760 mm) and 70 in. (=1780 mm) length and 6 in. (=150 mm) and 24 in. (=600 mm)
width, respectively, were employed. Each of the top ones was in its turn rigidly fixed to the
upper load platen, being still free to rotate through the machine's spherical joint. Both sets
of beams were ground to a top-to-bottom flange parallelity of 0.008 in. (=0.2 mm) and
surface flatness of 0.004 in. (=0.1 mm). The wider bottom load beam was also embedded
in a thin layer of Hydrostone to minimize the effect of minor flatness deviations. The vertical
unloaded edges of all panels were sandwiched between pairs of steel angles bolted snugly
together (see Figure 13.69) to prevent lateral movement and bending. The angles were bolted
only to the upper and lower loading beams, permitting axial movement of the edges of the
panels. Interface friction and the resulting load transfer were minimized by strips of Teflon
tape.
The widest stiffened panels, of 2130 mm arc length, were tested in the large Southwark
Emery machine. The compressive load was applied through its upper cross-head hydraulic
ram with the aid of a V-whipple tree loading beam. The required 0.010 in. (=0.25 mm)
parallelity caused some problems, which eventually resulted in a thick-walled box configu-
ration of the loading beam. The vertical edges of the wide cylindrical panels were restrained
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

by pairs of steel angles with Teflon interfaces, as before (see Figure 13.69).
Though the instrumentation, test procedure and analysis of experimental results (presented
in [13.133]) are not discussed here, some details will be pointed out.

a. Deflectometers
Radial displacements during loading were measured with LMSC electrovisual displacement
transducers. These simple transducers were essentially standard 5-in. throw dial gages that
actuated clip gage springs. Strain gages were bonded to the springs to provide the electrical
output, which gave a resolution of 0.06 mm with an accuracy of ± 5 percent. A simple but
effective Lockheed in-house design, these deflectometers permitted visual and recorded mon-
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1026 Stiffened Shells

."'-

. '-._ -\SPACER

TEFLON TAPE

ANGLES

PANEL

BOTTOM

Figure 13.69 Lockheed-UC Berkeley tests on large stiffened panel tests-side support installation
(from [13.133])

itoring of deflections and rate of deflection during incremental loading, as usually provided
by commercial differential transformer transducers. The long throw of the deflectometer, 5
in. (~125 mm), however, minimized the possibility of damage experienced by other trans-
ducers.

b. Load Distribution
Mismatch between the test fixture bearing surfaces and the test specimen load ends presents
the major problem in obtaining uniform load distribution. In the full-scale tests [13.132], the
mismatch was adjusted through the use of brass shims and soft aluminum bearing pads.
These pads yielded during panel loading, compensating the small variations not adjusted by
the shims. Often, however, the panel tests did not develop sufficient bearing stresses to yield
the aluminum pads. Thus, a new softer load pad material was sought, and after many tests
a loading pad of an inner layer of polypropylene and an outer layer of Teflon was selected
and gave good results for flat panels.
The one-quarter scale stiffened cylindrical panels were found to be extremely sensitive to
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

the slightest mismatch. Although the bearing stresses were high enough to justify the use of
soft aluminum pads, they did not provide adequate redistribution of load. Teflon pads, or
Teflon plus soft aluminum, were therefore substituted, with excellent results, although the
bearing stresses were often high enough to cut through the Teflon into the soft aluminum.
Hence it appeared that better compensation for these irregularities was obtained when the
pad material was in the plastic state. The local depression of the soft aluminum seemed to
trap the Teflon between the bearing surface and the loading edge. It was concluded that
Teflon bearing pads were successful for mismatches up to 0.01 mm. But the Teflon pad had
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Curved Panels 1027

one disadvantage: its load versus head travel curves distorted due to creep and failure of the
Teflon. Therefore, care was advised in its use.

c. Scale Factor
The comparisons between the one-quarter and full-scale tests indicated that there was no
special scale factor. Hence geometrically similar scale models are admissible, as was shown
also by the dimensional analysis of Chapter 5.

d. Retest Technique
Twenty-six one-quarter scale panels and five one-quarter scale cylinders were successfully
tested without causing permanent damage to the specimen. It was demonstrated that elastic
buckling of the bar-stiffened configurations could be controlled so as to prevent permanent
damage to the test specimens, provided the initial buckling stress was low in comparison
with the yield stress of the material. All specimens were, however, recontoured before reuse
in a subsequent test. Many of the panels were retested another five times, indicating the
feasibility of repeated buckling testing if certain prerequisites are satisfied and special care
is taken. These test results reinforce similar conclusions reached at the Technion in the
combined loading testing of aluminum alloy stringer-stiffened cylindrical shells to obtain
interaction curves, discussed in Subsection 13.4.1 (or see [9.272]-[9.274] and [13.107]).

e. Experimental Method
The good agreement between the one-quarter scale cylinder buckling loads and the equivalent
loads determined from the one-quarter scale and full-scale panel test results, confirmed the
method of using stiffened cylindrical panels as test vehicles for evaluation of the buckling
and postbuckling behavior of complete stitfened cylindrical shells.

13.6.3 Lockheed Palo Alto Stiffened Cylindrical Panels


A more recent example of laboratory scale tests on stiffened curved panels is the experiments
on optimized stiffened graphite-epoxy cylindrical panels carried out by Bushnell and his co-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

workers at the Lockheed Palo Alto Laboratory in the mid-eighties [13.134].


A series of tests on graphite-epoxy curved, stiffened panels was initiated by the Lockheed
team, primarily to validate the PANDA 2 computer program, for minimum-weight designs
of stiffened composite panels subjected to several load combinations, developed by Bushnell
[13.135]-[13.137]. For the tests a new testing machine for large axially stiffened cylindrical
panels subject to axial compression was built. Its unique features warrant a detailed discus-
sion, which is now presented, following [13.134] and [13.138].
Tests of rather small cylindrical shells, plates and columns under axial compression are usually
performed in what is called a 'standard test machine': a hydraulic press with one fixed platen and
one platen that translates. Both platens are so massive that they remain, for all practical purposes,
undeformed during the test. If the platen faces are parallel, and if there is virtually no rotation of
the movable platen during the compression test, the test specimen will be subjected to uniform end
shortening. If the specimen is installed vertically with great accuracy, if it is not curved in the axial
direction, and if its ends are trimmed with great accuracy, the axial strain in a large region of it
away from its loaded edges will be uniform compression prior to local buckling of the panel skin
between stiffeners. This ideal prebuckling situation is sought in the tests on the rather large, axially
stiffened, cylindrical panels to be discussed. In the load range above local buckling of the skin
between stringers, the axial strains in all the stringers should be equal, and axial bending of the
stringers should be small at loads below that corresponding to general instability.
[In practice, however,] the ideal situation is seldom encountered because of manufacturing tol-
erances during fabrication of the panels and test machine [sometimes called a test frame]. It is
difficult to test large cylindrical panels or shells in a standard test machine not only because the
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1028 Stiffened Shells

panels might not fit in the space provided, but also because it is extremely difficult to trim the
ends of the panel accurately enough to ensure reasonably uniform axial strain in the prebuckling
phase of the loading. One therefore needs a test frame with one fixed platen and one movable
platen, in which the movable platen can rotate, at least in the initial phase of the loading, in order
to position itself so that the axial strain in the prebuckling phase is as uniform as possible through-
out the panel (skin and stringers), and in the early post-local-buckling phase the axial strain is the
same in all stringers. This optimum positioning of the movable platen must be accomplished by
means of a control mechanism based on axial strains measured at appropriate locations in the test
specimen.

The axially stiffened, composite panels tested were optimized for minimum weight with
PANDA 2, yielding
a configuration in which local buckling between stringers is predicted to occur at only about 11
percent of the ultimate failure load. At loads above the local buckling load, the axial strain will no
longer be uniform in all parts of the panel because axial load will be shed from the panel skin to
the stringers as the amplitude of the local buckles deepens with increasing end shortening. The
skin and perhaps the web(s) and fiange(s) of the stringers will experience significant bending strains.
In order to ensure uniform end shortening over the entire panel cross section in the post-local-
buckling regime, the platens in the test frame must be massive enough not to deform significantly.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
In other words, the test frame should "permit application of progressively increasing,
uniform end displacements to the end planes of a rather wide panel being loaded in com-
pression, while accommodating significant non-parallel end planes that inevitably exist on
any large test specimen. (It should be emphasized, however, that the ends of the trimmed
panel must be planar. There should not be significant waviness of these end surfaces!)"
A conceptual sketch of a panel test frame, with a curved hat-stiffened panel in it, that
fulfills these requirements is presented in Figure 13.70. The frame has the following principal
components:
"1. There are two thick platens (henceforth called plates) one at each end of the panel.
The planes of these plates are nominally perpendicular to the axis of compression
loading, which is vertical.
2. There are three tension links, each nominally parallel to the axis of compression loading
and arranged so that their connection points to the upper and lower loading plates form
equilateral triangles. The centroids of the two triangles coincide with the load centroid
of the test panel.
3. Each of the three tension links consists of a hydraulic actuator, able to create tension,
in series with a load cell able to measure the tension created. Each tension link is
connected to the two loading plates by universal joints [called U-joints in Figure 13.70]
able to accommodate small angle changes where the centerlines of the tension links
connect to the loading faces of each of the two end plates.
4. The six universal joints mentioned in Item 3 cause the test frame to be literally unstable;
it is a mechanism. In order to correct this condition, a very stiff post parallel to the
axis of load is rigidly attached to the lower of the two thick end plates. This post
extends to where it is connected by a universal joint and a pair of linear ball bushings
to the upper end plate. The linear bushings allow the upper plate to travel freely a
short distance up or down, that is, in a direction parallel to the compression loading
axis. The universal joint, which is a large spherical bearing, allows the upper plate to
pivot about either of the two axes lying in the plane of the upper plate. The linear
bushings and universal joint restrain the upper plate from translational motions in its
plane.
5. The upper plate is still free to rotate relative to the lower plate about a vertical axis
through the center of the universal joint. To eliminate this last undesired degree of
freedom, a cross beam [not drawn in Figure 13.70, but visible in the photograph of
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Curved Panels 1029

Cl::IHROIOAL
AXIS

UPPER
END PLATE---.....

L1 NEAR
BEARtNC
(TYP]

U-JOlNT

POST

Figure 13.70 Lockheed Palo Alto tests on graphite-epoxy curved stiffened panels loaded in axial
compression-conceptual sketch of the panel test frame (from [13.134]). The unique
feature of this test machine is the freedom of the upper end plate to rotate about two
axes in the plane of the plate, thereby accommodating panels with non-parallel end
planes

the actual test frame in Figure 13.71] is rigidly connected to the top of the post. Two
linkage rods connect the ends of the cross beam to the side edges of the upper plate."
The actual test frame (see Figures 13.71 and 13.72) differs somewhat in the details:
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

"1. There are six tension links instead of three. For the system to be statically determinate
they are arranged in three pairs. the centroid of each pair falls at the vertices, D 1, D2,
D3, of the equilateral triangle discussed previously [see Figure 13.71]. The two actu-
ators within any given pair of tension links are hydraulically in parallel, so that the
forces in each tension link of the pair are always equal. This means that the resultant
force produced by a tension-link pair acts at the midpoint between the two links, that
is. at the vertices, D1, D2, D3 of the equilateral triangle. There are three reasons for
using pairs of tension links: (a) Smaller actuators are required since the load is shared
among twice as many; (b) Six smaller loads distributed [as shown in Figure 13.71]
cause less distortion of the two end plates than three loads" [shown in Figure 13.70];
"(c) It is desirable to be able to measure the relative axial movement of the end plates
at the vertices, D1, D2, D3." [A deftectometer can conveniently be placed at the vertex
between the two tension links.]
2. [The hydraulic actuators are not incorporated here into the tension links as in Figure
13.70.] Instead they are "through-hole" actuators located beneath the lower end plate
[J in Figure 13.71]. The actuators are not shown in [Figure 13.71], but some of them
are visible in [Figure 13.72]. The actual tension links (T), shown in [Figure 13.71]
pass through six holes in the lower end plate J and on through the centers of the
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1030 Stiffened Shells

31 1n.

PLAN VIEW
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

1..---" ~'
DETAIl

813
Figure 13.71 Lockheed Palo Alto tests on graphite-epoxy curved stiffened panels loaded in axial
compression-actual test frame (from [13.134]). There are six tension rods rather than
three and the resultant force of each pair of tension rods acts at a vertex Dl, D2 or D3
of the equilateral triangle, the centroid of which should coincide with the centroidal axis
of the axial stiffness of the test specimen

actuators. [In each tension link the actuator pushes against a 'mushroom head' on the
end of the link (which is actually the head of a bolt that screws into the end) away
from the end plate in order to create the tension force.] The tension links pass through
holes in the upper plate H. Their upper ends are fitted with large nuts, especially visible
in [Figure 13.72]. The nuts are used to adjust the lengths to the links prior to application
of the load.
3. [The load cells are not independent units in the tension links as suggested in the
conceptual sketch (Figure 13.70), but each link has strain gages attached to it to form
a load cell, as can be seen in Figure 13.73]. "Back-to-back axial strain gages are affixed
to diametrically opposed, deep, milled slots in each tension link. The milled slots
reduce the cross section of the tension link and provide fiat surfaces near the neutral
axis of the link, upon which to mount the back-to-back axial gages. In this way the
axial strain is increased locally, thereby increasing the accuracy of the output. Bending
coupling is virtually eliminated because both gages are near the centerline of the ten-
sion link and output from the back-to-back gages is averaged."
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Curved Panels 1031

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
(a) (b)
Fl j:ure 13.72 Lockheed Palo Aho test' on graphite-cpox) curved Miffcncd panels-te\1 frame (from
113.13-1). councsy of Dr. D. Bu,hncll): (a) (root \iew. (b) back"""· The large round
plates on the upper ~urface of the upper end plate are nuts attached to the tcn,ion links.
In the center of the figure one can 'ec two tcll\ion lmks with a deOcctomcter between
them

4. !The end plates II and J have their corner~ cut off at an angle (sec Figure 13.71).
which reduces their weig ht and provides ea~icr acce's to the test specimen. I
..5. The detail at the bottom of [Figure 13.71) provides a better view of the linear bushing
and uni versal joint at the top of the post C. The universal joint U is fixed to the press-
fit pin P. Therefore. it travel s up and down with the upper plate H. The universal joint
location on the pin i'> ~ct so that its center fall~ on the plane of the loading face of the
upper plate H. This feature keeps the lateral movement of points on the loadong face
to very low levels a' rotation of the plate occurs. The linear ball bushings arc press fit
into the block A. which also holds the outer race of the spherical bearing. The linear
bushings in Block 1\ arc fr~e to move up o r down on two shafts that arc press fit into

Fl~:u•-c 13.73 Lockheed l'nlo Aho tc>l> on graphite-epoxy cu1vccl stiffened panels- c loseup back view
of test frame. showing the univer.al joinl at lhe top of lhe vertical (J0\1 (C in Figure
13.7 1), load cell' (axial strain gages in slot\) and a deneciOmctcr hctwcen two tension
links (fromJI3. 134J, courtesy of Dr. D. Bu,hnell)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1032 Stiffened Shells

Block B, which is rigidly bolted to the top of the post C. Some of these details are
visible in [Figures 13.72b and 13.73].
6. The torsion-restraining cross beam mentioned previously is fixed to block A and there-
fore moves up or down with the upper end plate P. Torsional load is therefore carried
by lateral forces on the two linear bushings and the pins on which they travel. The
cross beam and attachments can be seen in [Figures 13.72b and 13.73].
7. There are three defiectometers [omitted in Figure 13.71], each consisting of an ex-
tendible rod spanning the "clear distance" between the loading faces of the end plates
H and J. The axes of these rods are nominally parallel to the axis of compressive load.
They intersect the end plates at the vertices, D 1, D2, D3, of the equilateral triangles
shown in [Figures 13.70 and 13.71]. The ends of each of the three extendible rods are
connected to the loading faces of the end plates by means of small universal joints."
[No load is transmitted by the telescoping rods and LVDT's are used to measure the
relative motion of the two ends of the rods.] Dial gages are also installed on the
assembly [see Figure 13.73] so tha:t direct deflection readout is possible to ±0.001
inches. Electrical readout from the LVDT enables reading the resetting the end plate
locations to a tolerance of ±0.000005 inches (~0.00013 mm).
8. [The two thick end plates Hand J are identical, made of 7075-T6 aluminum, and are
six inches thick. A simplified but very conservative analysis shows that one of these
end plates deflects less than 6 percent of the elastic shortening of a typical test panel,
under a load of say 50,000 lb. Hence if the test specimen is assumed to be unbuckled
and perfectly aligned, the maximum nonuniformity of load caused by end plate defor-
mation cannot exceed 12 percent of the elastic shortening, though in practice it is
probably less than half that figure.]
The description of this test frame was rather detailed, as it showed some typical design
considerations for a modern test rig. For further details the reader is referred to [13.138].
Before any large curved stiffened panels were made, several tests were conducted on small
fiat panels in order to acquire efficient fabrication techniques as well as develop appropriate
test procedures.
The fabrication and test procedure of the three large stitiened curved graphite-epoxy panels
(whose dimensions are shown in Figure 13.74) is detailed in [13.134]. Only a few significant
test items will be outlined here.

a. Potting of Panel Ends


After trimming, the panels were mounted 011 0.125-in. (~3.2-mm) thick 6061-T6 aluminum end
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

plates, which served three purposes: (I) preventing of gouging of test frame loading surfaces, (2)
yielding locally under local high loads and thereby somewhat correcting the nonuniformity of
prebuckling strains, and (3) acting as casting base for potting the ends. As indicated in the previous
subsection, it is important not to use too soft a material for the end plates, lest the panel edge cut
right through the end plates or rotation occurs.
The potting of the large panel ends was a bit tricky, mainly due to alignment problems and
because the panels were quite flexible and easily distorted prior to plotting. A reference frame was
therefore used for aligning the panels. The Lockheed team concluded, however, that "in future
tests it would be a good idea to manufacture a rigid frame or template for holding the entire panel
in a true circular cylindrical shape during the entire potting process. This special fixture must
penni! turning the panel over without releasing it from its holding restraints." A fixture of this type
would reduce the spring-back experienced with use of their alignment frame, from which the panel
had to be removed and reinstated for potting of the second end.

b. Installation of Panel in Test Frame


The centroidal axis of the axial stiffness of the curved panel must be positioned so that it coincides
as accurately as possible with the centroidal loading axis of the test frame .... Proper positioning
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Stiffened Curved Panels 1033

(~)

.11. 0 in.

-.--
5. 0 in.

1
15.5 in.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
j_
~0.6 in. POTTING TYP. \ I I
\_EMBEDMENT PLATE TYP.

Figure 13.74 Lockheed Palo Alto tests on graphite-epoxy curved stiffened panels-dimension of large
panels and locations of strain gages. Gages R were used to monitor overall bending of
the panel and uniformity of axial compression across its width. The axial locations of
gages P, Q, T and Z were determined by first loading the panel, without these gages,
well into the post-local-buckling regime to observe the positions of the buckles and the
resulting maximum strain concentrations (from [13.134], courtesy of Dr. D. Bushnell)

of the panel can be achieved by a trial-and-error process; low loads are applied, differences in the
loads at the vertices of the equilateral triangle are noted; the loads are removed and the panel is
repositioned; low loads are again applied, etc. During this stage the loads should not exceed the
local buckling load of the panel skin.

c. Application of Loads in Test Frame


The load at each vertex of the equilateral triangle shown in [Figure 13.701 can be controlled
independently. In fact, the deflection at each vertex, as measured by the deflectometer there [Figure
13.73] can be set and held by the servo-system that operates the pairs of actuators at that vertex.
The feedback signal for the hydraulic system at a vertex is the deflection signal generated by the
deflectometer associated with the actuator pair at the vertex. In a system with displacement feed-
back, the function of the servo is to maintain the feedback signal at a constant level, that is, to
maintain a commanded dellection until a new command signal is issued to alter this deflection.
This means that if a displacement change is commanded at one vertex, loads will change at all
vertices as required to create that displacement change. The predominant load change will be at
the vertex where the displacement change command is issued. The resulting new load values will
be the loads required to hold the original dellections at two of the vertices and to hold the new
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1034 Stiffened Shells

deflection prescribed at the third vertex. If commands for equal deflection changes are issued
simultaneously to all actuators, new loads will be applied so as to create equal deflections at the
three vertices .
. . . After the panel is carefully centered in the test frame, increments of load are applied by
issuing commands to each of the three servo systems to ramp the deflection simultaneously. Elec-
tronic circuitry was provided so that a comr1on ramp command signal to all three systems causes
uniform end deflection.
Ideally, the common ramp command should create equal loads on all of the load cells. Actually
the increments are not equal because of misalignment of the panel, imperfections in it, or effects
of local buckling.

With aid of the 'survey' strain gages (labeled R in Figure 13.74), the operator could adjust
the actuator system to compensate for any significant load differences.
All loads and gage readings inevitably change a little each time a command is issued to a single
servo system, and these changes are not always in the desired or expected direction. Therefore,
single-vertex corrections usually have to be applied by a series of successive approximations, each
of which is smaller than its predecessor. If the corrections do not converge, the panel is probably
not centered with enough accuracy in the test frame. If the corrections do converge the next
displacement increment can be applied. Experience has shown that once the panel is properly in
the test frame most of the corrections are needed only at fairly low loads .
. . . After local buckling occurs a fairly extensive set of corrections is usually required. The
magnitude and need for further corrections once more diminishes as the loading continues into the
post-local-buckling regime.

The instrumentation and comparisons between test and theory are presented in detail in
[13.134], as well as the conclusion arrived at. One of the recommendations for similar future
tests warrant reiteration here:
"Many more strain gages should be attached to the skin of each panel tested. This should
be done in order to capture the maximum strains in the postbuckling regime. Axial strain
gages should also be attached to the webs of the hats in order to determine if there is any
premature crippling there due to delaminations between the cloth layers of the webs," since
early delamination in the webs seems possible. The method of location of strain gages
mentioned in Figure 13.74 seems reasonable.
However, gages must be affixed at more axial locations in order to compensate for the propensity
of the buckle pattern easily to "slide" in the axial direction: Upon reloading the panel into the
postbuckling regime after affixing the gages. this pattern may have a different axial location from
that observed upon first loading. Perhaps it would be a good idea for the panel to be loaded into
the postbuckling regime, unloaded, removed from the test frame, and reloaded into the post buckling
regime several times in order to provide a "shakedown" for greater stability of the buckling pattern
and hence maximization of the useful information obtained from a limited number of strain gages.

13.7 Special Stiffened Shells


13.7.1 Corrugated Shells Subject to Axial Compression and Bending
Corrugations are essentially a continuous form of closely spaced stiffeners in fiat or curved
sheets. Corrugated skin was widely used in the early metal airplanes, the best-known example
being the German Junkers transport aircraft of the twenties and thirties. In the sixties, sev-
enties and eighties, corrugations made their comeback, mainly as ring-stiffened corrugated
cylindrical shells for intertank and interstage structures of large launch vehicles and for
primary structures of satellites, or as circumferentially corrugated, axially reinforced agri-
cultural silos (see for example [4.48], [9.249], [13.139] and [13.140]).
This motivated extensive theoretical and experimental studies on the stability of ring-
stiffened corrugated cylindrical shells at NASA Langley, NASA Marshall and Lockheed Palo
Alto in the sixties and seventies (see [9.249], [9.250], [13.99] and [13.139]).
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Special Stiffened Shells 1035

The experimental studies at NASA Marshall Center, Huntsville, for example, were part of
the Saturn V launcher development program and included 11 ring-stiffened corrugated cyl-
inders subject to axial compression, one subject to bending and one under combined bending
and axial compression (see [13.99]). The shells were rather large, of 1.25 m, 2.50 m and 10
m diameter, 0.84-6.83 m length, (R/t) = 1230-1360, and all made of 7076-T6 aluminum
alloy. Of the axially loaded cylinders, five failed in general instability, whereas the remaining
ones failed earlier either by panel instability or local crippling. For the shells that failed in
general instability, the agreement between the experimental results and a linear theory that
considers the eccentricity of the rings was very close ( ± 1-12 percent). For the two shells
subject to bending, the agreement was not as good, but became similar after a customary
bending factor of 1.2 had been applied.
During the tests, an interesting phenomenon was exhibited by the strain gages located on
the corrugated skin. Two gages were located opposite each other at various points on the
skin (see Figure 13.75a) to indicate longitudinal bending of the corrugation. As the axial
load increased, the strain versus load plot of the two gages in each couple remained linear
and almost coincident until just below the failure load (see Figure 13.75b). Then at various
locations on the shell the strains diverged nonlinearly as shown, indicating appreciable bend-
ing and that failure was imminent. General instability failures of the shells finally occurred
very rapidly. Dickson and Brolliar of the NASA Marshall Center proposed using this fore-
boding behavior for nondestructive determination of the general instability load of ring-
stiffened corrugated cylinders. Actually, employment of the Southwell method (as originally
suggested by Galletly and Reynolds, see Subsections 9.6.1 and 9.6.4 and [9.214]) would
facilitate this nondestructive prediction. However, the disadvantage of the proposed method
for axial compression is that usually, to indicate imminent failure, the load increments have
to be raised too close to the failure load to guarantee nondestructiveness.
One should note that ring-stiffening in longitudinally corrugated cylindrical shells will
always be eccentric with respect to the shell mid-surface (center line of corrugation), in

gage no.2
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(a) ,
g

gage noJ

load
(b)

Figure 13.75 NASA Marshall Center, Huntsville, tests on large corrugated cylindrical shells subject to
axial compression (from [13.99]): (a) typical strain gage locations, (b) strain versus load
for typical test shell
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1036 Stiffened Shells

aerospace applications usually on the inside (due to aerodynamic considerations). Ring ec-
centricity significantly influences the buckling load of the shell, and corrugated cylinders
with outside rings can have about twice the buckling load of similar shells with inside rings.
A special experimental investigation of this eccentricity effect was carried out at NASA
Langley by J.K. Anderson [9.250]. Two nearly identical corrugated shells, which differed
only in ring location, were tested in bending. The shells with corrugated walls (see Figures
13.76 and 13.77) had a diameter of 1.95 m, a length of 2.03 m, (Rit) ~ 1870 and fairly
shallow hat rings of 11.1 mm height (similar to the 11.7 mm height of the corrugations),
spaced 170 mm apart (resulting in an effective ring area ratio of about 0.24 ). Each cylinder
was fabricated from seven corrugated sheets of 707 5-T6 aluminum alloy, fastened together
by double rows of spot welds. The rings were attached to the corrugated wall by rivets. The
corrugation sheet thickness was taken as the average of 35 p,m measurements around the

-IIStiffe'l rtngs
at 6.67(16.

I__ I: ~corrugot ions-

attoc"ment cing~
II
u
(a)
'------'------ ~----____L_Ju

r
t~ 6.67------!
I I 6 94 I
0 I

ri[ij}J A

~----( 2~ :~~ ,-----1

I %f t :Q 15 2 ;= :;;=:T;;

- 6 0

- ( I[.

(c) ~------Test section----

Figure 13.76 NASA Langley Center bending tests on large corrugated cylindrical shells with eccentric
ring stiffeners-construction details of two test cylinders (from [9.250]. courtesy of
NASA Langley Research Center): (a) side view of cylinder 2, (b) end ring and stiffening
ring location for cylinder 1 (outside stiffening rings), (c) end ring, steel attachment ring
and stiffening ring location and details for cylinder 2 (inside stiffening rings). Dimensions
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

are in inches, with em in brackets


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Special Stiffened Shells 1037

Figure 13.77 NASA Langley Center bending tests on larg.: ring-stiffened con-ugmed cylindrical
shells-closeup of a test cylinder, with inside ~ti iTcning rings. in position in bending
test rig (from [9.249], councsy of NASA Langley Research Center)

circumference at the middle of the ; hell (each cylinder being cut in half after the test for
this purpose). The stiffening rings location is shown in Figure 13.76 (cylinder I with outside
ri ngs (b) and cylinder 2 with inside ones (c)). The test section of the test shells was bounded
by two large, open hat-section rings (see Figure 13.76c). The corrugated wall outside the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

test section was doubled in thickness to restrict fai lure to the test section . The doubling was
achieved by bondi ng and riveting another corrugated sheet to the wall. The steel attachment
rings were also riveted to the doubled wall.
The two cylinders were tested in the large NASA Langley bending test rig (see Figure
9.75b). Both test shells buckled at approximately 70 percent of the predicted loads, calculated
wi th a small-defl ecti on theory for bendi ng, which takes the eccentricity of the rings into
account. T he test cylinder with outside stiffening rings huckled at a load 2.3 times the
buckling load of the corresponding shell with inside stiffeni ng rings. The same ratio (2.3)
of buckling load for the cylinder with outside stiffening to that for the shell with inside
stiffening was also predicted by theory. Hence it appears that the eccentricity effect of the
stiffeners is adequately predicted by small deflection theory. This reinforces the relevant
conclusions summarized in Subsection 13. 1.4.
An additional series of bending tests on five large, 2-m diameter. corrugated 7075-T6
aluminum alloy cylindrical shells, similar to the one wi th inside rings of [9.250] (see Figure
13.76). was carried out at NASA Langley in the mid-sixt ies in order to study the effect of
ring spacing on their general instabi lity [9.249]. The cyli nders indeed failed by general
instability, except one that buckled into a panel -instability mode prior to buckling in the
global buckling mode. The cylindrical shell s failed at about 75 percent of the general insta-
bility load for bending. computed by small-deflection theory.
Two supplementary tests were conducted on some of the test cyli nders to try to isolate
the reasons for the discrepancy between prediction and experiment. In the first test a small
torsional load was npplied to one of the test shells. prior to loading it in bending. in order
to detemune the shear stiffness of the corrugated wall, wh ich strongly depends on the con-
ditions at the ends of the corrugations. If the corrugations are not restrained against rolling.
their shear stiffness may be significantly lower. The optically measured twist yielded a shear
stiffness 88 percent of the elementary calculated one. However. check calculations indicated
that if the measured shear stiffness were used instead of the elementary one. the predicted
buckli ng load would only be lowered by 2 percent.
The purpose of the second supplementary test was an assessment of the effect of the
attachment between the stiffening rings on the buckling load. One of the cylinders was
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1038 Stiffened Shells

therefore retested after trapezoidal attachment clips had been added between the rings and
the corrugated wall to stiffen these joints. For this supplementary retest, the cylinder was
rotated 180° from its original test position, so that the former tension side was now loaded
in compression. The design of the clips was such that they were expected to increase some-
what the bending and axial stiffness of the rings as well as significantly restrict possible
deformations between the outer corrugation crests and the reinforcing rings. The test cylinder,
however, failed at essentially the same load as before, indicating that the clips were rather
ineffectual and that the type of deformations restricted were not important in the kind of
shells tested.
An additional similar auxiliary test was carried out by Anderson at NASA Langley a few
months later (see [9.250]) to complement the earlier supplementary tests. Similar stiffening
clips were added between the stiffening rings and corrugated wall of cylinder 2, and it was
again retested after having been rotated 180°. The result, however, reconfirmed the earlier
one that the addition of the stiffening clips contributed very little towards an increase in
buckling load.

13.7.2 Corrugated Shells Subject to External Pressure


As mentioned at the beginning of the previous subsection, circumferential corrugations have
been widely employed as stiffening against external pressure, primarily in agricultural silos.
These silos are generally loaded by a combination of axial compression and external pressure
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

and are therefore designed as circumferential corrugations (usually thin sinusoidal steel cor-
rugations) with discrete external axial stiJieners (often with many axial stiffeners, such as
40 or so). These stringer-stiffened corrugated shells proved to be very efficient and econom-
ical designs, but not without problems. These motivated extensive testing (see for example
[13.140] and [13.141]), some of which will be discussed in Chapter 15 in connection with
vibration correlation.
A different interesting application of circumferential corrugations against external pressure
loading is the swedge-stiffened (corrugated) submarine hull proposed by Ross of Portsmouth
University, U.K. [13.142]-[13.146]. As is well known, a submarine structure in its basic
form consists of two hulls, the outer hull being a hydrodynamic one, and the inner one being
a pressure-resisting one, as shown in Figure 13.78. The pressure hulls of most traditional
submarines consists of long lengths of circular cylindrical shells, which are stiffened by ring-
stiffeners to increase their buckling resistance.
An alternative method of increasing the shell and general instability resistance of these
cylinders is to swedge-stiffen (corrugate) them (see Figure 13.78a). Theoretical studies in-
dicated some years ago (see [13.142]) tha1: swedge-stiffened cylinders are structurally more
efficient than traditional ring-stiffened vessels (as shown in Figure 13.78b) from considera-
tions of general instability. Additional advantages of swedge-stiffened cylinders appeared to
be their better suitability for noise insulation and their better shock resistance during torpedo
launches, resulting from their being more elastic in the axial direction. A few years later
another theoretical (finite element) investigation on the influence of the swedge-cone angle
on the general instability of the swedge-stiffened shell was performed at Hsinchu, Taiwan
[13.143].
Ross then initiated a test program on small models at Portsmouth University [ 13.144].
Nine swedge-stiffened carbon steel cylinders, in three series (each of three nominally equal
shells), of diameter D = 73-99 mm, wall thickness t = 0.14-0.27 mm, swedged (corrugated)
length L = 68-110 mm and a number of swedges (corrugations) N = 14-28, were tested.
The corrugated models were made from ordinary commercial tin cans. The ends of the
tin can were cut away with a can opener and the cans were then push-fitted onto the end
closure plates. Then they were sealed at their ends with a silicone household sealant.
The corrugated tin cans were made by the can manufacturers by their usual process: First
a flat rectangular plate was rolled into a circular cylinder and seam welded longitudinally.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Special Stiffened Shells 1039

(a) pressure hull

water /' casing

----

(b) pressure hull

Figure 13.78 Submarine pressure hulls-schematic (from [13.144]): (a) swedge-stiffened pressure hull,
(b) ring-stiffened pressure hull

Then the ends of the circular cylinder were flanged to stiffen the can, which was then spun
about its axis. The corrugations were introduced with appropriate tools. The male and female
corrugating tools laterally compressed the curved surface of each can. Often the teeth of
these tools were of saw-tooth form.
Since this process is fully automatic, the dimensions of the mass-produced corrugated tins
have to be accurate or the production machine will stop. The process therefore includes a
built-in dimensional quality control. Hence accurate specimens are made rather inexpensively
if they are taken from a regular production run.
Here the specimens used had (Rit) = 184-261 and an effective (L/R) = 1.14-2.22.
A few years later a second test program was initiated at the University of Portsmouth on
partially corrugated cylindrical shells, in which the corrugation-stiffened parts are separated
by unstiffened shell elements (see [13.145]). The tests were again carried out on small mild-
steel models of diameter D = 70-153 mm and wall thickness t = 0.15-0.28 mm, with
relevant ratios (Rit) = 229-274, similar to the earlier test series. Nine shells were tested in
four series, of 2, 3, 3 and 1 specimens each, respectively. In one series of three models the
corrugations covered most of the shell with only narrow stiffened connecting parts. These
shells failed by general instability, as had the earlier fully corrugated ones, but the others
with wider unstiffened subshells failed by a special local buckling mode, in which the cor-
rugations at the borders of the widest (central) unstiffened shells buckled locally, at a sig-
nificantly lower pressure.
More recently, a further similar test program was carried out at Portsmouth [13.146] on
six slightly larger steel corrugated models, of diameter D = 153-154 mm and wall thickness
t = 0.31-0.35 mm. The relevant nondimensional ratios (R/t) = 218-248 and (LIR) = 194-
2.32 were, however, similar to the earlier tests. Two series, each of three nominally equal
shells (again to allow for repeatability), were tested. One series consisted of three fully
corrugated shells, as in the earliest experiments, and one of three partially corrugated ones,
as in the second test program. The shells were tested to destruction in a test rig (shown in
Figure 13.79 with a partially corrugated shell) in which they were hung from the top by a
special end ring. This facilitated attachment of strain gages to the internal midspan of the
cylinders. In the earlier test setup (of [13.144] and [13.145]) the shells were sealed off by
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1040 Stiffened Shells

<---
21 water under
~~~~~~~~~~--1·~~ pressure

Figure 13.79 Portsmouth University tests on small steel corrugated cylindrical shells under uniform
external pressure-test rig with a partially corrugated model in position (from [13.146])

two end caps and hung by a bolt from the center of the lid of the pressure vessel. In all
three test programs the end caps were simply pushed into ends and sealed with silicone.
Typical swedge (corrugation) shapes for the models were carefully measured in the earlier
tests with the aid of a toolmaker's microscope and in the later ones with the aid of a
coordinate measuring machine (CMM). The corrugations were of a sinusoidal nature, al-
though in the computer analysis the corrugations were assumed to consist of truncated con-
ical shell elements. Since earlier computations had shown that there was very little difference
in predicting the eigenvalues with truncated cone elements instead of the more sophisticated
truncated, doubly curved axisymmetric shell elements, it was considered justified to use the
simpler conical elements to represent the measured doubly curved shapes.
Ovality measurements were taken for each model shell with their end caps fitted. In the
earlier tests the shells were supported on \!-blocks and measured with a 0.01-mm graduated
dial gage, while in the later tests a coordinate measuring machine was used. The initial out-
of-roundness e (the maximum difference between the innermost and outermost points on the
outer surface of the cylinder) was recorded for all the models. Complete initial ovality plots
(the initial imperfection plots), as well as postbuckling out-of-roundness plots, were, however,
presented only for some shells. Since the magnitudes of the initial out-of-roundness were
significant, a more complete analysis of the actual initial imperfection shapes, in the manner
discussed in Chapter 10, might have partially explained the values of elastic knock-down
observed.
The fully corrugated shells failed in all test programs by general instability (see for ex-
ample Figure 13.80). Partially corrugated shells can fail by a local mode, or their failure
may be triggered by such a local mode (see [13.145]). Hence their design and analysis may
be more difficult.
Further investigations, which should include also other modes of failure than general
instability, are needed to determine the possible superiority of corrugation stiffening over the
conventional ring stiffening. Such investigations should also consider the influence of out-
of-plane and in-plane boundary conditions on the buckling and collapse pressure of corru-
gated shells and could probably benefit from the use of vibration correlation methods, dis-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

cussed in Chapter 15.

13. 7.3 Elastically Supported and Sandwich Shells


Elastically supported shells are a form of special stiffened shells, since the elastic supports
stiffen the shell against buckling and the combined flexural stiffness of the shell and the
elastic supports resist instability. Though there are many types of applications of elastically
supported shells (see for example [13.1471 and [13.148]), the most common type of elastic
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Special Stiffened Shells 1041

Figure 13.80 Portsmouth University testS on small steel corrugated cylindrical shells under uniform
external pressure-general instability of typical swedge-stiffened (corrugated) model
shell MBS 3, with D = 99 mm, (R/1) = 206, (L/ R) = 1. 14 and 14 c ircumferential
corrugations (courtesy of Professor C.T.F. Ross)

support is the soil support in buried flexible tubes, which has been of particular interest to
the oil industry and has therefore been extensively studied in the last decades.
Among the surveys of the field, a recent review of theory and experiment on the buckling
of buried flexible tubes [13.149] appears to be rather comprehensive and includes also many
references of experimental studies. Moore pointed out in that review that two types of buried
tube buckling theories have been developed:

I. Linear "multiwave" theories, which consider the li near elastic buckling strength of the
structure as a number of buckles form around the circumference
2. "Single-wave" theories, in which the stabil ity of the structure is considered as a single
buckle develops and moves into the cavity in the tube.

A preferred model and theory was selected. on the basis of both experimental and theo-
retical considerations, recommending the use of linear multi wave theory. There Moore con-
sidered the relative merits of the Winkler model (which models the soil as a series of in-
dependent springs resisting the radial deformation of the structure) and the elastic continuum
model and concluded that the elastic continuum theory represented most of the experimental
data bener, though it overpredicted the buckling loads. Hence, for structural analysis and
design, linear multiwave buckling solutions based on elastic continuum representation of the
soil appear most suitable.
Further reexamination of experimental data brought out the importance of the nonuni·
formity of the hoop force, which had not been considered by earlier investigators, who
compared their predictions with the average hoop force. When the comparisons with exper-
imental data used the maxi mum hoop thrust instead of the average value, the agreement
between prediction and experiment was significantly improved. Figure 13.81 shows the cotn-
parison of continuum multi wave theory with maximum hoop force for four major experi-
mental studies [13.150]- [13.153]. The agreement is indeed quite good.
One of the major difficulties in the tests has been the accurate assessment of the repre-
sentative elastic module of the ground surroundi ng the buried tube. Improvements in mea-
surement techniques would therefore be of great value.
Available buckling data are primarily for deep tubes (with burial depth ;;;,: tube diameter).
and hence there is a need for experimental studies of the effect of shallow burial. These and
further investigations on inelastic buckling seem to be major future research directions for
buried tubes.
In practice, buried tubes are sometimes further stiffened by circumferential ribs, rings or
longitudinal corrugations. Only very few theoretical studies (see for example [1 3.154]) and
--`,`,`````,`,```

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1042 Stiffened Shells

• • Allgood and Ciani (1968)A}


13 150
o o Allgood end Ciani (1968) 8 REf. ·
I 4
0
• • Howard (19721 REF. 13.151 •
0
• • Gumbel (19831 REF. 13.152
• 0
• • Crabb and C·order(I985)-
-I~EF.I3.153
~~H
«>W
103 -Continuum
'0
E
z

Y:~
El
Figure 13.81 Elastic buckling of buried flexible tubes-experimental data and theory, related to max-
imum hoop force N"'"' (from [ 13.149]). The graph is a logarithmic plot of the nondi-
mensionalized maximum hoop force per unit length versus the relative ground stiffness
Y (where E~ = [EJ(l - v~)] and E,, v., are the elastic constants of the soil)

no experimental investigations have been reported in the literature, though they are certainly
warranted.
Sandwich shells represent another form of special stiffened shells. Obviously, the discus-
sion on sandwich plates in Section 12.4 also applies here, in particular the typical sandwich
modes of failure (shown in Figure 12.33). Similarly, the special experimental problems of
sandwich construction outlined there for sandwich plates also apply to sandwich shells.
In the early sixties a substantial test program on sandwich shells was carried out at NASA
Langley as part of their buckling investigations of large cylindrical shells related to the
development of large space launchers (see for example [13.155] and [13.156]). The NASA
bending experiments [13.156] can be taken as a typical example of modern buckling tests
on sandwich shells, though they were performed three decades ago.
The 1.7-m diameter honeycomb sandwich cylinders were constructed with 7075-T6 alu-
minum-alloy face sheets, about 0.5 mm thick, and with 3003-H19 aluminum alloy, hexagonal
cell, bonded, honeycomb cores, having 6.35-mm cells with 0.051-mm cell walls, yielding
an overall sandwich radius-to-thickness ratio of 92-157. The cylinders were made with six
longitudinal splices (all with scalloped doublers bonded on) in each face sheet, but without
circumferential ones. They were assembled on a male-type mandrel and cured in one oper-
ation. Three shells were tested in the large Langley bending test rig (see Figure 9.75b). The
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

test procedure was similar to that of the other bending experiments discussed in Subsections
13.4.2 and 13.7.1.
The analysis employed in the assessment of the experimental results was an improved
version of earlier NASA analyses, the improvements being inclusion of both face sheet
inelasticity and transverse shearing effects as well as consideration of core and adhesive
contributions to the wall stiffness. The significance of these latter contributions to the stiffness
were observed in auxiliary compressive tests on small rectangular samples cut from the walls
of the test cylinders. As a result of inclusion of all these effects, the agreement between
calculated and measured strains was satisfactory. The experimental buckling loads were a
little below the calculated ones, 87-92 percent, comparable (though slightly lower) to other
stiffened cylindrical shells of similar radius-to-thickness ratios.
Buckling occurred suddenly in all the shells without prior visible evidence of impending
buckling, but strain gage data indicated that buckling was preceded by considerable wall
bending. Buckling was characterized by a wave that traveled around from the compression
to the tension side and left a line of failure, which made it appear as if the shell had been
sheared transversely.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Special Stiffened Shells 1043

Peterson and Anderson's conclusions [13.156], which have applications beyond their three
cylinder tests to sandwich structures in general even today, were:
1. That the adhesive and core in bonded honeycomb sandwich plates and shells contribute
substantially in carrying the load and in enhancing the stitinesses
2. That considerable core buckling, which significantly affects the core contribution to wall
stitiness, usually precedes cylinder buckling
3. That cylinder buckling for sandwich shells with moderately heavy cores can be predicted
by linear theory with some empirical reduction factors (though today these calculations
would be supplemented by more sophisticated nonlinear ones).
Before closing this subsection, it may be worth mentioning one type of sandwich con-
struction, the truss-core sandwich, or corrugated-core sandwich, which is essentially a cor-
rugated panel or shell with covering face plates. This type of sandwich construction was
already in use in the fifties (see for example [13.157], which includes also careful experi-
ments) and has been revived in recent years (see for example [13.137]).

13.7.4 Tube-Stiffened Shells


Another interesting idea was recently proposed by Ross of the University of Portsmouth:
stiffening of shells with pressurized tubes to increase their resistance to buckling under
external pressure or axial compression [13.158]. Essentially the idea is an extension of the
stiffening of shells by internal pressure, used in the fifties and sixties in some of the inter-
continental ballistic missiles (see for example [9.168], [9.184] or [9.185]), but here the con-
cept was taken one step further.
For example, a circular cylindrical shell, subjected to external pressure and/ or axial com-
pression, could be stiffened by closely spaced internal circumferential elastic tubular stiff-
eners (probably wound continuously as a helical coil), which would be under internal pres-
sure. This would stiffen the shell significantly to both types of loading and could yield a
higher strength-to-weight ratio of the shell. In the case of a submarine, for example, an
"intelligent" or "smart" hull could be designed that would increase its stiffening by increase

/Bolt
Disc

Tube
----Dome

/Water
/

"-,,
"'-- Co~m~ction to pump

(a)

(b)

Figure 13.82 Portsmouth University buckling tests on tube-stiffened prolate domes (from [13.158]):
(a) test setup for external pressure loading, (b) assembly of stiffening tube to dome
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1044 Stiffened Shells

of internal pressure in the stiffening tubes as the submarine dives deeper. Obviously much
work is still required to explore feasible designs and the actual overall efficiency gains
possible from this kind of stiffening.
Ross and his co-workers also carried out a preliminary experimental investigation on three
prolate hemi-ellipsoidal domes made of solid urethane plastic (SUP), which were stiffened
by nylon tubing (see Figure 13.82a). The domes had a base radius of about 250 mm, height
about 493 mm and wall thickness 3.8-3.9 mm. The nylon stiffening had an external diameter
of 9.5 mm and a wall thickness of 2 mm. The nylon tubing was stuck to the inside of each
dome with Evostick glue (except in the first specimen). The simple process of tubing at-
tachment, using an inverted dome as a reel, is shown in Figure 13.82b. Prior to testing, the
out-of-roundness of each dome was measured at three heights. Pressurizing the stiffening
tubing with water of 4-8 bars was found to increase the external buckling pressure of the
dome, measured in the pressure vessel, by 17-43 percent.
These experiments were presented primarily to indicate that there are alternatiVe ways of
stiffening shells that warrant assessment, and that some preliminary tests (which can some-
times be relatively simple) are an essential ingredient of such evaluations.

References
13.1 Singer, J., Buckling of Integrally Stitl'ened Cylindrical Shells-A Review of Experiment and
Theory, in: Contributions to the Theory of Aircraft Structures, Delft University Press, Rotter-
dam, 1972, 325-357.
13.2 Singer, J. and Baruch, M., Recent Studies on Optimisation for Elastic Stability of Cylindrical
and Conical Shells, in: Aerospace Proceedings 1966, Proceedings of the 5th International Con-
gress of the Aeronautical Sciences. London, Sept. 1966, Macmillan, London, 1967, 751-782.
13.3 Burns, A.B. and Alrnroth, B.O., Structural Optimization of Axially Compressed Ring-Stringer
Stiffened Cylinders, Journal of Spacecraft and Rockets, 3, (1), 1966, 19-25.
13.4 Milligan R., Gerard, G., Lakshmikantham, C. and Becker, H., General Instability of Orthotropic
Stiffened Cylinders under Axial Compression, AIAA Journal, 4, (11), 1966, 1906--1913.
13.5 Croll, J.G.A., Stiffened Cylindrical Shells under Axial and Pressure Loading, in: Shell Struc-
tures-Stability and Strength, R. Narayanan, ed., Elsevier, London and New York, 1985, 19-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

56.
13.6 Kendrick, S., Ring-Stiffened Cylinders under External Pressure, in: Shell Structures-Stability
and Strength, R. Narayanan, ed., Elsevier, London and New York, 1985, 57-95.
13.7 Green, D.R. and Nelson, H.M., Compression Tests on Large-Scale Stringer-Stiffened Tubes, in:
Buckling of Shells in Offshore Structures, J.E. Harding, P.J. Dowling and N. Agelidis, eds.,
Granada, London, 1982, 25-43.
13.8 Walker, A.C., Andronicou, A. and Sriclharan, S., Experimental Investigation of the Buckling of
Stiffened Shells Using Small Scale Models, in: Buckling of Shells in Offshore Structures, J.E.
Harding, P.J. Dowling and N. Agelidis, eds., Granada, London, 1982, 45-71.
13.9 Dowling, P.J. and Harding, J.E., Experimental Behaviour of Ring and Stringer Stiffened Shells,
in: Buckling of Shells in Offshore Structures, J.E. Harding, P.J. Dowling and N. Agelidis, eds.,
Granada, London, 1982, 73-107.
13.10 Gerard, G.. Minimum Weight Design of Ring Stiffened Cylinders under External Pressure,
Journal of Ship Research, 5, 1961, 44-49.
13.11 Crawford, R.F. and Burns, A.B., Minimum Weight Potentials for Stiffened Plates and Shells,
AIAA Journal, 1, 1963, 879-886.
13.12 Bodner, S.R., General Instability of a Ring-Stiffened Cylindrical Shell under Hydrostatic Pres-
sure, Journal of Applied Mechanics, 24, (2), June 1957, 269-277.
13.13 Thielemann, W.F., New Developments in the Nonlinear Theories of the Buckling of Cylindrical
Shells, in: Aeronautics and Astronautics, Proceedings of the Durand Centennial Conference
1959, Pergamon Press, Oxford, 1960, 76-119.
13.14 Singer, J. and Fersht-Scher, R., Buckling of Orthotropic Conical Shells under External Pressure,
Aeronautical Quarterly, 15, (2), May 1964, 151-168.
13.15 Singer, J., Fersht-Scher, R. and Belser, A., Buckling of Orthotropic Conical Shells under Com-
bined Torsion and External or Internal Pressure, in: Proceedings of the 6th Israel Annual Con-
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1045

ference on Aviation and Astronautics, Feb. 1964, Israel Journal of Technology, 2, (1), 179~
189.
13.16 van der Neut, A., The General Instability of Stiffened Cylindrical Shells under Axial Com-
pression, Report S314, NLL Amsterdam, Reports and Transactions, 13, 1947, 57~84.
13.17 Singer, J., Baruch, M. and Harari, 0., Inversion of the Eccentricity Effect in Stiffened Cylin-
drical Shells Buckling under External Pressure, Journal of Mechanical Engineering Science
(England), 8, (4), Dec. 1966, 363~373.
13.18 Baruch, M., Singer, J. and Weller, T., Effect of Eccentricity of Stiffeners on the General Insta-
bility of Cylindrical Shells under Torsion, Proceedings of the 8th Israel Annual Conference on

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Aviation and Astronautics, Israel Journal of Technology, 4, (1), Feb. 1966, 144~154.
13.19 Thielemann, W. and Esslinger, M., Uber den Einfluss der Exzentrizitat von Langssteifen auf
die axia1e Beullast di.innwandiger Kreiszylinderschalen, Stahlbau, 34, 1965, 332.
13.20 Bushnell, D., Computerized Analysis of Shells-Governing Equations, Computers and Struc-
tures, 18, (3), 1984, 471~536.
13.21 Card, M.F. and Jones, R.M., Experimental and Theoretical Results for Buckling of Eccentrically
Stiffened Cylinders, NASA TN D-3639, Oct. 1966.
13.22 Bums, A.B., Structural Optimization on Axially Compressed Cylinders, Considering Ring-
Stringer Eccentricity Effects, Journal of Spacecraft and Rockets, 3, (8), August 1966, 1263~
1268.
13.23 Block, D.L., Buckling of Eccentrically Stiffened Orthotropic Cylinders under Pure Bending,
NASA TN D-3351, 1966.
13.24 Simitses, G.J., Buckling of Eccentrically Stiffened Cylinders under Torsion, AIAA Journal, 6,
(10), Oct. 1968, 1856~1860.
13.25 Baruch, M. and Singer, J., General Instability of Stiffened Circular Conical Shells under Hy-
drostatic Pressure, Aeronautical Quarterly, 16, (2), May 1965, 187~204.
13.26 Simitses, G.J. and Cole, R.T., General Instability of Eccentrically Stiffened Thin Spherical
Shells under Pressure, Journal of Applied Mechanics, 37, Series E, (4), Dec. 1970, 1165~1168.
13.27 Soong, T.C., Influence of Boundary Constraints on the Buckling of Eccentrically Stiffened
Orthotropic Cylinders, presented at the 7th International Symposium on Space Technology and
Science, Tokyo, May 1967.
13.28 Weller, T., Baruch, M. and Singer, J., Influence of In-Plane Boundary Conditions on Buckling
under Axial Compression of Ring-Stiffened Cylindrical Shells, Proceedings of the 5th Annual
Conference on Mechanical Engineering, Israel Journal of Technology. 9, (4), 1979, 397~410.
13.29 Weller, T., Further Studies on the Effect of In-Plane Boundary Conditions on th Buckling of
Stiffened Cylindrical Shells, TAE Report No. 120, Technion Research and Development Foun-
dation, Haifa, Israel, Dec. 1971.
13.30 Weller, T., Combined Stiffening and In-Plane Boundary Condition Effects on the Buckling of
Circular Cylindrical Stiffened-Shells, Computers and Structures, 9, 1978. 1~16.
13.31 Singer, J. and Rosen, A., Design Criteria for Buckling and Vibration of Imperfect Stiffened
Cylindrical Shells, in: !CAS Proceedings 1974, Proceedings of the 9th Congress of the Inter-
national Council of the Aeronautical Sciences, Haifa, August 1974, R.R. Dexter and J. Singer,
eds., Weizmann Science Press of Israel, Jerusalem, 1974, 495~517.
13.32 Singer, J. and Rosen, A., Influence of Boundary Conditions on the Buckling of Stiffened Cy-
lindrical Shells, in: Buckling of Structures, Proceedings of IUTAM Symposium, Harvard Uni-
versity, Cambridge, Mass., June 17~21, 1974, B. Budiansky, ed., Springer-Verlag, Berlin, 1976,
227~250.
13.33Bushnell, D., Almroth, B.O. and Sobel, L.H., Buckling of Shells of Revolution with Various
Wall Constructions, 1~3, NASA CR's 1049, 1050 and 1051, 1968.
13.34 Harari, 0., Singer, J. and Baruch, M., General Instability of Cylindrical Shells with Non-
Uniform Stiffeners, Proceedings of the 9th Israel Annual Conference on Aviation and Astro-
nautics, Israel Journal of Technology, 5, (1), Feb. 1967, 1214~ 1228.
13.35 Baruch, M., Equilibrium and Stability Equations for Discretely Stiffened Shells, Israel Journal
of Technology, 3, (2), 1965, 138~146.
13.36 Singer. J. and Haftka, R.T., Buckling of Discretely Ring-Stiffened Cylindrical Shells, Proceed-
ings of the lOth Israel Annual Conference on Aviation and Astronautics, Israel Journal of
Technology, 5, (1~2), February 1968, 125~137.
13.37 Singer, J. and Haftka, R., Buckling of Discretely Stringer-Stiffened Cylindrical Shells and Elas-
tically Restrained Panels, AIAA Journal, 13, (7), July 1975 (synoptic), 849~850.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1046 Stiffened Shells

13.38 Block, D.L., Influence of Discrete Ring Stiffeners and Prebuckling Deformations on the Buck-
ling of Eccentrically Stiffened Orthotropic Cylinders, NASA TN D-4283, Jan. 1968.
13.39 Sewall, J.L. and Naumann, E.C., An Experimental and Analytical Vibration Study of Thin
Cylindrical Shells with and without Longitudinal Stiffeners, NASA TN D-4705, 1968.
13.40 Rosen, A. and Singer, J., Vibrations of Axially Loaded Stiffened Cylindrical Shells, Journal of
Sound and Vibration, 34, (3), June 1974, 357-378.
13.41 Singer, J., Vibrations and Buckling of Imperfect Stiffened Shells-Recent Developments, in:
Collapse: The Buckling of Structures in Theory and Practice, J.M.T. Thompson and G.W. Hunt,
eds., Cambridge University Press, Cambridge, 1983, 443-481.
13.42 Almroth, B.O. and Bushnell, D., Computer Analysis of Various Shells of Revolution, AJAA
Journal, 6, (10), Oct. 1968, 1848-1855.
13.43 Weller, T., Singer, J. and Batterman, S.C., Influence of Eccentricity of Loading on Buckling of
Stringer-Stiffened Shells, in: Thin Shell Structures, Theory, Experiment and Design, Prentice-
Hall, Englewood Cliffs, N.J, 1974, 305-324.
13.44 Weller, T. and Singer, J., Further Experimental Studies on Buckling of Integrally Ring-Stiffened
Cylindrical Shells under Axial Compression, Experimental Mechanics, 14, (7), July 1974, 268-
273.
13.45 Len'ko, O.N., The Stability of Orthotropic Cylindrical Shells, Raschet Prostranstvennykh Kon-
struktsii, (4), Moscow, 1958, 499-524, Translation NASA TT F-9826, July 1963.
13.46 Card, M.F., Preliminary Results of Compression Tests on Cylinders with Eccentric: Longitudinal
Stiffeners, NASA TM X-1004, Sept. 1964.
13.47 Katz, L., Compression Tests on Integrally Stiffened Cylinders, NASA TM X-55315, August
1965.
13.48 Almroth, B.O., Influence of Imperfections and Edge Restraint on the Buckling of Axially Com-
pressed Cylinders, NASA CR-432, April 1966.
13.49 Rosen, A. and Singer, J., Further Experimental Studies on the Buckling of Integrally Stiffened
Cylindrical Shells, TAE Report 207, Technion Israel Institute of Technology, Department of
Aeronautical Engineering, Haifa, Israel, August 1975.
13.50 Koiter, W.T., Buckling and Post-Buckling Behavior of a Cylindrical Panel under Axial Com-
pression, Report S.476, National Luchtvaartlaboratorium (NLL) Amsterdam, Reports and Trans-
actions, 20, 1956.
13.51 Hutchinson, J.W. and Frauenthal, J.C., Elastic Postbuckling Behavior of Stiffened and Barreled
Cylindrical Shells, Journal of Applied Mechanics, 36 (Trans. ASME, 91, Series E), 1969, 784-
790.
13.52 Weller, T. and Singer, J., Experimental Studies on the Buckling under Axial Compression of
Integrally Stringer-Stiffened Circular Cylindrical Shells, Journal of Applied Mechanics, 44, (4),
Dec. 1977, 721-730.
13.53 Slankard, R.C. and Nash, W.A., Tests of the Elastic Stability of a Ring-Stiffened Cylindrical
Shell, Model BR-5 (A = 1.705), Subjected to Hydrostatic Pressure, U.S. Navy Department.
David Taylor Model Basin, DTMB Report 822. 1953.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

13.54 Wenk, E., Jr., Slankard, R.C. and Nash, W.A., Experimental Analysis of the Buckling of Cy-
lindrical Shells Subjected to External Hydrostatic Pressure. Proceedings, Society for Experi-
mental Stress Analysis, 12, (1), 1954, 163-180.
13.55 Kirstein, A.F. and Slankard, R.C., An Experimental Investigation of the Shell Instability
Strength of a Machined, Ring-Sti!Iened Cylindrical Shell under Hydrostatic Pressure (Model
BR-4A), U.S. Navy Department, David Taylor Model Basin, DMTB Report 997, 1956.
13.56 Galletly, G.D., Slankard, R.C. and Wenk, E., General Instability of Ring-Stiffened Cylindrical
Shells Subject to External Hydrostatic Pressure-A Comparison of Theory and Experiment,
Journal of Applied Mechanics, 25, 1958, 259-266.
13.57 Hom, K. and Couch, W.P., Hydrostatic Tests of Inelastic and Elastic Stability of Ring-Stiffened
Cylindrical Shells Machined from Strain-Hardening Steel, U.S. Navy Department, David Taylor
Model Basin, DTMB Report 1501, Dec. 1961.
13.58 Couch, W.P., Hydrostatic Pressure Tests of a Ring-Stiffened Cylinder of Oval Cross Section
(Major-to-Minor Axis Ratio of 1.5), l.S. Navy Department, David Taylor Model Basin, DTMB
Report 1788, March 1964.
13.59 Midgley, W.R., Johnson, A.E., Jr., Experimental Buckling of Internal Integral Ring-Stiffened
Cylinders, Experimental Mechanics, 7, 1967, 145-153.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1047

13.60 Ross, C.T.F., Aylward, W.R. and Boltwood, D.T., General Instability of Ring-Reinforced Cir-
cular Cylinders under External Pressure, Transactions RINA, 113, 1971, 73-92.
13.61 Ross, C.T.F., Pressure Vessels under External Pressure: Statics and Dynamics, Elsevier Applied
Science. London and New York, 1990.
13.62 Odland, J., An Experimental Investigation of the Buckling Strength of Ring-Stiffened Cylin-
drical Shells under Axial Compression. Norwegian Maritime Research, 4, (9), 1981, 22-39.
13.63 Yamamoto, Y., Homma, Y., Oshima, K., Mishiro, Y., Terada, H., Yoshikawa, T., Morihana, H.,
Yamauchi, Y. and Takenaka, M., General Instability of Ring-Stiffened Cylindrical Shells under
External Pressure, Marine Structures, 2, 1989, 133-149.
13.64 Yamamoto, Y., General Instability of a Reinforced Cylindrical Shell Clamped at Both Ends
under External Pressure, in: Proceedings of the Fourteenth Japan National Congress for Applied
Mechanics, 1964, 63-71.
13.65 Walker, A.C. and Davies, P., The Collapse of Stiffened Cylinders, in: Steel Plated Structures,
P.J. Dowling, J.E. Harding and P.A. Frieze, eds., Crosby Lockwood Staples, London, 1977,
791-808.
13.66 Walker, A.C., Andronicou, A. and Sridharan, S., Local Plastic Collapse of Ring-Stiffened Cyl-
inders, Proceedings of the Institution of Civil Engineers, 71, Part 2, June 1981, 341-367.
13.67 Dowling, P.J. and Harding, J.E., Current Research into the Strength of Cylindrical Shells Used
in Steel Jacket Construction, in: BOSS 79, Proceedings of the 2nd International Conference on
the Behaviour of Offshore Structures, Imperial College, London, August 1979, 327-340.
13.68 Dowling, P.J., Harding, J.E., Agelidis, N. and Fahy, W., Buckling of Orthogonally Stiffened
Cylindrical Shells Used in Offshore Engineering, in: Buckling of Shells, Proceedings of a State-
of-the-Art Colloquium, Stuttgart, May 1982, E. Ramm, ed., Springer-Verlag, Berlin, Heidelberg,
New York, 1982, 239-273.
13.69 Dowling, P.J., Harding, J.E., Fahy, W. and Agelidis, N., Report on the Testing of Large and
Small Scale Stiffened Shells under Axial Compression, U.K. Department of Energy Report OT-
R-8107, August 1981.
13.70 Scott, N.D., Harding, J.E. and Dowling, P.J., Fabrication of Small Scale Stiffened Cylindrical
Shells, Journal of Strain Analysis, !MechE, 22, (2), 1987, 97-106.
13.71 Grove, T. and Didriksen, T., Buckling Experiments on 4 Large Ring-Stiffened Cylindrical Shells
Subjected to Axial Compression and Lateral Pressure, Report No. 77-431, Det norske Veritas,
Oslo, 1977.
13.72 Miller, C.D. and Kinra, R.K., External Pressure Tests of Ring-Stiffened Fabricated Steel Cyl-
inders, 13th Annual Offshore Technology Conference, Houston, May 1981, Paper No. OTC
4107, Proceedings OTC '81, 3, 1981, 371-386.
13.73 Palchevskii, A.S. and Polyakov, P.S., Experimental Investigation of the Stress-Strain State and
Stability of Rib-Reinforced Cylindrical Shells with Sizable Rectangular Holes, Prikladnaia
Mekhanika, 12, Oct. 1976, 118-122 (in Russian).
13.74 Grinenko, N.I., Khishchenko, I.M. and Chernoglazov, G.S., Experimental Investigation of the
Stability of a Reinforced Cylindrical Shell in Pure Bending, Prikladnaia Mekhanika, 12, May
1976, 50-55 (in Russian).
13.75 Miller, C.D., Frieze, P.A., Zimmer, R.A. and Jan, H.Y., Collapse Tests of Fabricated Stiffened
Steel Cylinders under Combined Loads, presented at the ASME 4th National Congress of
Pressure Vessels and Piping Technology, Portland, Ore., June 1983, Paper 83-PVP-7.
13.76 Chen, Y., Zimmer, R.A., de Oliveira, J.G. and Jan, H.Y., Buckling and Ultimate Strength of
Stiffened Cylinders: Model Experiments and Strength Formulations, presented at 17th Annual
OTC, Houston, Tex., May 1985, Paper OTC 4853.
13.77 Groth, H., Buckling of Cylinders with Longitudinal Stiffeners Subjected to Axial Compression,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

FFA Technical Note HU-2184, Stockholm, 1979.


13.78 Esslinger, M. and Geier, B., Buckling and Postbuckling Behavior of Discretely Stiffened Thin-
Walled Circular Cylinders, Zeitschrift fur Flugwissenschaften, 18, (7), 1970, 240-253.
13.79 Miller, C.D., Experimental Study of the Buckling of Cylindrical Shells with Reinforced Open-
ings, presented at the ASME/ ANS Nuclear Engineering Conference, Portland, Ore., July 1982.
13.80 Tennyson, R.C., Booton, M. and Hui, D., Changes in Cylindrical Shell Buckling Behavior
Resulting from System Parameter Variations, in: Theoretical and Applied Mechanics, Proceed-
ings of the 15th IUTAM Congress, Toronto, 1980, F.P.J. Rimrott and B. Tabarrok, eds., North-
Holland, Amsterdam and New York, 1980, 417-430.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1048 Stiffened Shells

13.81 Williams, J.G. and Davies, R.C., Buckling Experiments on Stiffened Cast-Epoxy Conical Shells,
Experimental Mechanics, 15, (9), Sept. 1975, 329-338.
13.82 Mungan, I., Buckling Stresses of Stiffened Hyperboloidal Shells, Journal of the Structural
Division, ASCE, 105, (ST8), August 1979, 1589-1604.
13.83 Tillman, S.C., Some Effects of Rib-Reinforcement Arrangement on Spherical Dome-Buckling,
Experimental Mechanics, 18, (10), Oct. 1978, 396-400.
13.84 Babel, H.W., Christensen, R.H. and Dixon, H.H., Design, Fracture Control, Fabrication and
Testing of Pressurized Space Vehicle Structures, in: Thin Shell Structures, Theory, Experiment
and Design, Y.C. Fung and E.E. Sechler, eds., Prentice-Hall, Englewood Cliffs, N.J., 1974,
549-603.
13.85 Slysh, P., Dyer, J.E., Furman, J.H. and Key, J.E., Isogrid Structural Tests and Stability Analysis,
Journal of Aircraft, 13, (10), Oct. 1976, 778-785.
13.86 Heard, W.L., Anderson, M.S. and Slysh, P., An Engineering Procedure for Calculating Com-
pressive Strength of Isogrid Cylindncal Shells with Buckled Skins, NASA TN D-8239, June
1976.
13.87 Horton, W.H., On the Elastic Stability of Shells, NASA CR-145088, 1977.
13.88 Walker, A.C. and Sridharan, S., Buclding of Compressed, Longitudinally Stiffened Cylindrical
Shells in: BOSS 79, Proceedings of the 2nd International Conference on the Behaviour of
Offshore Structures, Imperial College, London, August 1979, 2, 341-356.
13.89 Harding, J.E., Dowling, P.J, Valsgard, S. and Walker, A.C., The Buckling Design of Stringer-
Stiffened Shells Subject to Combined Pressure and Axial Compression, Paper OTC 4473, in:
Proceedings of 15th Annual OTC, Houston, Tex., May 1983, 267-276.
13.90 Miller, C.D. and Grove, R.B., Current Research Related to Buckling of Shells for Offshore
Structures, Paper OTC 4474, in: Proceedings of 15th Annual OTC, Houston, Tex., May 1983,
277-288.
13.91 Tsang, S.K., Harding, J.E., Walker, A.C. and Andronicou, A., Buckling of Ring Stiffened Cyl-
inder Subjected to Combined Pres>ure and Axial Compressive Loading, Paper 83-PVP-4,
ASME, New York, 1983.
13.92 Chen, Q., Elwi, A.E. and Kulak, C:.L., Bending Strength of Longitudinally Stiffened Steel
Cylinders, Structural Engineering Report 192, Department of Civil Engineering, University of
Alberta. Edmonton, Alta., Canada, August 1993.
13.93 Stephens, W.B., Imperfection Sensitivity of Axially Compressed Stringer Reinforced Cylindrical
Panels under Internal Pressure, AIAA Journal, 9, (9), Sept. 1971, 1713-1719.
13.94 Singer, J. and Abramovich, H., Vibration Techniques for Definition of Practical Boundary Con-
ditions in Stiffened Shells, AIAA Journal, 17, (7), July 1979, 762-763.
13.95 Singer, J. and Abramovich, H., The Influence of Practical Boundary Conditions on the Vibra-
tions and Buckling of Stiffened Cylindrical Shells, TAE Report 288, Technion-Israel Institute
of Technology, Department of Aeronautical Engineering, Haifa. Israel, March 1978.
13.96 Singer, J., Recent Studies on the Correlation between Vibration and Buckling of Stiffened
Cylindrical Shells, Zeitschrift ji"ir Flugwissenschaften und Weltraumforschung, 3, (6), Nov.-
Dec. 1979, 333-343.
13.97 Kinra, R.K., Hydrostatic and Axial Collapse Tests of Stiffened Cylinders, Paper OTC 2685, in:
Proceedings of 8th Offshore Technology Conference, Houston, Tex., May 1976. 765-788,
13.98 Peterson, J.P. and Dow, M.B., Compression Tests on Circular Cylinders Stiffened Longitudinally
by Closely Spaced Z-Section Stringers, NASA MEMO 2-12-59L, March 1959.
13.99 Dickson, J.N. and Brolliar, R.H., The General Instability of Ring-Stiffened Corrugated Cylinders
under Axial Compression, NASA TN D-3089, Jan. 1966.
13.100 Card, M.F., Bending Tests of Large-Diameter Stiffened Cylinders Susceptible to General Insta-
bility, NASA TN D-2200, April 1964.
13.101 Akiyama, H., Yuhara, T., Shimizu, S. and Takahashi, T., Limit State of Steel Cylindrical Struc-
tures under Earthquake Loadings, in: Inelastic Behaviour of Plates and Shells, IUTAM Sym-
posium Rio de Janeiro 1985, I. Bevilacqua, R. Feijoo and R. Valid, eds., Springer-Verlag, Berlin,
Heidelberg, 1986, 263-290.
13.102 Yuhara, T. and Masanori, T., Recent Experimental Research into Buckling Design Methods for
FBR Components, International Journal of Pressure Vessels and Piping, Elsevier Science, 37,
1989, 243-261.
13.103 Kawamoto, Y. and Yuhara, T., Buckling of Fabricated Ring-Stiffened Steel Cylinders under
--`,`,`````,`,````,,``,``,,`-`-

Axial Compression, in: Advances in Marine Structures, C.S. Smith and J.D. Clarke, eds., El-
sevier Applied Science, London, New York, 1986, 262-280.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1049

13.104 Ross, C.T.F., The Collapse of Ring-Reinforced Cylinders Under Uniform External Pressure,
Transactions of the Royal Institute of Naval Architects (RINA), 107, 1965, 375~394.
13.105 Dow, D.A., Buckling and Postbuckling Tests of Ring-Stiffened Circular Cylinders Loaded by
Uniform External Pressure, NASA TN D-3111, Nov. 1965.
13.106 Dow, D.A. and Peterson, J.P., Test of a Large-Diameter Ring-Stiffened Cylinder Subjected to
Hydrostatic Pressure, NASA TN D-3647, Oct. 1966.
13.107 Abramovich, H. and Singer, J., Correlation between Vibration and Buckling of Stiffened Cy-
lindrical Shells under External Pressure and Combined Loading, Proceedings of the 20th Israel
Annual Conference on Aviation and Astronautics, Israel Journal of Technology, 16, (1 ~2), 1978,
34~44.
13.108 Walker, A.C., Segal, Y. and McCall, S., The Buckling of Thin-Walled Ring-Stiffened Shells,
in: Buckling of Shells, Proceedings of the State-of-the-Art Colloquium, Stuttgart University,
Germany, May 6~ 7, 1982, Springer-Verlag, Berlin, Heidelberg, New York, 1982, 275~304.
13.109 Walker, A.C. and McCall, S., Buckling Tests on Stringer Stiffened Cylinder Model Subjected
to Load Combination, Det norske Veritas Report 82-0299, 1982.
13.110 Seleim, S.S. and Roorda, J., Buckling Behaviour of Ring-Stiffened Cylinders; Experimental
Study, Thin-Walled Structures, 4, 1989, 203~222.
13.111 Kendrick, S., The Buckling under External Pressure of Circular Cylindrical Shells with Evenly
Spaced Equal Strength Circular Ring Frames-Part I, U.K. Naval Construction ResearchEs-
tablishment Report No. 211, Feb. 1953.
13.112 Chen, Q., Elwi, A. E. and Kulak, G.L., Longitudinally-Stiffened Large Diameter Fabricated Steel
Tubes, Journal of Constructional Steel Research, 35, 1995, 19~48.
13.113 Chen, Q., Kulak, G.L. and Elwi, A.E., Flexural Tests on Longitudinally-Stiffened Fabricated
Steel Cylinders, Journal of Constructional Steel Research, 34, (1), 1995, 1~25.
13.114 Kulak, G.L., Stephens, M.J. and Bailey, R.W., Bending Tests of Large Diameter Fabricated
Steel Cylinders, Canadian Journal of Civil Engineering, 15, 1988, 183~189.
13.115 Dunn, L.G., Some Investigations of the General Instability of Stiffened Metal Cylinders VIII-
Stiffened Metal Cylinders Subjected to Pure Torsion, NACA TN 1197, March 1947.
13.116 Milligan, R. and Gerard, G .. General Instability of Orthotropically Stiffened Cylinders under
Torsion, AIAA Journal, 5, (11), Nov. 1967, 2071~2073.
13.117 Shalev, D., Segal, A. and Frostig, Y., Post Buckling of Laminated Composite Stiffened Curved
Panels Subjected to Cyclic Shear and Compression, in: !CAS Proceedings 1992, Proceedings
of the 18th Congress of the International Council of the Aeronautical Sciences, Beijing, People's
Republic of China, Sept. 1992, (distributed by AIAA, Washington, D.C.), 1460~1466.
13.118 Segal, A., Frostig, Y., Shalev, D., Weller, T. and Sheinman, I., Post Buckling Behaviour of
Stiffened Composite Panels Loaded in Cyclic Compression and Shear, Collection of Papers,
33rd Israel Annual Conference on Aviation and Astronautics, Tel-Aviv and Haifa, February 24~
25, 1993, 225~236.
13.119 Baruch, M., Singer, J. and Harari, 0., Instability of Conical Shells with Non-Uniformly Spaced
Stiffeners under Hydrostatic Pressure, Proceedings of the 7th Israel Annual Conference on
Aviation and Astronautics, Israel Journal of Technology, 3, (1), Feb. 1965, 62~71.
13.120 Singer, J., Berkovits, A., Weller, T., Ishai, 0., Baruch, M. and Harari, 0., Experimental and
Theoretical Studies on Buckling of Conical and Cylindrical Shells Under Combined Loading,
TAE Report 48, Department of Aeronautical Engineering, Technion, Israel Institute of Tech-
nology, June 1966, Section 5.
13.121 Singer, J. Eckstein, A., Fersht-Scher, R. and Berkovits, A., Buckling of Isotropic, Orthotropic
and Ring-Stiffened Shells, TAE Report 30, Department of Aeronautical Engineering, Tech-
nion-Israel Institute of Technology, Sept. 1963. Section 1.
13.122 Weller, T., Buckling of Ring-Stiffened Conical Shells under Axial Compression, Torsion and
Combined Axial Compression and Torsions. M.Sc. Thesis, Faculty of Aeronautical Engineering,
Technion-Israel Institute of Technology, Haifa, Israel, May 1967 (in Hebrew).
13.123 Heard, W.L., Anderson, M.S., Anderson, J.K. and Card, M.F., Design, Analysis, and Tests of
a Structural Prototype Viking Aeroshell, Journal of Spacecraft and Rockets, 10, (l ), Jan. 1973,
56~65.
13.124 Anderson, J.K. and Davis, R.C., Buckling Tests of Two 4.6-Meter-Diameter Magnesium Ring-
Stiffened Conical Shells Loaded under External Pressure, NASA TN D-7303, 1973.
13.125 Ross, C.T.F. and Johns, T., Buckling and Vibration of Ring-Stiffened Cones under Uniform
External Pressure, Thin- Walled Structures, 6, 1988, 321 ~342.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1050 Stiffened Shells

13.126 Ross, C.T.F. and Hamer, H., Inelastic Buckling of Ring-Stiffened Circular Conical Shells under
External Pressure, in: Advances in Structural Engineering Computing, B.H.V. Topping, and M.
Papadrakakis, eds., Civil-Camp. Press, Edinburgh, 1994, 109-113.
13.127 Ross, C.T.F. and Johns, T., Vibration of Hemi-EIIipsoidal Axisymmetric Domes Submerged in
Water, Proceedings of the Institution of Mechanical Engineers, 200, 1986, 389-398.
13.128 d'Herouville, M., Methodes experimentales resultats d'essais sur des coques de revolution,
Revue Francaise de Mecanique, (62-63), 1977, 61-65.
13.129 Sobel, L.H., Weller, T. and Agarwal, B.L., Buckling of Cylindrical Panels under Axial Com-
pression, Computers and Structures, ti, 1976, 29-35.
13.130 Sobel, L.H., and Agarwal, B.L., Buckling of Eccentrically Stringer-Stiffened Cylindrical Panels
under Axial Compression, Computers and Structures, 6, 1976, 193-198.
13.131 Agarwal, B.L., Postbuckling Behavior of Composite-Stiffened-Curved Panels Loaded in Com-
pression. Experimental Mechanics, 22, June 1982, 231-236.
13.132 Lunsford, L.R., Structural Panel Stability Program, Final Report-Phase I, Lock11eed Missiles
and Space Company, Sunnyvale, CA, Report LMSC A746888, April 1965.
13.133 Structural Panel Stability Development Program, Final Report-Phases II and III, Lockheed
Missiles and Space Company, Sunnyvale, Calif., Report LMSC A770422, Oct. 1965.
13.134 Bushnell, D., Holmes, A.M.C., Flaggs . D.L. and McCormick, P.J., Optimum Design, Fabrication
and Test of Graphite-Epoxy, Curved, :Stiffened, Locally Buckled Panels Loaded in Axial Com-
pression. in: Buckling of Structures: Theory and Experiment, I. Elishakoff, J. Arbocz, C.D.
Babcock and A. Libai, eds., Studies in Applied Mechanics 19, Elsevier, Amsterdam, New York,
1988, 61-131.
13.135 Bushnell, D., PANDA 2-Program for Minimum Weight Design of Stiffened, Composite, Lo-
cally Buckled Panels, Computers and Structures, 25, 1987, 469-605.
13.136 Bushnell, D., Use of PANDA 2 to Optimize Composite, Imperfect, Stiffened, Locally Buckled
Panels under Combined In-Plane Loads and Normal Pressure, in: Design and Analysis of Com-
posite Material Vessels, D. Hui and T. Kozik, eds., PVP, 121, ASME, 1987, 21-42.
13.137 Bushnell, D., Truss-Core Sandwich Design via PANDA 2, Computers and Structures, 44, (5),
1992, 1091-1119.
13.138 Holmes, A.M.C., Design, Operation and Use of the Stringer-Stiffened Panel Test Frame (A
User's Guide), Lockheed Palo Alto Research Laboratory, LMSC Report F110854, Palo Alto,
Calif., May 1986.
13.139 Bushnell, D., Crippling and Buckling of Corrugated Ring-Stiffened Cylinders, Journal of Space-
craft and Rockets, 9, (5), May 1972, 357-363.
13.140 Ory, H. and Hoffmann, H., Buckling and Vibration of Corrugated Shells, in: Buckling of Struc-
tures, Theory and Experiment, I. Elishakoff, J. Arbocz, C.D. Babcock and A. Libai, eds., Studies
in Applied Mechanics 19, Elsevier, Amsterdam, New York, 1988, 285-312.
13.141 Ory, H. and Hoffmann, H., Stability of Corrugated Shells, in: Proceedings of the ECCS Col-
loquium on Stability of Plate and Shell Structures, Ghent University, April 1987, P. Dubas and
D. Vandepitte, eds., Ghent, Belgium, 1987, 567-576.
13.142 Ross, C.T.F., A Novel Submarine Pressure Hull Design, Journal (Jf Ship Research, 31, (3), Sept.
1987, 186-188.
13.143 Yuan, K.-Y., Liang, C.-C. and Ma, Y.-C., Investigation of the Cone Angle of a Novel Sweclge-
Stiffenecl Pressure Hull, Journal of Ship Research, 35, (1), March 1991, 83-86.
13.144 Ross, C.'f.F. and Palmer, A., General Instability of Sweclge-Stiffenecl Circular Cylinders under
Uniform External Pressure, Journal of Ship Research, 37, (1), March 1993, 77-85.
13.145 Ross, C.T.F. and Humphries, M., The Buckling of Corrugated Circular Cylinders under Uniform
External Pressure, Thin- Walled Structures, 17, 1993, 259-271.
13.146 Ross, C.T.F. and Heigl, T., Buckling of Corrugated Axisymmetric Shells Under Uniform Ex-
ternal Pressure, in: Structural Dynamics and Vibrations 1995, ASME PD-Vol. 70, B.A. Ovung,
1.1. Esat, A.B. Sabir and V. Karadag, eels., ASME, New York, 1995, 199-205.
13.147 Ory, H., Die Stabilitat eines vorgespannten Doppelmanteltanks mit elastischer Zwischenschicht,
in: Hans Ebner Gedachtnis-Kolloquium am 27128 Oktober 1977 in Aachen, Mitteilung, Institut
fur Leichtbau, RWTH Aachen, April 1978, 268-315.
13.148 Yamamoto, Y. and Matsubara, N., Buckling of a Cylindrical Shell under External Pressure
Restrained by an Outer Rigid Wall, in: Collapse: The Buckling of Structures in Theory and
Practice, J.M.T. Thompson and G.W. Hunt, eels., Cambridge University Press, Cambridge, 1983,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

493-504.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1051

13.149 Moore, J.D., Elastic Buckling of Buried Flexible Tubes-A Review of Theory and Experiment,
Journal of Geotechnical Engineering, ASCE, 115, (3), March 1989, Paper No. 23627, 340-
358.
13.150 Allgood, J.R. and Ciani, J.B., The Influence of Soil Modulus on the Behaviour of Cylinders in
Sand. Highway Research Rec. No. 249, 1968, 1-13.
13.151 Howard, A.K., Laboratory Load Tests on Buried Flexible Pipe, Journal of the American Water
Works Association, 64, (10), 1972, 655-662.
13.152 Gumbel, J.E., Analysis and Design of Buried Flexible Pipes, Ph.D. thesis, University of Surrey,
Surrey, U.K., 1983.
13.153 Crabb, G.I. and Carder, D.R., Loading Tests on Buried Flexible Pipes to Validate a New Design
Model, Research Report 28, Transportation and Road Research Laboratory, Department of the
Environment, Crowthorne, U.K., 1985.
13.154 Moore, l.D., Influence of Rib Stiffeners on the Buckling Strength of Elastically Supported
Tubes, International Journal of Solids and Structures, 26, (5/6), 1990, 539-547.
13.155 Cunningham, J.H. and Jacobson, M.J., Design and Testing of Honeycomb Sandwich Cylinders
under Axial Compression, in: Collected Papers on Instability of Shell Structures-I962, NASA
TN D-1510, 1962, 341-360.
13.156 Peterson, J.P. and Anderson. J.K., Structural Behavior and Buckling Strength of Honeycomb
Sandwich Cylinders Subjected to Bending, NASA TN D-2926, August 1965.
13.157 Libove, C. and Hubka, R.E., Elastic Constants for Corrugated-Core Sandwich Plates, NACA
TN 2289, Feb. 1951.
13.158 Ross. C.T.F. and Popken, D., Buckling of Tube-Stiffened Prolate Domes under External Water
Pressure, Thin- Walled Structures, 22, 1995, 159-179.
13.159 Tong, L., Tabarokk, B. and Wang, T.K., Simple Solutions for Buckling of Orthotropic Conical
Shells, International Journal of Solids and Structures, 29, (8), 1992, 933-946.
--`,`,`````,`,````,,``,``,,`-`-`,,`

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
14
Composite Structures

14.1 Background
Advanced lightweight laminated composite structural elements are increasingly being intro-
duced into new designs of modern aerospace structures for enhancing their structural effi-
ciency and performance. In parallel, in recognition of the numerous advantages that com-
posites offer, there is a steady growth in replacement of metallic components by composite

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
ones in marine structures, ground transportation, robotics, sports and other fields of engi-
neering.
Like metallic isotropic and orthotropic structures, composite laminated structures are sus-
ceptible to buckling when they are exposed to significant compression, shear, pressure and
combinations of these loadings. Their buckling load-carrying capacity is affected by initial
geometrical imperfections, boundary conditions and load eccentricity. The extent of these
effects depends on the type of structure under consideration. In particular, they are empha-
sized when the theoretically predicted critical loads of shell elements are compared with
experimental observations. Usually there exists a significant discrepancy between theory and
experiments, unless the boundary conditions and load eccentricity are as closely as possible
realistically modeled in the computations and the actual measured initial geometrical imper-
fections are introduced into the calculations as well.
The buckling and postbuckling behavior of laminated composite structures are affected by
additional factors, and new failure modes appear that are not experienced in structures made
of conventional materials. These additional factors make the design of composite structures
more challenging than that of metallic isotropic and orthotropic structures. Elastic couplings
are introduced into the mechanical behavior when unbalanced and more pronounced unsym-
metric laminates are employed in the design. These couplings can be exploited to achieve
critical performance advantages, as in aeroelastic tailoring. On the other hand, their presence
degrades the load-carrying capacity and stiffness. This occurs, for example, in the case of a
drop-off in a laminate construction (see for instance [14.1] and [14.2]), where due to design
constraints the laminate thickness is tapered by dropping off individual plies in an unsym-
metric fashion, and when local delaminations exist. Delaminations may also stem from im-
pact damage and poor fabrication, or result from significant interlaminar stresses developed
in highly stressed and deformed elements. In particular, they develop in the webs of stiffened
structures designed to operate in the postbuckling region in order to enhance their structural
efficiency, thus avoiding complete exploitation of the postbuckling strength capacity of these
structures. High interlaminar stresses leading to delamination are also experienced along free
(exposed fibers) unstiffened/unreinforced edges of laminates, as well as discontinuities (cut-
outs).

CopyrightBuckling
Wiley Experiments: Experimental Methods in Buckling of Thin-Walled Structures: Shells, Built-Up Structures, Composites
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
and Additional Topics – Volume 2. J. Singer, J. Arbocz and T. Weller
No reproduction or networking permitted without license from IHS
Copyright © 2002 John Wiley & Sons, Inc.
Not for Resale, 02/13/2019 01:32:58 MST
1054 Composite Structures

An additional degrading factor, particularly associated with buckling and postbuckling of


stiffened composite structures, is stitiener/web separation. Microbuckling phenomena (see
for example [14.3] and [14.4]), which reduce the compressive strength of laminated com-
posite structures and lead to their premature failure, may also significantly decrease the
buckling load of laminated composite structures as well as affect their postbuckling char-
acteristics. Microbuckling can stem from presence of manufacturing imperfections in the
form of voids, poor interlaminar bonding and misaligned fibers, or from inherent character-
istics of the fiber-reinforced composites that influence their compressive strength capability:
fiber volume fraction, fiber array, specimen geometry, fiber diameter, fiber and matrix prop-
erties and fiber end fixity within the laminated material.
The aforementioned problems and shortcomings, associated with the introduction of light-
weight laminated composites were identified and recognized by the researchers involved in
their development. They have realized that in order to exploit fully the potential of com-
posites and the advantages provided by them, the analysis, testing and design of composite
structural elements should, in addition to satisfying traditional design requirements typical
of metal structures, also account for the characteristics and failure modes unique to com-
posites. In order to provide reliable analytical tools for efficient designs, which are validated
by experimental observations, the buckling and postbuckling characteristics and the special
features of laminated composites had to be better understood and properly addressed in the
design of composite test specimens and :test setups. Test procedures, in compliance with the
special characteristics and anticipated failure modes of composites, had to be developed for
reliable testing of composite elements, as well as for detection and identification of unknown
features peculiar to composites, to facilitate the development of improved and dependable
analytical tools.
Though laminated composite structures constitute a relatively new field of engineering,
numerous studies on their buckling and postbuckling behavior, particularly theoretical ones,
were carried out. Leissa reviewed abom 400 articles on buckling of laminated composite
plates and shell panels in [14.5], and many additional studies in [12.89]. A large number of
investigations are also cited and reviewed in r14.6]-[14.11].
In the following sections of this chapter, experimental studies on buckling and postbuck-
ling behavior of composite structural elements will be described and evaluated . First, buck-
ling and postbuckling experiments with fiat unstiffened and stiffened panels subjected to
axial compression, shear and combined Loading will be considered. Crippling strength tests
of thin-walled stiffener elements and postbuckled stiffened panels will also be addressed.
Next, experimental programs on wing-box structures and on curved panels and cylindrical
shell elements, unstiffened and stiffened, under axial compression, shear, pressure and com-
bined loadings will be described and discussed. Some theoretical aspects and considerations
pertinent to thin-walled laminated, fiat and curved, structural elements will also be presented.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

14.2 Flat Panels


14.2. 1 Unstiffened Panels
a. Theoretical Considerations
The methodology for determining and analyzing the buckling behavior of laminated com-
posite plates is in essence identical to that applied to isotropic plates. As in isotropic plates,
it involves the solution of an eigenvalue problem associated with a governing set of homo-
geneous differential equations and a prescribed set of homogeneous boundary conditions (see
Section 2 of Chapter 1 and Section 3 of Chapter 3).
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1055

In the case of isotropic plates exact buckling solutions are available only for a few com-
binations of loading and boundary conditions. Most buckling problems are solved by ap-
proximate methods: Rayleigh-Ritz, Galerkin, series, finite element and finite difference. As
will now be shown, solution of the buckling behavior problem of laminated composite plates,
as reflected by the bifurcation buckling equations, poses an even tougher challenge due to
complications inherent to laminated materials.
Here the relation between the in-plane stress and out-of-plane moment resultants N,, N,.,
M,, M, and M,1 to the midplane strains and curvature changes, assuming the Kirchhoff
hypothesis, is given by (see [14.12]):

N, A,, A,2 A,r, 8,, 8,2 8,6 E~~


N, A,2 An A26 8,2 822 826 e~~
Nn. A,, A26 A66 8,6 826 866 i:,
(14.1)
M, 8" 8,2 8,6 D" D,2 D,6 Kx
M, 8,2 822 B2r. D,2 D22 D26 K,

Mn 8,6 82(, 866 D,, D26 D66 K_n'

where Au, 8,i and Du are the stiffness matrix coefficients given by:
h/1

A,i' =
f --hi 2
Cudz

(14.2)

Cu are constants that change from layer to layer during the integration (see for example
Chapter 4 of [14.12]), his the laminate thickness and z is the distance of the mid-plane of
a ply measured relative to the laminate mid-plane. Note that the 3 X 3 Au, B,i' and DiJ
submatrices and the 6 X 6 stiffness matrix of Eq. (14.1) are symmetric.
The tem1s D 16 and D26 , in the stitiness coefficients matrix ofEq. (14.1), introduce bending-
twisting coupling, and the 8,j terms lead to bending-stretching coupling. The existence of
these coupling terms depends upon the alignment of the fibers in the plies that constitute
the laminate relative to the boundaries of the plate, as well as the stacking sequence of the
plies that compose the laminate. Plates in which the fiber orientation in a ply at a distance
(+ z) from mid-plate is identical with the fibers in the ply at (- z) and the plies are of the
same material and identical thickness are said to be "symmetric". Otherwise, they are "un-
symmetric". Symmetric plates in which the fibers in the laminae are oriented in oo and 90°
to the plate axes are designated as orthotropic cross-ply laminates. It is apparent from Eq.
(14.2) that the 8u terms vanish in "symmetric" laminates, thus simplifying the bifurcation
buckling equations associated with the analysis of "symmetric" plates.
The bifurcation buckling equations for an unsymmetric plate are (see [14.5], [12.89] and
[14.10]):

(L 33
L.,
L
-
13

F)
] {u} {0}
v
w
= 0
0
(14.3)

where
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1056 Composite Structures

a4 a4 a4 a4 a4
L 11 =D 11 ax'
~
+ 4D 16 - " - + 2(D 12 + 2D 66 ) -2 -
ax'ay ax ay 2
+ 4D7- 6 - - , + D 77 -;----
axay' -- ay 4
a2 a2 a2
L~'
-
= L-. 1
"
=A 16 -
ax2
+ (A 1, + A66 )
-
--
ilxay
+ A26 -
ay 2

Ln = L-,t
·
= -Btt -axa', -3B,6 -.aray
a' a' a'
-,-- -(Bt2 + 2B66) -.- , - B26 - ,
axay ay
iJ 1
a' a' a'
Ln = L,2 = -B,6 -ax , -(B,2 + 2B66) -2-
ax ay
3 - B26 ~-
axay
B22 - ,
ay

(14.4)
For symmetrically laminated plates the (B;)'s bending-stretching terms vanish and Eq. (14.3)
becomes
a4 w a4 w a4 w iJ 4 w
D, 1 - 4 + 4D, 6 - 1 - + 2(D 17 + 2D"6 ) - 7- , + 4Do1 - -
ax iJx iJy - iJx-iJy- -' iJxay'
iJ 4 w a2 w iJ 2 w a2 w
+ D22 - 4
= Nx - + 2N,, - - + N, (14.5)
ay ax 2 . axay iJy-
7

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Note that Eq. (14.5) is the buckling equation as for a homogeneous, anisotropic plate. The
only difference is how the stitincss coefficients D,1 in this equation are calculated.
Equation (14.5) can be further simplified for symmetric cross-ply laminates, where the
fibers of the laminae are aligned with the in-plane axes of the plate and for which the
bending-twist coupling vanishes. Since in these plates D 26 = D 16 = 0, Eq. (14.5) becomes:

a2w a2w a2w


= N, - " + 2N, -.- + N,. -:--;- (14.6)
ax- · axay · ay-
This equation is of the same form as the buckling equation for a homogeneous orthotropic
plate.
Solutions of Eqs. (14.3), (14.5) and (14.6) that are of order eight have to comply with
four out-of-plane and in-plane boundary conditions applied at each edge of the plate. In
plates either simply supported or clamped along their edges, one out of the four following
combinations of boundary conditions has to be satisfied:
''S'' ''C''

Sl or Cl:
aw (14.7)
w = M" or - = u" = U1 = 0
an
iJw
S2 or C2: w = M, or -:- = N, = U1 = 0
i!n
iJw
S3 or C3: w = M" or - = u, = N"' = 0
iJn
aw
S4 or C4: w = M" or an = N" = N" 1 = 0
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1057

Here the subscripts n and t denote the coordinates normal and tangent to a boundary under
consideration; "S" and "C" stand for simply supported and clamped boundaries, respec-
tively. (For further discussion about the boundary conditions see [14.5], [12.89] and [14.10].)
Leissa in his excellent comprehensive reviews on buckling of laminated composite plates
([14.5] and [12.89]) cited and summarized the numerous studies associated with the efforts
to provide solutions to Eqs. (14.3), (14.5) and (14.6). He presented many solutions and results
for three laminate types, orthotropic, anisotropic and unsymmetric, subjected to various types
of loadings, unidirectional and combined, and prescribed combinations of boundary condi-
tions. In the following presentation, which is based on Leissa's monograph [14.5], buckling
solutions for various types of laminated plates are given. Also, shear deformation effects and
postbuckling behavior are discussed.

Buckling of orthotropic plates- As shown in [14.5] the solution of the buckling mode of a
(a x b) rectangular orthotropic plate, simply supported, (w = M, = 0) along its four edges
and subjected to uniaxial or biaxial loading (zero shear stress Tw = 0), is given by

Wmn =
. m7TX .
C 11111 Slll -a- Slll bnny (nz, n = 1, 2, 3 ...
) (14.8)

Substitution of Eq. (14.8) into the buckling equation (14.6) yields the critical value of the
buckling stress:

(1_)
(b
2
n:
m-
_ u,hb 2
(14.9)
rr2D22

where u, and u, are the stresses applied along the edges of the plate parallel to the longi-
tudinal x and transverse y axes, respectively and m and n represent the number of buckling
half-waves in the x and y directions, respectively.
Plots ofEq. (14.9) corresponding to uniaxial loading, u, = 0, are presented in Figure 14.1
for the nondimensional buckling stress parameter KJ 7T 2 versus the plate aspect ratio (a/ b),
for (D 12 + 2D 66 )/ D 22 = 1 and various other values of D 11 I D 22 • It appears from Figure 14.1
that for composite plates simply supported along all edges (type SSSS), which are stiffer in
the direction of loading, that is (D 11 > D 22 ), KJ 7T 2 is larger than 4 (isotropic plate);
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,

1.0 5.0
Aspect ratio alb

Figure 14.1 Uniaxial buckling parameter (u, = T" = 0) of a plate simply supported along its edges
with various stiffness ratios D,, I D 22 , for (D 12 + 2D66 )/ D 22 = 1 (from [14.5])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1058 Composite Structures

however the critical mode is associated with fewer longitudinal waves. Conversely, if the
fibers are primarily aligned perpendicular to the direction of loading, D 11 < Dw the buckling
load is lower and there are more axial buckles in the critical mode.
Curves obtained from Eq. (14.9) for biaxial loading are depicted in Figure 14.2 (hydro-
static loading, crJcr, = 1) and Figure 14.3 (tension-compression cr)cr, = -1, pure shear on
the 45o planes). It is seen from Figure 14.2 that the critical mode corresponding to hydrostatic
pressure loading is associated with one half-wave in each direction, immaterial of the plate
dimensions, and KJ 71' 2 has a minimum value approaching I. In the case of compression-
tension loading of Figure 14.3, more waves than for uniaxial loading are observed in the
direction of the compressive loading, and as anticipated, the buckling stress is higher due to
the stabilizing effect of the tension loading. Equation (14.9) can, of course, be also used for
determining the critical stress of other combinations of biaxial loading.

Buckling of anisotropic plates- Due to the existence of the bending-twisting coupling terms,
D 16 and D 26 in Eq. (12.5), there is no exact solution of the buckling equations for plates of
finite dimensions. Following Jones [14.12] Leissa pointed out in [14.5] and [12.89] that the
presence of even very small D 16 and D 26 terms (e.g. angle-ply plate with a large odd number
of alternating plies) may cause results significantly different than those obtained for zero D 16
and D 26 .
The results of applying a finite difference energy method to solve the buckling problem
of Eq. (14.5) in the case of angle-ply graphite-epoxy simply supported and clamped plates
subject to either compression or shear loading, obtained by Housner and Stein [14.13], are
shown in Figures. 14.4 and 14.5, respectively. It is apparent from Figure 14.4 that the critical
buckling stress has a maximum value for a ± (} laminate configuration of approximately
± 4SO. The magnitude of this maximum value is more than twice that of oo or 90° unidirec-
tional laminates. Also, a range of fiber orientation exists for which the buckling stress of the
graphite-epoxy plates exceeds that of corresponding to equal weight, aluminum plates having
the same dimensions and boundary conditions.
Similar conclusions can be drawn from Figure 14.5 for the shear loading case. In this
case, however, the angle of optimum fiber orientation exceeds 4SO as the plate aspect ratio
(a I b) becomes larger than unity.

Buckling of unsymmetric laminates - As shown above, the buckled mode of this type of
plate is associated with coupling between bending and mid-plane stretching.
Exact closed form solutions of Eqs. (14.3) are available only for two cases of uniform
biaxial loading (cr, and cr, constants, T,, '= 0): cross-ply plates with S2 boundary conditions
and angle-ply plates having S3-type edges. It was shown in [14.5] that for the unsymmetrical
cross-ply laminates with S2 boundary conditions exact solution of Eq. (14.3) takes the form:

mTrx nTry
u = A cos -a- sin b

mTrx nTry
v = B sin ~ cos b (14.10)

mTrx nTry
w = C ·;in
.
-X- sin

-b-

and for the antisymmetrical angle-ply laminates with S3 boundaries, the exact solution is
given by:
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1059

--~-

-:~-o·"·O-··
-a,
.
•E -a,,
/' "[}__2]_ -'- 1 (isotropic)
I --+----1
0
g_ 011
~
H-----1------+--~o/.o; = 1o-+----+-------1
~ I
s"'
~
[IJ
0
0 3.0 4.0 5.0
Aspect ratio alb

Figure 14.2 Hydrostatic stress buckling parameter (cr) a, = l) of a plate simply supported along
its edges with various stiffness ratios D 11 I D 22 , for (D 12 + 2D 66 )1 D 22 = l (from [14.5])

. m7TX n7Ty
u = A sm - - cos - -
a b
m7TX n7Ty
v = B cos - - sin - - (14.11)
a b
m7TX n7Ty
w = C sin -a- sin b
These solutions lead to the following equation for determination of the critical buckling
stress parameter:

c + 2Cl2cl3c23--- cllc:;3--- CnC73


(14.12)
" cllc22 - cT2

where for the cross-ply laminates the (C;;)'s are given by

5G
\1
43 \ 1\ 1\ 'I _,
v v2 \../"\...J A. .....
"'= 1 3 4 5 6 7
40
\E_u ~ 10
D,,
o' ay -

6''-'"'"""""'0
'" I
!!.JJ.=0.1 -ail{ -ax
-

-
l/D, -
\'A
a,
~--J""- ..-I .l
m-- 1 2 3 4 5 6 7 8
,, '~·
m=1 2 3 4 5 I I
1.0 2.0 3.0 4.0 5.0
Aspect ratio alb

Figure 14.3 Tension-compression buckling parameter (a) a, = ---1) of a plate simply supported along
its edges with various stiffness ratios D 11 I D, 2 , for (D 12 + 2D 66 )1 D 22 = 1 (from [14.5])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1060 Composite Structures

y'

Equal-weight
aluminum plates
alb o Simply supported
0 Clamped

Integer

All edges
5, ~ -·--Clamped
---Simply supported

0
oL---~1~5----~3~0-----4~5----~6~0----~7~5-----!90"
Fiber ori<!'ntatlon ±0, degrees

Figure 14.4 Compressive buckling parameter for graphite-epoxy angle-ply plates (from [14.13])

cl2 c21 = (Al! + Aoo)a{3 (14.13)

= B 11 at 3 + (B 12 + 2B66 )a{3 2
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Cu c,l

c2, c,2 = (B 1 ! + 2B 66 )a 2{3 + B-22 {33


with
m7T n7T
a=- {:.:
a b
and for the angle-ply plates they are:
ell -(A II a2 -- A66{32)

Cn -(A22{32 -- A""a2)

C, = D 1a 4 + 2(D 12 + 2D 66 )a2 {32 + D 22 {3 4

-(A 12 + A 6 c)a{3 (14.14)

C,l 3B (p2{3 + B2of3'

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1061

y T"Y

f--"------+---+-~£..-f->"~~1-''--"~-~ lib
~

"'
~ I

'"' ~~
I

~
-.;
E
!:!
8.
tl)
tl)

!;
tl)
Equal-weight
"'
.5:
:;;:
aluminum plates
u o Simply supported

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
"
.£l D Clamped

~
.c
Ill
0
0 15 30 90
Fiber orientation ±O,degrees

Figure 14.5 Shear buckling parameter for graphite-epoxy angle-ply plates (from [14.13])

m7T n7T
with a=- {3 = -
a b

and the coefficients A;r B;i and Du are defined by Eq. (14.2).
Plots of the uniaxial nondimensional buckling stress yielded by Eqs. (14.12) and either
Eq. (14.13) or (14.14) are shown in Figures 14.6 and 14.7. It is observed in these figures
that the effect of unsymmetry is emphasized in laminates composed of two plies; however,
it significantly decreases with increase in the number of plies in the laminate. The bending-
stretching coupling effect vanishes for a plate composed of an infinite number of layers,
N = cxo, and the plate behaves as if it were orthotropic and symmetric. It is observed in
Figure 14.7 that for angle-ply laminates (}is optimum at ±45°.

Effects of shear deformation - In analyzing laminated plates with classical laminated plate
theory, the Kirchhoff hypothesis is assumed. This assumption neglects the transverse shear
deformations, an effect that becomes important as the plate thickness increases relative to
the other dimensions of the plate. In this case the flexibility added to the plate, due to the
shear deformations, may be of more importance in laminated plates than in isotropic ones.
In Figure 14.8 the results of a study [14.14] on the influence of shear deformations on
buckling of a ( ± 4SO) angle-ply graphite-epoxy symmetrical laminate, uniaxially loaded and
having all edges simply supported, are compared with those corresponding to classical plate
theory. This figure depicts the considerable decrease in buckling stress due to shear defor-
mations in relatively thick plates, a/ h < 30. As pointed out in [14.5], this reduction in load
was observed in many investigations that were cited there.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1062 Composite Structures

---4
(a) Aspect ratio alb

:----------------6
t

X 0
ZIZ
I
--.2

(b)
'IODULUS RATIO. ~
2

Figure 14.6 Comparison of antisymmetrical and orthotropic solutions (from [14.5)): (a) with varying
aspect ratios (a/ b) for uniaxially loaded, cross-ply plates with SS3 type edge conditions,
(b) with varying moduli ratios (E, I E 2 ) for uniaxially loaded, cross-ply square plates with
SS3-type edge conditions

Postbuckling behavior- The previous subsections discuss the linear, bifurcation buckling
analysis used to determine the critical value of stress associated with the occurrence of
instability in a particular plate subjected to a given load and prescribed boundary conditions.
However, like metal plates, composite plates exhibit stable postbuckling behavior and are
capable of sustaining loads considerably in excess of their buckling capacity before collapse
or reaching their ultimate load (see Subsections 8.1.2 and 8.1.3 of Chapter 8).
Theoretical analysis of postbuckling behavior of plates is nonlinear. The nonlinearity stems
from the additional strains/ stresses induced by the transverse displacements. A comprehen-
sive theory to handle the large displacement behavior of all types of laminated plates is
furnished in the excellent book by Chia [14.15].
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Equation (14.3) can be generalized for postbuckling studies by adding the nonlinear terms,
due to stretching, to the strain-displacement relations. The postbuckling equations remain
coupled, even when a laminate is symmetric, because of the nonlinear terms, and a system
of eighth-order nonlinear differential equations has to be treated. Though these equations are
homogeneous, the boundary conditions are not. The solution methodology of the nonlinear
equations involves the prescription of either in-plane stresses or displacements along the
plate edges. Then the resulting equilibrium problem is solved with aid of the equations for
the lateral displacement in terms of the prescribed edge values.
Chia and Prabhakara [14.15] and [14.16] studied the postbuckling behavior of unidirec-
Copyright Wiley
tional orthotropic plates, simply supported along
Provided by IHS Markit under license with WILEY
their four edges, made of graphite, glass
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1063

80.---.---.----.---.------~

WiM
Orthotropic solution -a

70f---I---IB,1--6_=_s,-+6_=_ol*--f Tlb
'}o
!<!
.a
60 f----t------.rf----7 ~
0~

..!. 50f---+l+-k--t--l----'t-\ ~--~


~

~
E
d

"
~
~
~

"'c
""'u
"
.a
.,;
-~
-.,
:>

(a) Lammation angle 6, degrees

•o.---------------------,
N O'HHOTROPIC SOWTION

30
ru

E
2
g_
"0
~ 20

X
d
iii
E1=4o ii12~.s
.o -" E2 E2
z" I tlJ O L_~l:_:_:Nx:__o_:_;NLY:__,_;_:_cN,__,,y_~0:.:)'--~--------'
0 15 30 •s
(b) LAMINATION ANGLE 6

Figure 14.7 Comparison of antisymmetrical and orthotropic solutions (from [14.5]): (a) with varying
lamination angle for uniaxially loaded, angle-ply square plates with SS3-type edge con-
ditions, (b) with varying lamination angle for biaxially loaded angle-ply square plates with
SS3-type edge conditions

and boron fibers embedded in epoxy resin, and compared their behavior with that of an
isotropic plate. The variation of uJacr with increase in (w,lh) obtained in [14.15] and
[14.16], for uniaxially loaded square plates, is shown in Figure 14.9 (note that ux is the
applied stress, u" is the critical stress and w, is the deflection at the center of the plate). It
is apparent from Figure 14.9 that the isotropic plates have significantly higher postbuckling
carrying load capability. Furthermore, to reach a certain increase in postbuckling load, the
composite plates require greater deflections. These larger deflections may excite failure mech-
anisms inherent to composites (e.g. delamination) under relatively small additional stresses.
For further discussion, details and studies on postbuckling behavior, the interested reader
is referred to [14.5], [12.89] and [14.15].
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

b. Axial Compression Loading Experiments


The development of a technique for testing of fiberglass panels in edgewise compression
was presented by Hoff, Boley and Coan in 1948 [8.53]. The test equipment, test techniques,
evaluation of the test data and the test apparatus described there are discussed in Subsection
Copyright Wiley
Provided8.2.6 of Chapter
by IHS Markit 8 (for
under license with WILEY details on loading deviceLicensee=McDermott
see also Figure 8.45).
Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1064 Composite Structures

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
(a) Thickness ratio a/h

80

70 Elasticity
Classical lamination
~ plate theory
E so 0
Shear deformation
E' theory
g_

10

0.5 1.0 1.5 2.0 2.5 3.0


Aspect ratio
..':.
(b) b

Figure 14.8 Shear deformation effects (from [14.5]): (a) influence of plate thickness ratio on buckling
of a unixially loaded, ± 45 angle-ply square plate having an infinite number of layers,
simply supported along the edges, (b) comparison of exact elasticity, classical plate theory
(CPT) and shear deformation theory (SDT) solutions for uniaxially loaded cross-ply lam-
inates, simply supported along their edges

OL___~L----~--~~---7
0 1.0 2.0 3.0 4.0
Central deflection '«:lh
---T~~

Figure 14.9 Variation of postbuckling uniaxial stress with increase in central de!lection for square
plates made of ditierent materials simply supported along the edges (from [14.5])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1065

Other pioneering experimental investigations were reported by Davis and Zender [14.17],
Ashton and Love [14.18], Kicher and Mandell [14.19], Willey [14.20], Verchery [14.21]
and A. Chailleux, Y. Hans and Verchery [8.82]. In [14.17], Davis and Zender determined
experimentally the buckling and postbuckling behavior of 12-ply (0/60/120b and 18-ply
(0/60/ 120) 1 s quasi-isotropic glass filament- reinforced plastic plates. Details of the test spec-
imens are shown in Figure 14.1 0. The specimens were loaded in the loading frame depicted
in Figure 14.11, with the unloaded edges being simply supported as displayed in Figure
14.12. The unloaded edge fixtures were made of 17-4PH stainless steel heat-treated to H900
condition. They were adjustable to accommodate plates of various thickness. In order to
avoid penetration of the knife edges of the fixtures into the relatively soft epoxy, a 0.04-cm
radius was provided on the contact edges of the edge fixtures (see Figure 14.12). Further-
more, to ensure that the load introduced by the testing machine was solely supported by the
plate, except for negligible friction loads between the surface of the plate and edge fixtures,
the fixtures were made 9.5 mm shorter than the test plates.
Compressive load was continuously applied until failure at a deformation rate of 21
f.Lm/s. Shortening of the plate was measured by a DCDT (Direct Current Differential Trans-
former) (see Figure 14.11) and lateral displacement at five vertical locations midway between
the edge supports by another five DCDT's (see Figure 14.13).
Comparisons of the test results with theoretical predictions of buckling loads, corresponding
to all of the 15 plates tested in the program, and postbuckling responses corresponding to
plates designated 8 and 9 in the test program, are shown in Figures 14.14 and 14.15, respec-
tively. Very good agreement between theory and experiments is observed in these figures.

The experimental program of Ashton and Love [14.18] aimed at determining the bucking
loads of boron-epoxy panels and comparing the test results with those predicted by the
analysis of Ashton and Waddoups [14.22]. Twenty plates fabricated from Narmco 5505
prepregged tape bonded together with Shell Epon 828 (60 parts by wt) and General Mills
Versamid 1150 (40 parts by wt) were employed in the test program. The plates were 11 in.
(279.4 mm) square and of various laminate configurations and thicknesses (see Tables 1 and
2 in [14.18]). Prior to testing, the thickness of each plate was measured at nine predetermined
points and the results were then averaged to yield the thicknesses presented in Table 2. In
the tests the panels were installed in a square frame that clamped their loaded edges and

25.4

6.35

(all dimensions In mm)


--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 14.10 Davis and Zender's test specimens [14.17])


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1066 Composite Structures

1'!-.::
~
.
t:.U:'"":J
...::: I

Figure 14.11 Plate tested by Davis and Zender installed in Lhe test machine (from [14.17])

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
mai ntai ned uniform displacement while the load was increased. The unloaded edges of the
plates were either completely fixed (Figure 14.16a) or simply supported by knife-edge sup-
pons (Figure 14.16b). After the edges of the panels were secured, their actual inside testing
dimensions measured 10 x 10 in. ' (254 x 254 mm' ). In preparation for testing and in order
to reduce possible introduction of lateral shear forces between the plate loaded edges and
load bead, as well as tension and compression reactions in the edge supports and shear-load
transfer to the side supports. each panel was taped on all four edges with two layers of
Scotch 549 Tenon tape.
The test frame was mounted in a 200 kips (890 kN) test machine with its upper portion
attached to the middle head of the machine and the lower portion resting on the load table.
Parallelism between the heads of the machine was maintained between ± 0.003 in. (.076
mm) over the 11-in. (279.4-mm) test length by using shim stock.

Figure 14.12 Edge fixtures used in the loading rig of Davis and Zender (from 114. 17))
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1067

Figure 14.13 lnstrumenlation used tO measure lateral dellection of plates in the tests of Davis and
Zender (from [14.171)

The following procedures were followed in conducting the clamped-clamped tests:


I. The panel was placed in the fixture and all the pressure bolts forcing the four clamping
bars against the panel were tightened with a wrench.
2. The assembly was lifted from the table using the load head to allow the bottom to swing
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

free, thus aligning the head and base when lowering the assembly back to the table.
3. The rack of transformers, three linear displacement transducers (LOT's, Sanbord Model
5850T-I 00) spaced evenly along the vertical centerline of the panel to record its lateral
deflections, was put in place. The core of each LOT was bonded to urethane tabs bonded

<
.§ .010
;;
d
.v
~ .008

! .006

.oo•

.002

panf( wlc!th
( poneI th1<:k l'\lf'5.S
)2 ( b
=7
)2
Figure 14.14 Comparison of buckling data obtained in the tests of Davis and Zender (from [ 14. 17])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1068 Composite Structures

tnd shoftming

Figure 14.15 Comparison of experimental suess-unit-shonening curves obsef\·ed in the tests of Davis
and Zender and theoretically predicted SltC,~·unit-shoneoiog CUf\'CS (from [ 14.17))

to the back surface of the panel and the transfonncrs were adjusted to ind icate zero
displacement.
The panel was then loaded in 500 and 1.000 lb (2.2 and 4.5 kN} increments unti l ap-
proximately 0.05 in. ( 1.27 mm} de flection was achieved. The panel was now un loaded to
assure zero shift of tbe LOT. Then the LOT was readjusted to zero and the procedure was
repeated until tbe system returned to zero satisfactorily and the panel was ready for test. To
asccnain compatibility with small-deflection theory. the maximum deflection in each test was
restricted to less than 0.06 in. ( 1.52 mm). Each test configuration underwent at least two
data runs and then a clamped-simple support test was perfomted. To obtain this test condi-
tion. the same procedures were followed, except that the side clamps were rotated. thus
obtaining the conditions presented by Figure 14. 16b. The side suppon tighteni ng bolts were

(O) (b)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 14.16 Top view of compression frame employed in the tests of Ashton and Lo•c (from ( 14.18():
Copyright Wiley (a) showing panel in clamped edge suppon.
Provided by IHS Markit under license with WILEY
(b) showing panel in knife edge suppon
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1069

left only hand tight in this position. After completion of the clamped-simple test, each panel
was rotated 90° and the tests were repeated. A set of four buckling curves was thus obtained
for each panel.
Some of the panels were characterized by very high Poisson ratios leading to in-plane
normal stress-shear strain coupling. Uniformity of load introduction was therefore in doubt.
To evaluate the load input, 12 strain rosettes (6 back-to-back pairs) were bonded on some
of these plates, and they were tested under both clamped-clamped (up to 10,000 lb or 45
kN) and clamped-simple conditions (up to 8,000 lb or 36 kN). An acceptable uniaxial stress
condition uniformly distributed over the panel was observed.
The Southwell method was applied to determine the critical buckling load. A typical plot
reproduced from [14.18] is depicted in Figure 14.17. The critical buckling loads obtained in
the tests were compared with those predicted by anisotropic analysis. For the symmetrically
tested laminates the experimental results verified the analytical ones. It was also demonstrated
that unsymmetrical laminated panels can be adequately analyzed using the uncoupled anal-
ysis with the reduced bending stiffness matrix of [14.23].

Kicher and Mandell [14.19] conducted experiments with approximately square glass fiber
and graphite fiber-reinforced plates subjected to uniformly distributed compression edge load-
ing, in order to evaluate the applicability of classical buckling theories and the effects of
membrane-bending coupling. Two types of loading were examined: loaded edges simply
supported and unloaded edges either free or simply supported. Wedge-shaped bearing points
were attached to the loaded edges to achieve simple supports and avoid crushing of the edges
of the brittle specimens. This added about 0.5 in. (12.7 mm) to the length of the 10 X 10
in. 2 (254 X 254 mm 2 ) specimen and introduced approximately a 1 percent error in the
calculated buckling load. On the unloaded edges the simply supported boundary condition
was achieved by confining the displacement of the plates between rollers, thus allowing the
unloaded edges to freely expand in the plane of the plate (see Figure 14.18a-c).
The test fixture, Figure 14.18a-c, is peculiar in that it introduces a uniform load rather
than the common uniform displacement. Uniform distribution of the load was maintained by
employing a soft loading head. Thus, nonuniform distributions in the end displacement did
not induce a nonuniform stress into the plate. To achieve this type of loading, each of the
bearing wedges mentioned above was fitted into a V-grooved spring supported piston, the

8 ~Panel No.: 8
Configuration: 0°
J 16

Test Condition :Clamped-Simple


1

'f'~ 6 / 1

,L_
X
"C

~
a..

/
-9: /I...-f...--- f.----
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

-c' 3 / -
.3"
c
r--
j/
-~ 2

l1'
u
104
"' P= 5 · 00 x •8600 1bs
't
0
cr 7.00-1.12

0
1/ 0
0 100 200 300 400 500 600
Deflection,o, Inches x 10" 4

Figure 14.17
Copyright Wiley
Typical Southwell plot and load deflection curve obtained in the tests of Ashton and
Love (from [ 14.18])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1070 Composite Structures

(a) (b)

(c)

Figure 14.18 Axial compression test fixture used in the tests of Kicher and Mandell (from [14.19])

spring of which was selected to match the axial stiffness of the plate. During load application
the springs displaced relative to each other and an approximately uniform load distribution
was achieved. Rigid body motion, necessitating the floating of the vertical side supports, was
introduced by the deformation of the lower set of springs. The simply supported boundary
conditions on the unloaded edges were provided by hardened steel rollers mounted in cages
that slid in fixed brass side rails. A cylindrically shaped load cell was employed to determine
the total load supported by the plate.
Lateral deflections of the loaded plate were measured by a micrometer that made an
electric contact with the plate. Maximum lateral deflections were measured to obtain the
load-deflection curve, and deflections at a number of additional points were recorded to
define the mode of the deflected plate surface. To maintain contact at the point of maximum
displacement during loading, the displacement transducer was accurately shifted vertically.
The Southwell plot technique was employed to determine the critical load from the load-
deflection curves (see Figure 14.19). It was pointed out in [14.19] that good correlation was
observed between the experimental buckling loads and the predicted ones, provided the
appropriate analytical tools were employed.
Willey [14.20], Verchery [14.21] and Chailleux, Hans and Verchery [8.82] indicated that
their investigations were aimed at obtaining experimental techniques for determining the
critical loads of composite plates. The program of Willey [14.20] focused on methods for
interpretation of the test data. This goal immediately raised the question of what data should
be acquired and what method should be employed to analyze these data. To answer this
question, Willey evaluated the strain reversal technique (see Subsection 8.3.1 of Chapter 8)
versus the Donnell large-deflection approach [4.26], for which the critica1load is given by
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Eqs. (8.13) (see Subsection 8.3.2 of Chapter 8).
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1071

OL-~~~~--~~-J
0 400 BOO 1200
Deflection (t,) 110_, in I

Figure 14.19 Southwell plot and load deflection curve observed in the tests of Kicher and Mandell for
a (0/90/90/0) Thornell-25 plate, with four sides simply supported (from [14.19])

2
W [ 3(1 - u )
P = Wp' 1+ (W + 2W0 )(W + W0 ) ] (8.13)
Wo + 812

The Southwell expression for the critical load [4.12] given by Eq. (4.11) (see Subsection
4.5 .1 of Chapter 4) is

W = P, W- W: (4.11)
p ()

If the requirement is imposed that W0 /W be small, Eq. (8.13) becomes


P = P,(l + yW 2 ) (14.15)
3(1 - v 2 )
where y=

Since y is positive and W is real, this equation implies that it applies only to P > P,, i.e.
to the postbuckling range of loading.
To achieve the objectives of the program, 20 square fiberglass panels of various lamination
angles were tested under axial compression with their loaded edges fixed and unloaded sides
simply supported. Once the suitability of the Southwell method was established, the exper-
iments were performed to study the effects of changing boundary conditions and initial
geometrical imperfections, as well as to verify the large-displacement technique of Eq.
(14.15) as a method for determining the critical load.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Typical results, depicting the three methods are given in Figure 14.20 reproduced from
[14.201. On the basis of the experimental data reduction it was concluded in [14.20] that the
strain reversal technique is meaningless in the interpretation of the test data. On the other
hand, both the Southwell method and large-displacement technique (P versus 82 ) provided
identical results that could be associated with theory. Furthermore, the large displacement
technique appeared to be preferable because of the ease of its application and the more
positive definition of the straight lines obtained from the reduced data.
The investigations of Verchery [ 14.21] and Chailleux, Hans and Verchery [8.82] aimed at
applying the Southwell method and the dynamical approach, based on the change of trans-
verse natural frequencies with change in applied load, for determining the critical load of
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1072 Composite Structures

1< .002" CLEARANCE


400
2< .004
3' .006

300

@
a..
'0
200
]
100

(a)
10 14
Deflection vs. Locd and Strain Reversal vs.Load.

100 200 6' 400


300

250

o9
a..
-u 200
]

150

(b)
8 6 12 16 20
Southwe!l and P vs. 6 2 Plots.

Figure 14.20 Willey's large deflection test results at various side support clearances (from [14.20]):
(a) deflection versus load and strain reversal load for determination of critical load, (b)
Southwell and large deflection plots for determination of critical load

square plates (see Subsection 15.2.3 of Chapter 15). It was concluded in [8.82] that with
certain restrictions the Southwell plot using the transverse deflection can be successfully
applied for determining the critical load. The dynamical method, however, must be used
with caution since a linear relation between the applied load and the squared eigenfrequency
is not always observed. A criterion of the validity of this approach, based on the range of
the linear zone, was given.
Banks and Harvey [14.24] and Marshall and Banks [14.25] studied the buckling and in
particular the postbuckling behavior of reinforced plastic panels subjected to unidirectional
in-plane loading. The experimental program of [14.24] aimed at validating the theoretical
studies of [14.26] and [14.27]. To achieve this goal, glass fiber orthotropic plates, using
unidirectional E-type Tyglas embedded in a Crystic 272 polyester resin (Scott Bader Co.
Ltd.), were fabricated. The plates were tested in the rig shown in Figure 14.21, that was
designed and built by Fok [14.28]. This rig was similar in many aspects to that of Hoff et
al. [8.53]. To permit testing of plates with various aspect ratios, alterations were introduced.
The rig was designed to provide a uniform displacement along the loaded edges, stress-free
unloaded edges and simply supported flexural boundary conditions along all edges (see
details of Figure 14.21). Plates 250 mm wide and of various lengths, up to a maximum
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

aspect ratio of 2, could be accommodated in the loading rig.


The loading rig consisted of a lower base, a removable upper platen and two identical
channel side supports, which were bolted lo the lower base. The upper platen was constrained
to move vertically during loading of the compressed panel by rigidly attached guides. Load
was introduced into the specimen through a series of needle bearings mounted in steel blocks,
which were guided into the platens. Slotted holes in the webs of the side supports allowed
adjustments of the side supports knife edges to accommodate variations in plate thickness.
The knife edges provided line contact along the unloaded length of the plate. They were
lubricated during loading to prevent restraint on the vertical displacement of the panel and
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1073

tui'PO!ffS I'OR l"un


LOI'ID&D &OQI.I

Figure 14.21 Banks and Harvey's compression test rig for laminated plates (from [14.24])

permit free rotation about the edges, thus ensuring that the total applied load would be
introduced into the plate.
A 0.9 MN Tinius Olsen tension-compression testing machine was used to load the panels.
Displacement-controlled loading was applied and was measured through the base of the
machine by load transducers. The response of the panel was monitored by pairs of strain
gages. two 90° rosette gages bonded face to face along the horizontal centerline of the panels,
and by dial gages that were used to measure the center out of plane deflection of the plate
and its in-plane end shortening.
To initiate a test, the top platen of the loading rig was removed and the panel was posi-
tioned in place by sliding it until its bottom rested on the slotted rods in the base. Then the
side knife edges were positioned against the panel to contact the thickest points along the
unloaded edges, and tentatively clamped. After it was confirmed that the plate was not rigidly
supported (in which case the axial displacements along the unloaded edges of the plate would
become restrained and some of the load would have been lost in friction). the knife edges
were fully tightened and the upper platen fixed in position. Following this procedure, the
panel was subjected to a small load of about 0.1 of its buckling load. This procedure was
used to settle the plate into the loading rig and eliminate any relative movements between
components of the rig. Then the load was removed, the strain gages and dial gages were
zeroed and the test of the panel was performed.
The experimental results, described by large deflection behavior and stress distribution,
were compared with the theoretical predictions of [14.26] and [14.27]. Some typical results
that were obtained in the test program of [14.24] are depicted in Figure 14.22a-c. Based on
the comparison between experiment and theory. it can be concluded that the test rig repro-
duced the nominal assumed boundary conditions. In general, the test results confirmed, to a
reasonable degree, the analytical predictions up to about three times the buckling load.
In [14.25] an attempt was made to highlight the influence of geometrical imperfections
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

on the buckling and postbuckling characteristics of composite structural elements. It was


demonstrated that presence of these imperfections may significantly reduce the load-carrying
capacity of composite elements.

The TELAC (Technical Laboratory for Advanced Composites, MIT) experiments- Extensive
theoretical and exceptionally well-documented experimental studies on buckling and post-
buckling behavior of fiat, symmetrically and unsymmetrically laminated and sandwich com-
posite panels were conducted at TELAC (Technology Laboratory for Advanced Composites),
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1074 Composite Structures

~
o_u
c::-
0
9 2f----+----t-----/
...J
<(
~
1-
<I:

~
<(
9
(a)
3 4 s
END DISPLACEMENJ/
..U/.Ucr
CRITICAL END DISPLACEMENT

3.---------.---------,r--------.
/(

~" f - - - - - - - 1 - - - - - -
9

1 2 3
OUT OF PLANE DEFLECTION/PLATE THICKNESS -WJ-i\_

0
~
...J
_,2 X <r;;' EXPERIMEHTRL
<(
u
;:::
a:1
~
<(
0
...J
(c)
10
STRESS IN Ml-l/rn 2

Figure 14.22 Typical comparison of Banks and Harvey's tests results with theoretical prediction curves
for a plate with aspect ratio equal to I (from [14.24]): (a) load-end displacement, (b)
load-out-of-plane deflection curves, (c) variation of longitudinal membrane and bending
stresses at center of plate

Department of Aeronautics and Astronautics, Massachusetts Institute of Technology (see


Finch, [14.7]; Jensen [14.9]; Jensen and Lagace [14.29]; Lagace, Jensen and Finch [14.30];
Minguet [ 14.31]; and Minguet, Dugundji and Lagace [ 14.32]). All of the experimental studies
included in these investigations were based on the TELAC loading jig developed by Finch
[14.7] and Lagace [14.33].
This buckling fixture is depicted in Figure 14.23. It was designed to meet as closely as
possible the requirements of applying the correct boundary conditions and introduction of
an even uniaxial compressive load along opposite edges of the panel, allowing specimens of
different thicknesses to be tested, and to be compatible with the hydraulic grips of a 445 kN
MTS 810 material test system. The jig was designed to sustain a buckling load of 245 kN.
Since the actual test loads were below 9 KN, the jig operated well within its design limit.
Since the top head of the MTS machine locks in place while the lower head displaces,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

the loading jig was of a free-floating type, gripped in the top head while being uniformly
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1075

-
~

c--GRIPPING TAB

::~~0 0 o~t~
_._
.
·~ . tJ
-254--
or,'

6 81 M
:a: 254
. t-~:·
H
. . . . l. I-
l~r--- GUIDE SHAFT

. . .
' ' '

''
' !II ,.
.l.§t .El.
.:
-
-. . li '
I'
LOAD ADJUSTMENT

' II'
lf I
0 0 0 : :
!


L:3 3~ L25.4
2 ~
~

25.4J

Figure 14.23 TELAC compressive loading jig (from [14.9]).

loaded due to the displacement of the bottom head. The load was introduced into the jig
through the top and bottom loading tabs of the jig (Figure 14.23) and was distributed across
102 X 102 mm 2 steel 1-beams. The lower I-beam was free to slide on four vertical 25.4-mm
diameter steel rod guide shafts. All sliding parts were well lubricated with molybdenum-
sulfur. The rig itself weighed 756 N. This weight was subtracted from the load reading prior
to recording the applied loads.
An important feature of the TELAC jig was the set of five adjustment 3/4 in. bolts, which
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

were located between the lower 1-beam and a secondary intermediate load distribution mem-
ber (Figure 14.23). These bolts allowed minor readjustments for obtaining an even load
distribution across the width of the tested panel. The fixture was designed to provide three
combinations of boundary conditions along the edges of the panels. whose loaded edges
were always nominally clamped, whereas the unloaded ones could be free, simply supported
or clamped. The clamped end boundary conditions were attached to the frame in the manner
shown in Figure 14.24. The clamp is made of two 16-mm steel bars separated by a 1-mm
steel spacer. The spacer was 13-mm shorter than the two bars, creating a 13 mm deep
channel. The loaded edge of the plate was inserted into this channel and then clamped by
the row of seven bolts (Figure 14.23).
The true clamped boundary conditions could be violated due to nonzero slopes induced
at the loaded edges of the panel by the imperfect construction of the clamp fixture of Figure
14.24. These misalignments were corrected by the slope-alignment bolts indicated in the
figure.
Clamped, simply supported and free boundary conditions could be applied by the side
supports of the loading jig. The fixtures producing the first two of these are presented in
Figure 14.25. The side supports consisted of steel channels into which the clamped and
simply supported boundaries were bolted into place. For interchangeability purposes, the
side supports were in modular form so that they could be easily removed from the jig,
interchanged and adapted to test plates of different widths and thicknesses.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1076 Composite Structures

CLAMPING BOLT

SLOPE
ALIGNMENT
BOLTS

CLAMP BARS "I'' BEAM

Figure 14.24 End clamps simulating clamped end boundary conditions in the TELAC loading appa-
ratus (from [14.9])

It is apparent from Figure 14.25 that the clamped side condition was produced in a similar
manner to that of the loaded edge, except that the sides of the plate were allowed to undergo
in-plane Poisson expansion perpendicular to the edge. In order not to restrict the in-plane
freedom, the pincer bolts in Figure 14.25 were only hand tightened. Likewise, the simply
supported boundary conditions in Figure 14.25 provided in-plane freedom because the pincer
bolts were also only hand tightened. The rounded knife edges (radius of 1.6 mm) provided
zero lateral deflection and zero applied moment and prevented stress concentrations along
the contact line.
[14.18] and [14.34] demonstrated the use of Teflon strips or tapes as a means to alleviate
the problem of extraneous shear load transfer to the specimen. Ashton and Love [14.18],
indicated that the presence of induced secondary stresses may reduce the buckling capacity
by one-third. Therefore, to minimize the etiects of these secondary stresses, two layers of
Teflon, 0.13 mm thick and 13 mm wide, were applied to the panel borders and were lubri-
cated by the same molybdenum-lithium compound used in the loading jig. This served as a
shear-transfer buffer, prevented the corner edges of the clamp from penetrating into the
specimen, and together with the hand tightening provided the "loosely clamped condition"
(see Chia [14.15]).
A problem may arise from employing rounded knife edges rather than sharp ones. This
problem, shown in Figure 14.26b, is associated with the moment that is exerted on the edges
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

STEEL CHANNEL

Figure 14.25 End clamps simulating simply supported side boundary conditions in the TELAC loading
Copyright Wiley
Provided by IHS Markit under license with WILEY
apparatus (from [14.9]) Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1077

Figure 14.26 Geometric difficulties presented by rounded knife edges (from [14.7])

when the angular deflection becomes large. In the experiments of the TELAC group, how-
ever, the angular deflections were so small that any induced moments could be neglected
and therefore the rounded edges did not present any geometric difficulties (Figure 14.26a).
The objectives of the experimental programs of [14.7], [14.9] and [14.29]-[14.33] were
twofold. The first goal was the determination of the buckling load by application of the
Southwell method. The second experimental objective focused on mapping the deflected
surface of the panel so as to compare the experimentally observed mode shapes with those
predicted by the theoretical buckling analysis developed in these investigations.
Application of the Southwell method requires high precision in the measurement of very
small displacements, on the order of half the panel thickness (0.50-0.75 mm). This require-
ment arises because the applicability of the method is limited to plate deflections not greater
than about half the plate thickness [4.40]. To obtain as many data points as possible within
the allowable deflection region, the primary requirement of the system is that the measuring
device must be characterized by high precision and fine resolution. For comparison of the
measured mode shapes with the analytical ones, enough measured points are needed to
represent adequately the deflected shape of the entire plate surface.
To meet the above objectives two back-to-back strain gages were bonded at a distance of
13 mm on either side of the panel vertical centerline (see Figure 14.27). Their strain readings
were used to display the bifurcation buckling behavior and determine an effective longitu-
dinal elastic modulus for estimating the local longitudinal stress. Also, because of the sym-
metry of the test and antisymmetry of the fibers about the loading axis, the two pairs of
gages provided an indication of the proper load introduction and panel alignment in the
fixture.
A ± 25 mm Trans-Tek model 354 LVDT was mounted on the side of the buckling jig to
supplement the strain measurements and provide a better measurement of the average relative
end-shortening of the plate as a function of the applied load. Measuring this displacement
locally, rather than using the testing machine head displacement, eliminated the inaccuracies
due to the slippage in the load grip and "play" in the heads and test fixture (see Figure
14.28).
Two types of lateral displacement were measured in the tests. One was associated with
scanning of the deflected shape of the plate at predetermined load levels, and the other with
continuous monitoring of the deflection under increasing load at predetermined fixed points
on the panel. Scanning was performed with the deflection tracker of Figure 14.29 [14.7] and
[14.9]), which was mounted at the front of the loading rig. This tracker was in essence a
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyrighttwo-degree-of-freedom
Wiley scanning system resembling a conventional X-Y plotter. The contin-
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1078 Composite Structures

• Strain Gages
(front and back )

o Thickness Measurements
and Transducer Locations
--,.-- y

0 0

1,2 3,4

63 5 mm
- +
i-127mm
254 mm

'-o
'i
63.5 mm

_;-
25.4 mm

Figure 14.27 Strain gage, thickness measurement and transducer locations in the TELAC test speci-
mens (from [14.9])

uous local displacements measurements were executed with the transducer rack and cart
assembly shown in Figure 14.28 [14.9] which was mounted at the back of the rig.
The deflection tracker employed a ±25 mm Sanborn type 7DCDT-1000 transducer that
was mounted on the scanning device. The DCDT was held by a Derlin plastic block and
the assembly could slide freely along two parallel vertical runners and was connected by a
cord-pulley system to a rotary potentiometer, the voltage output of which was used to record
the vertical position of the tracker on the plate surface (the spacing of measured discrete
data points was preset before each test). This scanning unit could be positioned transversely
anywhere along the plate width by sliding it within a pair of guide channels. In order to
obtain consistent data, nine clamping positions were built into the horizontal guides, 25.4
mm apart (Figure 14.29).

Figure 14.28 Transducer rack and cart assembly of the TELAC buckling jig (from [14.9])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1079

+ DEFLECTION MEASUREMENT LOCATIONS

/LAMINATE
r--- ----l
I + + + + + + + I
I + + + + +- + + I
I + + + + + + + I
I + + + +- +- + + I
I + + + + + + + I
457mm I + + + + + + + I
I + + + + ++ + I
I
+ + +- + + + + I
I + +
L_ _ __ ___,I

(a) (b)
Figure 14.29 Out-of-plane deflection tracker of the TELAC buckling jib (from [14.9])

Deflection measurements were sampled by a computer at the 81 points of Figure 14.29.


A typical output of a X-Y plotter showing a set of traces along the centerline of a clamped-
end, free-side specimen at monotonically increasing loads, is depicted in Figure 14.30. The
data obtained by the tracking system were used to generate isodeflection contour maps, which
were compared with the theoretical predictions (Figure 14.31 ).
The deflection tracker was also used to position appropriately and align completely the
boundary conditions. Corner readings and vertical profiles of the two extreme horizontal
mapping stations of the plate (Figure 14.29) taken with the scanner provided the information
necessary to correctly align the translational out-of-plane degree of freedom and the slope
of the boundaries. The four comers of the plate were aligned to within ± 0.04 mm relative
to a plane parallel to the deflection tracker, and the slope of the boundary conditions was
adjusted with the aid of the alignment bolts (Figure 14.24) to within ± 0.2°.
The modular rack and cart assembly held six ± 25 mm Trans-Tek DC-DC Model 354
gaging transducers (three in the upper row and three in the lower row, see Figure 14.28) and
three ±51 mm Trans-Tek model 355 gaging transducers. The transducers were spaced 63.5
mm apart in a form presented in Figure 14.27.

A y

w.
w

X N,
I

Figure 14.30 Typical out-of-plane deflection scans obtained by the deflection tracker of Figure 14.29
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(from [ 14.7])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1080 Composite Structures

Experimental Rayleigh- Ritz


[0/i90/i90/f03 ]T- 2 [03 //903 //90 3 it03]T
Clamped Sides Clamped Sides
5.3 kN 9.0 kN

Figure 14.31 Typical comparison between experimental (left) and predicted (right) isodeflection con-
tour maps-(0,/90,/90)0,) 1 laminate with clamped side boundary conditions (from
[14.9])

The test setup enabled automatic computerized data acquisition of 16 channels of infor-
mation: data from the nine deflection transducers in the rack, an end-shortening displacement,
four strain gages and the load cell. A switch allowed data to be taken from the transducer
and potentiometer mounted on the deflection tracker. During the tests, signals were simul-
taneously sent to a computer and an X-Y plotter or a voltmeter for on-line monitoring.
Preliminary tests were conducted to verify the test equipment and validate and refine the
proposed testing procedures. It was realized from these tests that many tedious steps were
involved in the process of calibrating the scanner (Figure 14.29), transducers and loading
jig, installing and aligning the plate, and performing the multiphase test for each specimen.
This led to standardizing the test procedures in the form of a checklist. Since it represents
a good example of modern and consistent procedure, this check list is now reproduced
verbatim from [14.9].

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
A) Prepare Plate for Installation
1) Attach teflon tape to all borders of laminate (two layers).
2) Lubricate teflon tape (all surfaces).
B) Prepare Jig and Install Plate
1) Install probe.
2) Zero load reading.
3) Insert plate loosely in upper and lower boundary conditions.
C) Attach Upper and Lower Boundary Conditions to Plate
1) Load to approximately 890 N (Set Point).
2) Hand tighten upper and lower boundary conditions (horizontal bolts).
3) Loosely tighten with wrench (seat bolt + 11 16 turn).
4) Return stroke to zero (Set Point).
D) Position Upper and Lower Boundary Conditions with Respect to the Frame
1) Read displacement (computer units and volts) and potentiometer readings (com-
puter units) at four corners. Adjust (if necessary) and repeat corner readings.
2) Record scanner offset (average of four corner readings).
3) Verify center transducer (No. 5) signal (first time only).
4) Clamp upper and lower boundary conditions to frame (vertical bolts).
E) Align and Fix Side Boundary Conditions
1) Align side boundary conditions (simultaneously hand tighten front and back).
2) Clamp side boundary conditions to frame.
F) Update Database Information File ("GLOBAL")
G) Balance and Calibrate Strain Gages
1) Attach strain gage wires.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1081

2) Adjust bridge excitation voltage (1 volt) on conditioner.


3) Adjust amplifier balance (with excitation off).
4) Turn excitation voltage on.
5) Balance all four strain gages.
6) Run calibration program ("ADCAL").
7) Calibrate gain for each gage (13746 ohms shunt equals 2000 computer units).
8) Record stain gage offsets (computer units).
H) Adjust Load Introduction
I) Load to approximately 890 N to seat plate.
2) Return stroke to zero.
3) Align load introduction bolts (hand tighten).
4) Release load to small compressive value (Set Point).
I) Determine Ojj:vets
I) Position transducer rack against plate.
2) Switch electronics to "Rack".
3) Record offsets for remaining channels ("ADCAL").
J) Scan Plate with No Load
1) Switch electronics to "Scan".
2) Run scanner data acquisition program ("SCAN").
3) Switch electronics to "Rack".
K) Initialize Data Acquisition
1) Run data acquisition program ("TDAK").
2) Enter initial data offsets.
3) Set interval between data points to 0.25 seconds.
4) Record initial machine stroke.
L) Load to Stability "Limit Load"
1) Reset memory.
2) Start data acquisition.
3) Increase stroke at rate of 0.42 millimeters/minute.
4) Hold stroke when center transducer indicates a displacement equal to half of the
plate thickness.
5) Halt data acquisition and close files.
6) Record maximum load and stroke.
M) Scan Plate at "Limit Load"
I) Switch electronics to "Scan".
2) Run scanning data acquisition program ("SCAN").
3) Return machine stroke to zero.
4) Switch electronics to "Rack".
N) Reinitialize Data Acquisition
I) Run data acquisition program ("TDAK" ).
2) Enter initial data offsets.
3) Set interval between data points to 0.4 seconds.
0) Repeat Steps L, M and N for Scans at Higher Loads
P) Load to Failure
1) Reset memory.
2) Start data acquisition.
3) Increase stroke at rate of 0.42 millimeters/minute.
4) Record marks when audible damage occurs.
5) Hold stroke when load drops significantly.
6) Halt data acquisition and close files.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
7) Record maximum load and stroke.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1082 Composite Structures

Q) Photograph and Remove Plate


1) Withdraw rack.
2) Attach label and photograph plate.
3) Return stoke to zero.
4) Lower crosshead and turn pump off.
5) Remove plate from jig.

It was recognized that following such a checklist would ensure that all of the steps were
accomplished in a consistent and proper sequence. Also, prior to testing a summary sheet
of critical information for the test was prepared. Such data sheets facilitated collection of
data offsets and other essential parameters and enhanced the testing procedures. Note that a
similar procedure may also be useful in other tests.
Following the preliminary tests, the entire test setup was realigned and calibrated in prep-
aration for full-scale testing. Prior to each testing session, the loading frame was carefully
installed in the MTS machine. Alignment of the rig was ensured by lining up the steel tabs
with one side of the hydraulic tabs of the MTS machine (see Figure 14.23).
After the rig was installed in the MTS machine, the clamping upper and lower boundary
conditions were correctly positioned and aligned in the rig with the aid of an aluminum plate
that was placed in it. Next the side boundary conditions were adjusted. The aluminum plate
was loaded and all of the data acquisition equipment and software were operationally verified
and calibrated. Following the correct positioning and complete alignment of the boundary
conditions, the aluminum plate was removed from the fixture by loosening only the front
halves of the boundary conditions from the rig. The rear halves of the boundaries remained
fixed and served as a reference plane for alignment of the graphite-epoxy panels.
Before the composite plates were inserted into the loading fixture, the two layers of 0.14-
mm thick by 13-mm wide CHR-T Temp-R-Tape Teflon tape were bonded to the borders on
the plate. To minimize friction along the borders, which would introduce shear loading into
the specimen, the borders were also lubricated with Lubriplate molybdenum-lithium No. 2
multipurpose lubricant. Then the plate was inserted loosely in the upper and lower bound-
aries. A load of approximately 0.9 kN was applied to the plate to properly seat it in the
boundaries. The upper and lower boundary conditions were hand tightened and the load was
removed.
In this position the corner readings of the plate were taken with the deflection scanner
and the position of the boundary conditions was adjusted as necessary. The average of the
four corner measurements provided an initial offset of the position of the plate relative to
the zero position of the scanning transducer. This offset was subtracted from the readings
obtained in the scans. Now the upper, lower and side boundary conditions were securely
clamped to the frame. The acquisition software was initialized and the strain gages were
balanced and calibrated.
At this stage the specimen was again loaded to approximately 0.9 kN to ensure its proper
seating. The load was removed and the transducer rack was positioned against the plate,
with the transducers compressed to the center of their available range. The deflected shape
of the plate was scanned using the deflection tracker. Vertical scans at nine horizontal po-
sitions (Figure 14.29) were taken and the initial shape of the unloaded plate was defined.
The initial shape included the deformatiorr due to the lateral forces imposed by the trans-
ducers mounted on the rack of stationary transducers (Figure 14.28). These forces were also
considered in the analysis.
Following the deflection mapping, the plate was loaded until its center deflected one half
of the plate thickness. It was observed that during this loading, buckling had already oc-
curred. Therefore, the load necessary to cause the plate to deflect half of its thickness was
defined as limit load. A scan was performed at this load level and the load was released.
The data acquisition was reinitialized and the loading was repeated with scans taken at
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

predetermined load levels.


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1083

In the first test with a type of laminate and boundary conditions. the loading was halted
on ly at the limi t load. The plate was scanned, the load wa~ removed and the plate was loaded
to failure. Failure was defined as the point associated with significant decrease in the plate
load-carryi ng capability {more than 30 percent). Next, laminates of the same type and sup-
ported with the same boundary conditions were sca nned at higher loads when an interesting
mode shape developed during the initial test. The plate center deflection was monitored
during all of the tests. The plates were unloaded after each scan and data acquisition was
reinitialized. Following the final scan the panel was continuously loaded to failure. The above
process was repeated for each of the three types of the side boundary conditions studied in
the program {clamped, simply supported and free unloaded edges).
In summarizing the test results obtained with the TELAC test setup, it was noted that:
The experimental methods yielded reliable and repeatable strain and out-of-plane deflec-
tion data.
The test results of the specimens with the simply supported side boundary condit ions
agreed more closely with the analytical predictions than those of the plates with the
clamped side boundary conditions. It was suggested that this was a consequence of in-
complete clamping caused by the compliance of the Teflon tape at the plate borders.
In -plane sliding was restricted in the experiment because of friction at the boundaries.
particularly due to the normal loads along the unloaded edges. This contradicted the
assumption of in-plane freedom in the analysis.
1/re Struc/1/res Division m NASA Langley Research Cemer is one of the leading groups
in research of composite structures. Since the beginni ng of activities in this challenging field
of engineeri ng, this division has been extensively involved in theoretical and experimental
investigations on buckling and postbuckJi ng behavior of composite structures. These included
tests on Hat panels under axial compression {see for example (14.35] and [ 14.36] and the
many relevant studies cited in them). in which the loading rig depicted in Figure 14.32 was
used for introducing the axial compression load. In this fixture the loaded edges of the
specimen were clamped and the unloaded edges were simply supported by knife-edge re-
straints to avoid buckling of the specimen as a wide column. {This loadi ng fixture is essen·
tial ly a modification of that employed by Davis and Zender 114.17], which is depicted in
Figure 14.11). This loading rig was also used in the tests of Engelstad et al. [1 4.37].
STAGS [ 14.38] nonl inear general shell analysis computer code was used to pred ict the
behavior of the panels tested in 114.35]. It was shown there that the predicted behavior
correlated well with the postbuckling test results up to failure. Finite element analysis was
developed in [14.36)to predict the behavior of the panels. Comparison of the test resulL~ of
[ 14.36] with the predicted ones indicated that the finite element model overestimated the
postbuckling behavior. Finite-clement analysis was also developed in [ 14.37) to predict the
response of the tested panels. Contrary to the observation in [ 14.36], good correlation be·
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 14.32 NASA Langley compressive loading jig of unstiffened


plares (from 114.35))
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1084 Composite Structures

tween the experimentally observed and analytically predicted postbuckling response was
obtained.
Other studies worthwhile mentioning are those reported in [14.39]-[14.41]. In [14.39]
Buskell et a!. presented the results of postbuckling tests conducted at the Department of
Aeronautics, Imperial College, London, on thin rectangular quasi-isotropic graphite-epoxy
panels. The tests were executed in a 10-·ton capacity special-purpose built panel buckling
machine. In this machine the plates were mounted between two rigid platens, one fixed and
one screw driven, and loaded by applying a known displacement to the movable platen. The
stiffness of the specimen was much lower than that of the machine. Hence, careful control
and monitoring of the load and specimen shortening was possible up to failure of the plate.
To simulate the initial buckling mode of a very long panel typical of aircraft structures, the
loaded ends of the plate were clamped, whereas the unloaded sides were simply supported
by adjustable pairs of knife edges. The test results were compared with finite-element pre-

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
dictions obtained by the FINEL finite-element code, developed at the Department of Aero-
nautics, Imperial College, London. Good agreement was found between the axial response
observed in the tests and the finite element prediction. In comparing the observed lateral
deflections of the panel with the finite element predicted ones, it was found that as in [14.36],
good agreement was experienced in the prebuckling and early postbuckling range. However,
in the deep range of postbuckling the finite-element model overestimated the postbuckling
stiffness.
The experimental program of [14.40], carried out by Chang et al. at the General Dynamics
Convair Division, San Diego, California, aimed at validating the applicability of analytical
methods developed for predicting the buckling strength of carbon/ carbon panels subjected
to axial compression. The plates were tested in an Instron Universal Testing Machine. To
avoid brooming of the loaded edges of the panel, the ends were potted with Kerr dental
gypsum compound in steel blocks. This already provided clamped loaded ends. The unloaded
sides were simply supported. This was achieved by using slotted pipes, which were bolted
against the plate. The slots were machined with a 30° angle to minimize the clamping effect
of the pipe wall on the plate unloaded edges. The authors reported good agreement between
the test results and bifurcation buckling load predictions using linearized, specially ortho-
tropic, elastic plate buckling theory. They observed, however, that in general the analytical
methods overpredicted the bifurcation buckling loads.
Based on the above correlation studies, it appears that both analytical solutions and finite-
element analysis can be applied to predict the buckling loads of unstiffened composite com-
pression panels. Neither of these analyses, however, provides an adequate solution for the
postbuckling behavior of these panels. Because the finite-element solutions appear to be less
conservative than the analytical ones, it is suggested that unless the results of further com-
parison studies become available, finite elements solutions like STAGS will be used to predict
the postbuckling behavior of composite compression panels.
In [14.41] buckling characteristics and failure of panels made from carbon fiber-reinforced
plastics faced with aluminum alloy were investigated. The panels had clamped loaded ends
and simply supported unloaded sides. The author indicated that test results compared favor-
ably with theoretical predictions.
Theoretical and experimental studies on compression failure and buckling of sandwich
panels with graphite-epoxy faceplates were reported in [14.31], [14.32] and [12.95] (see
Subsection 12.4.2 of Chapter 12). The panels of [12.95] were simply supported, both along
their loaded ends and unloaded sides. To apply simply supported boundary conditions along
the loaded edges, the end supports presented in Figure 12.39 were used. It should be noted
that, except from transmitting the load into the panel, the aluminum blocks attached to the
ends of the panel were also used to prevent delamination of the faceplates at the loaded
ends. To obtain uniform end loading of the plate, the loading device of Hoff, Boley and
Coan [8.53] was used, with, as indicated by the authors, a notable degree of success. Fur-
thermore, the load distribution was only adjusted under relatively low load levels, between
0.05
Copyright Wiley
and 0.25 of the buckling load, and generally
Provided by IHS Markit under license with WILEY
one adjustment was found to be sufficient.
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1085

The loading rig of Figure 14.23 developed at MIT [14.7] and [14.9] was also used in the
MIT tests on sandwich plates (see [14.31] and [14.32]). Clamping of the loaded ends was
modified to accommodate the thick sandwich plates. Very good agreement was obtained
between the predictions, of both deflections and strains, and the experimental results. It was
also shown that, using analysis and simple failure criteria, panel failures can be predicted,
agreeing fairly well with the experimental data obtained in these tests.

c. Shear Loading
Investigations on buckling and particularly postbuckling behavior of composite structures
under shear loading were strongly motivated by the fact that designs of metal stiffened
structures demonstrate high structural efficiency when buckling of their webs significantly
below limit load is permitted. Consequently, adopting a nonconservative design approach,
stiffened structural components are designed to achieve their ultimate load-carrying capacity
well in a "deep" postbuckled state. Since composite materials possess higher weight effi-
ciency than metals, it was conceived that for similar design conditions composite materials
could provide more efficient structural elements than metals. Furthermore, the available data
in the literature indeed indicated that composite structures have significant postbuckling
strength (see for example [8.125] and [14.42]-[14.44]). Hence, it became obvious that in
order for composite structures to challenge and compete with their metal counterparts, their
postbuckling capability had to be exploited.
The fuselage of an aircraft lends itself best to this design philosophy. Hence, most of the
research on buckling and postbuckling under shear loading of composite structures was
associated with fuselage designs. In these investigations the attempt was made to replace the
conservative design approach and apply a bolder one that exploits the full weight savings
potential of composite fuselage structures. This called for the design of thin-gage composite
fuselage panels that could sustain loads many times higher than their buckling loads. It was
realized, however, that before such an approach could be accepted and fully exploited, the
buckling and postbuckling, as well as the failure characteristics of composite elements under
shear loading, must be well understood. This required an ability to predict adequately the
response of the structure and its failure. Hence, appropriate test apparatuses and testing
procedures were developed (for example [8.110], [14.43] and [14.45]). In [8.110] Farley and
Baker indicated that the commonly used shear test fixtures are satisfactory for introduction
of in-plane shear loading in metallic structures, but they are inadequate for testing composite
structures, particularly in postbuckling. In testing the more brittle composite materials, the
adverse compression and tension load stresses that develop in the corners of the panel and
intensify in the buckled web lead to unanticipated crimping and tearing failure modes and
low failure loads. These failures, which are not experienced with metal structures because
of their ductile behavior and ability to locally redistribute loads, of course have to be con-
sidered and eliminated in the development of the test fixtures of composite shear panels.
Another mode of failure inherent to composite structures and particularly associated with
postbuckling is delamination due to high interlaminar stresses.
Experiments on buckling and postbuckling behavior of shear panels were discussed in
Section 8.4 of Chapter 8. Test programs employing the most commonly used in-plane shear
test methods, diagonal tension and biaxial picture frame fixtures and cantilever and three-
point beam shear fixtures, are presented and evaluated there. Additional discussions on in-
plane shear test methods can be found in [8.110], [8.112], [14.45] and [14.46]. Here, test
programs that were specifically concerned with composite shear panels and their results are
discussed.
Kaminski and Ashton [14.47], Pimm [14.48], Foreman [14.49], G.L. Farley and Baker
[8.110], Bush and Weller [8.112] and Rouse [14.43] used the picture frame concept (see
Section 8.4 and Figures 8.78 and 8.79 of Chapter 8) for introduction of in-plane shear
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

loading. In [ 14.47] Kaminski and Ashton studied the postbuckling behavior of boron-epoxy
panels having an aspect ratio of 3:1. In the program, laminates of various laminate orien-
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1086 Composite Structures

lations, (0190);; (01 :t 60)s and ( ± 45)~ with variable stacking sequence were fabricated from
Narmco 5055 prepregged tape. T he shear frame used in their tests consisted of two sets of
two pin-conne~ted legs that were bonded along the edges of the shear panels and yielded a
6 x 18 in2 ( 152.4 X 457.2 mm 2) test section (see Figure 14.33). The panels were inspected
with an infrared light prior to and after testing to determine area.~ of delamination. Following
the analysis of [14.50) and [14.51]. the critical loads P.. were calculated and compared wilh
the test results. These results demonstrated that all of the panel laminate configurations could
be loaded till deformation into diagonal tension and that they sustained loads many times
their buckling loads. as well as induced deflections that were many times greater than the
panel thickness. A sequence of photographs depicting the development of the diagonal ten-
sion mode of one of the panels tested in the program of [14.47] (-451+451 +451 - 45) is
presented in Figure 14.33.
From their test results the authors concluded that linear buckling theory was adequate for
prediction of the critical loads. which were found to be strongly stacking sequence dependent.
Infrared and visual inspection showed that the primary cause of panel failure was stress
concentrations at the panel corners that induced tearing. which is governed by the in-plane
membrane strength of the laminate. Interlami nar strength was not a limiting factor for these
panels.
Within the framework of the Vought Corporation studie~ to define the optimum type of
load rransfer for fuselage shell structures. Pimrn (14.48) tested shear panels of two aspect
ratios. square alb = I and rectangular alb = 2.5. This choice of aspect ratio was based on
the fact that square panels yield the highest buckling stress for a given b dimension, whereas
the rectangular panel of aspect ratio 2.5 represents panel ~ with an infinite aspect ratio, i.e.
for alb greater than 2.5 the buckl ing stress is aspect rati o-independent. The webs of the

tal 1200 tbs (b) 40001bs


--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

tel 5000 lbs (d) 53001bs


Figure 14.33 Kaminski and Ashton·s pic1urc frame and buckled shcat panel (-45/45/45/ - 45) under
various load levels (from (14.47)): (a) 1200 lb (5 .34 kN), (b) 4000 lb ( 17.80 kN), (c)
5000 lb (22.23 kN). (d) 5300 lb (23.58 kN)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1087

panels tested in the program had various thicknesses and were made of graphite-epoxy and
Kevlar-epoxy prepregged materiaL Two forms of materials were used-tape and fabric. The
graphite-epoxy webs were fabricated from the following material systems:
1. GriEp tape-52081T300 (Narmco)
2. GriEp fabric-HMF 330C Style 133 (Fiberite 9341T300)
3. GriEp fabric-F2631T300 24 X 24 weave (Hxcell)
The edges of the specimens, along which the shear load was introduced, were reinforced
with fiberglass tabs. The panels were cured in a conventional autoclave process.
Except in the early tests, the response of the specimens was monitored by a 45o rosette
gage and three out-of-plane displacement transducers, whereas in the early tests as many as
six transducers were used. Head travel of the loading machine was also measured.
The specimens were tested in a 160,000 lb (711.2 kN) maximum capacity Riehle test
machine. The loads were applied incrementally till failure. In the range of out-of-plane
displacements smaller than half the panel thickness, small load increments were applied.
Upon reaching this magnitude of displacement, the transducers were removed and the panel
was loaded in larger increments to failure. Strain gage readings were recorded up to failure.
A comparison of the test results with the predicted ones (determined by the Vought Lam-
buck code, USAF RA-5) is shown in Figure 14.34 (reproduced from [14.48]). It is apparent
from this figure that correlation is quite good.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Strain readings and out-of-plane displacements were used to determine the buckling load
by the Southwell method. Typical strain readings versus applied load for a panel are depicted
in Figure 14.35, and out-of-plane displacements measurements versus applied load together
with the corresponding Southwell plot are shown in Figure 14.36. Comparison of Figure
14.35 with Figure 14.36 shows that the Southwell plot (Figure 14.36) yielded a critical load
significantly larger than that obtained from the strain readings (Figure 14.35). The inflection
point of the load versus displacement curve in Figure 14.36 is located at a load of about
P = 4,500 lbs (20.02 kN), which is identical to the buckling load determined by the strain
records in Figure 14.35.

1,000
,--~--------

PREDICTION BASED ON MEASURED THICKNESS~

..=
ACTUAL BUCKLING LOAD
PREDICTION BASED ON ------
THEORETICAL THICKNES?__
-n ~-

500 - -
-
400 - - rr
300

~ 2 00 I

~
I I
- -
"'
u
100
-

"

50
40

30

1 2 3 4 5 6 7 B 9 10 11 12 13 14 15 16 1718 19 20 21 34 35 36
SPECIMEN NO.

Figure 14.34 Predicted versus actual shear buckling loads observed in the tests of the Vought Cor-
poration (from [14.48])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1088 Composite Structures

2or------------~-A-P-H-ITE-/-EP-O-XY--TA_P_E____________l

5208/ TJOO
18
LAYUP: {Oj.:t:t,s;o 2;:!: 45;D 2ift.SjO]
t. 0.0643
1&
alb'" t.O
,. ptCR = 4,500
PFAIL • 11 1875
\ 12 PflpCR • 4.19

l,&oo 2.000 3,&&0 4,000 5,000 6,oo:J 7,000


s1rain < • MICROINCHES/inches

Figure 14.35 Typical midpoint strain versus load behavior observed in a test of a square shear panel

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
conducted by the Vought Corporation (from [ 14.48])

Another study by the Vought Corporation, aimed at demonstrating that application of


composite tension field construction provides weight savings over aluminum tension field
construction in lightly loaded fuselage structures, was performed by Foreman [14.49].
To assess the potential of postbuckled shear panel structures, static and fatigue tests with
8 X 8 in 2 (203 x 203 mm 2 ) and 8 X 20 in. 2 (203 X 508 mm 2 ) KEVLAR-epoxy and graphite-
epoxy panels were conducted in steel picture frame fixtures. The results obtained in these
tests are shown in Figure 14.37. They reveal that in the static case the KEVLAR-epoxy
panels possess considerably higher specific strength than graphite-epoxy ones, and certainly
than aluminum panels. It should, however, be noted that though initial buckling of the KEV-
LAR-epoxy panels occurred at significantly lower load levels than those sustained by the
graphite-epoxy ones, their postbuckling failure occurred at loads 30-60 times the buckling
load of the panel, as compared to 2-10 times only for the graphite-epoxy panels. It is
apparent from Figure 14.37 that fatigue loading of the composite panels almost equally
affected the residual specific strength of both material systems. Compared to aluminum, both
types of composite panels showed a potential of more than 200 percent improvement in the
fatigue strength-to-weight ratio when advanced composite tension-field shear panel construc-
tions are employed. The fatigue tests indicated that repeated buckling at the corners of the
panels were the main cause of the reduction in the fatigue life.

Rouse [14.43] reported the analytical and experimental results of a NASA Langley inves-
tigation on postbuckling response and failure characteristics of selected 8-, 16- and 24-ply
fiat square graphite-epoxy plates subjected to shear loading. The study focused on the de-

PCR P~OM SOUTHWELL

b
PCR FROM STRAINS

==-
2
,A A

\ __,- 6
A

tl_.,.,.
6--6---/ 6/P P = INVERSE SLOPE OF LINE "A"
CR
LINE "A"
PCR~ 6,276LB ( 27.93 kN)

0.002 0.006 0.01 0.14 0.18 0.22


LATERAL OISPLACMCNT 0-IN.

Figure 14.36 Midpoint out-of-plane deflection of the shear panel of Figure 14.35 and corresponding
Copyright Wiley Southwell plot (from [14.48])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1089

0.5
STATI C STRENGTH FATIGUE STRENGTH
I<EVlAit.Ef'OXY
H • 100,000 CVCUS
Jt a wl

O•.C GltAI'HITE-EI'OXY
SPlOFIC
STJtl NOTll
fsMA.X tP
IN. X 10·6 O.l
. e.
e
AlUMiNUM
707S-t4

GR.APIIIU · KEVlAR·
lP~XY E~XY

AlUM INUM
0 . 1 f-

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
"~"
Figure 14.37 Static and fatigue tension field shear strength efficiency observed in the tests of the Vought
Corporation (from [14.49])

termination of the parameters that affect the response and failure, and an attempt was made
to fully describe the innuence of orthotropy on the postbuckling stiffness of a composite
shear panel.
The test program employed 8-, 16- and 24-ply orthotropic (±45/0,)so quasi-isotropic
( :!: 45/0/90),, all ± 45° ply and (0/90)5 laminated nat panels. The panels were fabricated
from AS4/3502 (Hercules) unidirectional tapes and were cured in an autoclave, following
the procedures recommended by the manufacturers. The overall dimensions of the laminates
were 445 x 445 mm2 • Metal edge reinforcements were bonded to the specimens with a
room-temperature-cure adhesive. Except for some of the 8-ply specimens that had aluminum
edge rein forcemenL~. all of the specimens had steel edge reinforcements. Bolt holes were
ultrasonically drilled into the metal edge reinforcements of the specimens so as to match the
bolt hole pattern of the loading fixture. The test section of the specimen (Figure 14.38) was
305 x 305 mm2 •
The panels were loaded by the picture frame depicted in Figure 14.39, which was com-
posed of two back-to-back steel rails bolted to the edge reinforcements of the panels. The
rai ls were interconnected with high-strength pins, which did not extend through the test panel

A'HAl EOCE
RE INf O•!C" 1'1:'1'

TfSI SECTION
Figure 14.38 Typical NASA Langley specimen fo r picture frame testing (from (14.43])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1090 Composite Structures

Figure 14.39 NASA Langley piciUre-frame fixture (from [14.43])

and whose centerline was coincident with the corner of the test section of the panel. Locating
the pins at the corners of the test section assured correct picture frame kinematics and proper
load introduction into the test section of the panel by the frame. The load was introduced
into the frame uniaxially by applying a tension load from a 345 kN test machine in the
direction of the diagonal of the specimen.
To monitor the response of the panel, electrical-resistance strain gage rosenes and direct
current displacement transducers (DCDT's) were employed. The latter, which were mounted
on the loading clevis of the test fixture. were used to measure the extension along the tension
diagonal of the panel and the out-of-plane deflections of the specimen. An automated data
acquisition system was employed to record the load. strain and displacement data at pre-
scribed time intervals. The moire fringe technique was also used to demonstrate qualitatively
the out-of-plane displacements of the panel.
The author reponed that the test results correlated well with analytical predictions by the

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
STAGS nonlinear general shell analysis computer code [ 14.38)). Comparisons between the
test results of [ 14.43] and this analysis are depicted in Figure 14.40a and b.
It should be noted that the test results of [ 14.43] show the main deficiency of the pic!Ure
frame test rig. i.e. the development of significant bending strains in the direction of the
tension diagonal of the panel ncar the comers of the panel. However, it appears from these
results that no specimen failed due to this strain concentration. In the case of the thin 8-ply
specimens. high displacement gradients that occurred at the center of the panels. which were
associated with tmnsverse cmcks and delamination between plies, lead to their failure. Ad-
hesive failure between the graphite-epoxy specimen and the steel edge reinforcements, due

IU •N'PltlO lOAO A.l IUCkUHG


l·tMICOIS\

-~~.
0 1tsl
- .tMM.YSIS
e f"llUAl

:J~
NOilll""lllfOCllH I ·P'LN«:lUUCJI(P\ 'ftl l
HOIIMUltDD<JU; II G'< , •· II 6tr
IaI (b)

Figure 14.40 Comparison of the NASA Langley shear test results of a 24-ply quasi-isotropic panel
with STAGS predictions: (a) extension along the diagonal. (b) midpoint out-of-plane
deflections (from [14.43])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1091

to the bearing of fixture bolts on the graphite-epoxy laminate, caused the failure of the thick
16- and 24-ply specimens. In the latter, the adhesive separated from the steel edge reinforce-
ments and bearing stresses, due to the fixture bolts, on the graphite-epoxy laminate caused
failure of the panel.
In the shear tests o{ the Northrop Corporation, Agarwal [ 14.45] used the concept of a
cantilever beam shear fixture with eccentric loading to evaluate the behavior of realistically
configured multibay panels operating well into the postbuckled regime. With this test method
he assessed the adequacy of buckling methodologies and analytical tools for prediction of
the shear response of panels and provided confidence in the viability of this type of high-
performance structures.
A detailed discussion on the selection of a suitable test setup for shear web testing is
presented in [14.45]. The discussion there rules out the usc of the picture frame because of
the severe tension and compression strains that develop at the corners of the specimens, and
eliminates its application for tests on stiffened shear webs because the heavy edge members
of the picture frame preclude any appreciable compression load on the uprights. Similarly,
Agarwal also excludes the concept of a short beam loaded at its ends by eccentric loading
arms [14.45] because there the energy released at failure of the test web may deform the
chords (flanges) of the beam to a degree requiring replacement. Furthermore, this testing
concept has only a single test bay, which precludes realistic loading of the uprights.
It is pointed out in [14.45] that most of the tests with metal tension field beam used the
cantilever beam concept since it accurately simulates the response of an actual aircraft struc-
ture (see also the discussion in Subsection 8.8.4 of Chapter 8). In the testing, however, a
difficulty arises due to lateral instability of the compression chord of the beam. This problem
is overcome by increasing the dimensions of the chord, providing lateral supports or both.
Increasing the size of the chords obscures the shear concentration in the corners, which are
induced by deflection of the chords between the uprights. This concentration is a real-life
effect that has to be included in the test and evaluated by it. Introduction of lateral supports,
on the other hand, complicates the test setup and increases the cost of the experiment.
Another difficulty associated with cantilever beams is due to the incomplete rigidity of the
support of the beam. The presence of a slight rotation of the support also complicates the
determination of the stiffness of the beam. This difficulty can be eliminated by testing simply
supported beams with a reinforced center bay. Furthermore, it is required that each half of
the simply supported beam will be as long as an equivalent cantilever beam in order to
reproduce the upright loads as accurately as in the cantilever beam. Of course, this increase
in length significantly increases the lateral instability and substantially increases the cost of
the specimen.
A compromise between the concept of the cantilever beam specimen and the eccentrically
loaded short beam specimen is provided by the cantilever beam with an eccentric loading
arm at the free end, which was suggested in [14.52] and [14.53] and was adopted in the
tests of [14.45]. Figure 14.41 depicts this arrangement, which can have realistic chord areas
because of the reduction in the chord loads. This solution still requires some degree of lateral
support (see Figure 14.41). The choice of this testing configuration was based upon its
efficiency and effectiveness and its relatively simple adaptability to changes in web config-
urations without requiring major modifications.
The test specimens of [14.45] consisted of stiffened panels with two cocured stiffeners,
either hat sections or !-stiffeners (see Figure 14.42). The sectional properties of the stiffeners
were identical, except for their torsional stiffness. The two different stiffener shapes were
chosen in order to explore the different effects of closed and open sections on the behavior
of the panel. To allow different buckling loads, two different stiffener spacings were used,
330 mm and 228.6 mm for the hat-section stiffeners and 330 mm for the !-stiffeners. The
stiffeners were quite rigid to ensure that failure would occur in the skin of the panel. To
ensure that the proper load was introduced into the panel without failure, the thickness of
the panel edge was increased.
Copyright Wiley
--`,`,`````,`,```

Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1092 Composite Structures

LATERAL
~~~- TEST SPECI>lEN ~~-
SUPPORTS

I
I
G liil

1-
~ ~ ~ !! ~I
I
>
:>:w
:>:UJ
a"
I II
u

t2l [l

If
j ~
ECCENTRIC LOADING ARt-1

Figure 14.41 The cantilever beam shear fixture with eccentric loading used in the tests of the Northrop
Corporation (from [14.45])

The panels were fabricated from 3501-5 HMF woven graphite-epoxy and 3501-6 AS
graphite-epoxy tape. The removable rubber mandrel technique was employed for fabrication
of the panel, and the stiffeners were cocured with the skin. The stiffeners of the hat-stiffened
panels were laid over the rubber mandrel, and a vacuum bag was used to retain their shape,
while in the case of the !-stiffened panels rubber mandrels were used to support both sides
of the vertical legs of the stiffeners and then the whole assembly was vacuum bagged.
The panels were attached to the loading frame by 6.34-mm bolts at an approximately
25.4-mm spacing along all four edges of the panel, and 6.34-mm aluminum plates were used
as web end bays to transfer the shear load into the test panels. The caps (chords) of the test
fixture were made of steel and were quite stiff.
Load was applied to the test fixture with a hydraulic cylinder. To control the pressure
supplied to the cylinder and maintain fine control, an Edison' hydraulic proportioning unit
was used. The load was measured by a load cell, and strain and displacement data were
recorded at each load increment. The panels were instrumented with back-to-back strain
gage rosettes, which were used to determine the onset of buckling and the strain in the
tension diagonal corner, with back-to-back axial gages to measure the strains in the upright
stiffeners and deflection gages to measure deflections at several locations of the web as well
as those of the metal chord of the loading fixture. The normal deflection contours of the
panels were also visually monitored with aid of a moire grid. These observations were only
qualitative.
Agarwal's results [14.45] were compared with analytical predictions. It was shown that
the experimentally observed failure loads were only half of those predicted by application
of the metal-based tension field theory. This discrepancy was explained on the basis of the
experimentally predominant failure modes of the panels. compression failure and stiffener-
web delamination. Both these modes differ from those experienced with metal shear webs,
where failure is due to web rupture or clue to formation of permanent buckles in the web.
To improve the analytical predictions, calculations were carried out with the NASTRAN
[2.98] code, which allowed modeling of the failure of the skin in the tension field corners,
and the response of the edges near the skin-stiffener interface was studied. The analysis with
NASTRAN yielded good predictions of the postbuckling response, as well as of the failure
load for the compression failure mode in the diagonal tension corners. Comparisons of the
experimental results with the analytical ones for load versus applied vertical displacements,
out-of-plane displacements and maximum diagonal strains [14.45] arc depicted in Figs.
14.43-14.45, respectively. It is apparent from these figures that the correlations between
experiments and analysis are good.
Agarwal noted that large concentrations of normal forces near the diagonal tension corners,
where the stiffener-web clelaminations were observed, resulted from the constraint against
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1093

I~
A rA ___A_ ----~---- ---0.--,

1
_l__j
--l---- ·1·-- _::--1--
e u
go•:
I
I
I I
L ____ _ _ _ _ _ _ _ _j

l---o-1
(&,2;o;~2;o2;~2;0ls

c~Z:zl
i-2.4 CM-1--HI CM--1
SECTION A-A

SYM
[I..§;Ozl2_) 5 ~0.8•1 CM

.1?..4-==.!i_~--~u
'C:Jl
·-=· "'J
1 45 -o-45·01
"'-"'2· s I J.JCM

~ ~
I 1.27CM~ ~w--I I
1..1.1 2;01 5 2.54
4.oocM-
SECTION B-B IHAT STIFFENEnJ

SYM

1.2/CM-
-2.54CM-
~ I I
4.00CM-
I
SECTION ll-llli-STIFFENEnl

Figure 14.42 The Northrop Corporation shear beams-panel and stiffener details (from [14.45])

I' LOAD (KN I


u
;;;
>
!z Q6 1.50
w
~
w ~

3 0,4 1.00 u -
0..
Ill
>
0 0.2 0.50
0

""
..J
0..
0..
<( 4,000 B,OOO 12,000
LOAD ( LBS I

Figure 14.43 Comparison of experimental results for applied load and displacements observed in the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

shear tests of the Northrop Corporation with NASTRAN predictions (from [14.45])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1094 Composite Structures

Figure 14.44 Comparison of NASTRAN predictions of nonnal displacement contours at N,, = 175
kN/m with moire fringes at N., = 177 kN/m recorded in a shear test of the Northrop
Corporation (from [ 14.45])

normal displacements. However, the failure loads associated with this failure mode were not
calculated, since there are no provisions in the NASTRAN finite-element model to account
for this type of failure mode.

In Subsection 8.4.4 of Chapter 8 the extensive Technion Aircraft Structures Laboratory


investigations. employing the three-point beam shear fixture (Figure 8.87) for studying the
postbuckling behavior of metal shear panels, were discussed and assessed. The three-point
loading rig of Figure 8.87 was also adopted in the extensive Technion experimental studies
on postbuck.ling response and the behavior atier extensive repeated buckling of composite
shear webs carried out by Weller and Singer [8.125], and Weller eta!. [ 14.54] and [ 14.55].
In these studies the loading apparatus of Figure 14.46 (sec [8.1 251 and [ 14.56] was used in
parallel to that of Figure 8.87. It is apparent from Figure 14.46 that the manner in which
the reaction loads are introduced into the edges of the shear beam, by the loadi ng frame
depicted in this figure, differs from that of the rig in Figure 8.87. In the loading fixture of
Figure 14.46 the simple support edge loading pins were directly inserted into both the loading
holes at the edges of the beam and the loading holes in the supporting brackets, which were
rigidly bolted to the vertical guide columns of the MTS loading frame. Contrary to the
loading fixture of Figure 8.87. this type of end loading restrains the horizontal movement of
the beam during postbuckl ing and thus may affect the stress field experienced by the beam.
Nevertheless, comparisons of the strain measurements associated with the loading frame of

- THEORE1'1CAI. RESUt. T$
ejO E.XPER:IMENTAl R~SIJLT S
Z PAI'Itl lA
~
a:: .yb
EXPERIIMEN1AL RESUlTS
t;) 0.006 t-L--PA_" E_
l _ >_
A - -- - '
w
2:
m
w
g: 0.004
~
<.>

o.oo2

Figure 14.45 Comparison of experimental maximum diagonal strains observed in the Nonhrop Cor-
pOration shear tests with NASTRAN predictions (from [1 4.45])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1095

Figure 14.46 Recent version of three-point beam shear fixture used in the Technion ASL tests (from
[8. 125))

Figure 14.46 with those observed in the beam when loading it by the rig of Figure 8.87
revealed only insignificant differences.
As in Subsection 8.4.4 of Chapter 8 and in (14.44], two main considerations justified the
choice of the three-poi nt loading concept: realistic test conditions and a relatively high load
ratio compared with buckling. Furthermore, as pointed out there, the three-point shear beam
configuration, the Wagner beam-type structure, simu lates common aeronautical applications,
insofar as stiffening and load transmission arc concerned, and thus provides a realistic test
environment. Consequently, as in (8. 122)-(8.124], the core of the investigations of [8.125]
and ( 14.54]-( 14.56] consisted of instrumented repeated buckling tests on three-point loaded,
five-bay Wagner beams under realistic test conditions, supplemented by numerical analysis
with STAGS [ 14.38). The focus was on the influence of web configuration, the surrounding
structure and stiffeners, on the response of the shear panels and their durability under re-
peated buckling. To achieve these objectives, the effects of varying sizes of the stiffeners,
the magnitude of the initial buckl ing loads, the panel aspect ratio and the cycling shearing
force V,1 ,.. to which the panel is exposed, were evaluated.
To meet these research goals, shear beams of hybrid type construction were used. As a
result of adopting the test configuration of Figure 8.87, the beams were identical in dimen-
sions, except for the web, to the metal beams of Figure 8.88 (Subsection 8.4 of Chapter 8).
The beams were constructed from a 2024 T3 aluminum frame (L 20 x 20 x 2 mm~ Hanges
and uprights) bonded in a specially designed template by EA-9309.1 adhesive (Hyson comp.,
IAI code) to composi te webs separately fabricated from either graphite-epoxy T300/5208
unidirectional tapes or graphite-epoxy planeweave 3502/3K-70-PW I AS4 tapes. The webs
of the test beams had the following stacking sequence: 4-ply ( +451 - 45)s, 5-ply ( +45/ - 45:
O)s, 7-ply (+ 45/ - 45/90/0)s and 2-ply (±~s (where ± 45 stands for planeweave).
The beams were cured and vacuum bagged at room temperature to avoid compatibili ty
problems between the aluminum stiffeners and composite web. The controlled bonding pro-
cedures were developed by Israel Aircraft Company and complied with ai rcraft standards,
so as to be able to sustain the loadi ng spectrum and modes planned for the test program.
Pairs of strain gage roseltes were bonded face to face on the surfaces of the web according
to Figure 14.47. Gage rosettes were located at the edges of the tension field diagonal, adjacent
to the corners where the buckle of the web interacts with the stiff frame and where high
local stresses, which may lead to failure initiation and fatigue problems. develop. Rosettes
were also positioned at the center of the web to detect incipient buckling as well as to
eval uate the average overall response of the web. Si nce high bending stresses develop in the
Hanges of the beam at its center, pairs of unidirectional gages were also bonded face to face
to the flanges of the beam to monitor the strains developed there and their behav ior during
fatigue cycl ing of the beam.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1096 Composite Structures

1(11)
!(ttl

T(IT)

,. I
41'

~~
1...
ll
I
lZ
z
ll
l
14

15
5
lf
I
l7
7
ll

.
I
lt (lAO:)
I (fDTJ

·] ]· . .. . .' . ' ' ' ' '


01.
111U

...
0 I
112 '
'

........
Ol A
031

.
....
l(U)
(a) '

011

I( Ill
......., .
07.
07 I

.,.••
T(17)

10.
' '
I A•

·1 1(151
4CI4l


I I •
it A"
1Z •• . ' -
• see TENSION case (a)

(b)

Figure 14.47 Strain gage location in the Technion ASL shear webs (from [8.125]): (a) outermost panel
fibers in tension, (b) outermost panel fibers in compression

The measured strains were recorded by a multichannel data logging system (KIOWA
UCAM-lOA). In addition to strain gage measurements, the shadow moire technique was
employed for global qualitative observation of the deflected pattern of the web on the way
to and during buckling and in the postbuc:kled region. This technique was also used to track
dynamic damage propagation in the web, as reflected by a continuous change of the deflected
pattern of the web (see Figure 14.48).
In contrast to the metal shear beams of Section 8.4, the three-point loading test configu-
ration with the composite webs results in different stress states in the two test webs of each
beam, either tension or compression in the outermost fibers of the test web. This significantly
affects the test results because the modes of failure and the fatigue, residual and ultimate
strengths of the web are all dependent upon the state of stress in the fibers of the outermost
layers of the laminate. On the other hand, this unsymmetric behavior furnishes an advantage
since it can be exploited to conduct simultaneously a different test on each test section of
the web and thus save considerable testing time. Furthermore, the use of the dummy plate
mentioned in Section 8.4 is avoided, so that no unknown and undesirable strains are intro-
duced into the web stiffened by the dummy plate prior to its testing. In essence, the simul-
taneous testing of both test sections ensures that their testing commences from a state of
virginity.
Recognizing the drawback of the commonly used in-plane shear-test methods for testing
of composites, Farley and Baker [8.11 0] developed an improved shear test fixture (Figure
14.49) for accurate determination of the response of thin-gage composite panels subjected
to in-plane shear loading. Development of the shear test fixture was based on a finite-element
analysis of the effects of shear fixture kinematics and of the loading tab stiffness on the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

shear panel in-plane stress distribution. For this purpose the location of the pivot points
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1097

(a) (b)

(c)

Figure 14.48 Moire fringe paucms indicating dynamic damage propagation in Technion ASL shear
tests (from [8.1 25]): (a) N = 24.500 cycles. (b) N = 27.500 cycles. (c) N ~ 27,500
cycles. after complete failure

(corner pi ns in Figure 14.49) of the shear fixture rails was varied according to Figure l 4.50a.
The resu lts of this finite-element analysis are depicted in Figure 14.50b.
It is apparent from Figure 14.50b that location I (the centers of the pins located at the
comers of the test section) yields almost unifonn shear stress distribution all over the panel,
with negligible stress gradients and nonnal stresses. In this case the fixture kinematics re-
semble ideal shear panel deformation (Figure 14.51a and b). Locating the pins in position 2
corresponds to the fixture-tab ki nematics of Figure 14.5 1c. In this case the ends of the loadlng
tabs along the tensions diagona l auempt to overlap. thus inducing compressive in-plane
normal stresses. while the ends of the tabs along the ot her diagonal move apart, introducing
tensile in-plane normal stresses in other cornet-s. The compressive stresses may cause local
buckling of the panel, leading to crimping or folding of the panel, whereas the tensile stresses
can tear the panel apart along the diagonal or along the loading tab-panel interface.
The finite-element analysis also showed that the in-plane shear stress distribution was
significantly affected by the ratio or the in-plane stiO'ness of the loading tabs to the stiffness
or the panel (see Figure 14.52) and the distance from the comer pins to the nearest bolt used
to transfer load between the loading fixture and the panel. It was also shown by the analysis
that the method of load int.-oduction (Figure 8.77) is immaterial, provided that the fixture
corner pins were properly located and the loading tabs and fixtures were sufficicmly stiff
relative to the panel test section. Elimination or removal or the corner pins. as in beam tests.
aggravate the corner stress problem. --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1098 Composite Structures

(a) ( c)

(b)

Figure 14.49 Improved shear test fixture used in the testS of Farley and Baker (from [8.11 0]): (a)
improved picture frame fixture, (b) improved picture frame in test machine, (c) detailed
expanded view of improved picture frame

rtOAOING TABS
3"'\ I ;-_3
z-;:; - .;~-2
'-1 1-'

PAIL!

,rl h,
( a) 2 .. .. .._z
3../ \ "\..:3
CORNlR PIN POSITIONS 1, 2. 3

00~
CONTOOR INTERVAL =O.Ol COHIOUR Uff(IVAL- O.OS
POStliON I POSitiON 2
COHJOUR I NliR VA~ • 0,20
POSITION 1

Figure 14.50
( b) ~.,J ~1J ~~
CONTOUR INT(RVAL ,. 0.01 COHIOI.Ii INtUtVAl a 0.10 COHTOI..Iit INitllVAl • 0,)0

Finite-clement analysis of the effect of picture frame comer pin position on shear stress
and normal stress distributions in the shear panel (from (8. 110]) (a) comer pin locations,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley (b) stress distributions corresponding to pin position


Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1099

UNDEFORr.'.ED IDEAL DEFORMATION


(a)

/TASSp-\·.
.

0
.

SHEAR
FIXTURE • . ..•
CORNER

lJNDEFORMED
PINS DEFORMED
(b)

(c)

Figure 14.51 Fixture frame kinematics (from [8.110]): (a) ideal shear deformation, (b) Farley and
Baker's fixture tab kinematics with pins at corner of panel position 1 of Figure 14.50a,
(c) fixture tab kinematics with pins at position 2 of Figure 14.50a

Following the analytical results, the improved shear fixture (Figure 14.49) consisted of
three parts: the load introduction structure (Figure 14.49a), the picture frame fixture (Figure
14.49b) and the test panel. The picture frame assembly is bolted to the load introduction
structure, which is a built-up I-beam assembly bolted to the load platens of the test machine.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Vertical load is applied along one side and reacted by the other side. Application of the
vertical load is associated with horizontal contraction. This contraction is compensated for
by horizontal movement of the load platen, otherwise the fixture would bind and cause
nonuniform load distribution.
Two back-to-back picture frames constitute the picture frame fixture (Figure l4.49b ),
bolted to the load-introduction structure. Each frame is composed of four rails, which are
connected with high-strength pins. The pins are positioned with their centerline coincident
with the corner of the test section of the panel. The comer pins do not extend through the
test panel.

78

1
xy/ 1 applied 1
xy/ 1 applied 1 xy/ 1 applied
CONTOUR INTERVAL= 0.08 CONTOUR INTERVAL= 0.07 CONTOUR INITRVAL = 0.10
P~~~L s;;;:~~:is " 30 "10 "l

Figure 14.52 Finite-element analysis of the effect of tab stiffness on shear stress distribution in the
Copyright Wiley
Provided by IHS Markit under license with WILEY
panel (from [8.11 0]) Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1100 Composite Structures

Three combinations of composite fabric materials were tested in the improved shear fix-
ture: ( ± 45) 5 and ( ± 30/ -60); Kevlar-epoxy (Kevlar-49 (style 285)/5208 epoxy) and ( ± 45
Kelvar/-45 graphite) (Kevlar-49 (style 285)/T300 graphite (Satin weave)/5208 epoxy).
Stainless steel loading tabs, 3.2 mm thick, were bonded to all edges of the panel using a
room temperature adhesive to prevent thermal buckling of the panel. The corner of the
loading-tab was cut out using the tangent-circle concept (Figure 14.53). This concept was
chosen to eliminate a notch being cut into the corners of the test section and to allow large
in-plane panel deformation without contact of adjacent tabs. It should be noted that the holes
in the corner of the panel are not aligned with the corner pins.
The combination of design and sound analysis used to develop the shear fixture of [8.11 0]
represents an appropriate approach to the development of a test apparatus that meets the
desired goals of a test program. It should, however, be noted that in spite of the satisfactory
test results reported by the authors, the shear fixture constructed by this method appears to
be relatively complex and rather expensive compared with the other apparatuses and test
methods discussed in this section.

A large-scale shear test was performed at NASA Langley Research Center by Bush (see
[14.57]). In this test a three-point deep beam shear fixture (see Figure 14.54) was used for
buckling of a 36 X 47 in. 2 (914.4 X 1193.8 mm 2 ) graphite-epoxy (Thornell 300/Narmco
5208) sandwich shear web with 10-plies, ±45 facings (see Figure 14.55). The test aimed at
identifying problem areas for sandwich shear webs that are stability critical and at assessing
the adequacy of analytical tools for prediction of shear buckling.
The sandwich shear web was designed to be stability critical at a shear load of 5000 lb/
in. (875 N /mm). The 10-ply facings were found to comply with this requirement and pro-
vided an adequate strength margin. Buckling calculations using NASTRAN [2.98] and the
code of [14.13] were performed to determine the core thickness that would meet the load
objective. Based on the NASTRAN results, the web configuration of Figure 14.55 was se-
lected.
Three-point loading (Figure 14.56a) was employed for testing of the sandwich shear web.
The test frame consisted of a deep beam with built-up aluminum caps (Figure 14.56a). One
half of the beam, the dummy stiffened aluminum part, was conservatively designed. The test
shear web and the dummy stiffened plate were bolted to the beam caps. Apparently this test
method subjected the specimen to in-plane bending in addition to the desired shear stresses.
Based on preliminary designs, the magnitude of the bending stresses was anticipated to be
about 10-15 percent of the shear stress magnitude.
Fabrication of the shear web consisted of individual molding of each shear web facing,
vacuum bagging and autoclave curing at 350oF. After cure, each face underwent a C-scan
to locate defects. The aluminum H/C core edge (Figure 14.55) was filled, the complete
sandwich was assembled, bagged and both facings were simultaneously bonded to the core
with Hysol AE-951 in an autoclave. After bonding, each side of the panel was C-scanned
to identify defective bond areas.
The shear web was installed in the frame and surveyed to determine pretest out-of-plane
deformations. It was found that the center of the webs deviated from flatness by 0.012-0.20
in. (0.305-5.08 mm), whereas the periphery was essentially fiat.

STEEL TABS

A
BONDLINE

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 14.53 Loading tab assembly on the shear panels of Farley and Baker (from [8.11 0])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1101

Figure 14.54 NASA Langley three-point deep beam shear fixture (from [ 14.57))

The frame was installed in a 1.2 million lb (5.33 MN) compression test machine and
resu·ained laterally, as shown in Figure 14.54. The frame was supported at the bottom at
each end and load was applied at the top center. The tension bar lateral restraint system did
not affect the in-plane defonnation of the web. The total applied design load was 360 kips
( 1.6 X 10' kN).
The shear webs and the test frame caps were instrumented by 97, mostly back-to-back,
strain gages to survey the shear field as well as the beam bending strains (Figure 14.56a and
b). The moire fringe technique was employed to observe the lateral deAections of the web
and identify the buckling mode shape.

I .. I
"
1,.--·--.: :. :. -:.. ':, - .- .= :. .::. ~- ..- .: ___..._..__-:.: ---'
I•
II
..
,,
II f- ~1. 9

,, (!45·)~
s
I
,, 2,.9 - 44 $.

-4'"
II _.!!!D~
DI R,( ( TI Oti
,+
••
•• \\.
........ +.+...+ '"+........:--_ - ..=..::: '~.' ~- --- - - - - --
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

l..7S" 1~•.so"
I I

·W.@ ! ~ ! W#;! ll lllllllll ll llll11


/
i
I
'/,
I
L H AP( X 1221 ;.
ll. )I FI LLER l~ f.t.f' O.OSS..,( IO PLIE~,14S')~
fig ure 14.55 Design details fo r stability-critical shear web used in the NASA Langley large-scale test
(from [ 14.57])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1102 Composite Structures

OUTS!OE CAP- !.TRENGTH W£.9

P/-;r
~--------------------~

(b)

Figure 14.56 NASA Langley shear web test frame (from [14.57]): (a) strain gage instrumentation on
frame caps and uprights, (b) shear web strain gage layout

A comparison of test results with NASTRAN predictions is shown in Figure 14.57. It was
observed that failure occurred at 69 percent of the anticipated buckling load. It was assumed
that this difference was due to sensitivity of the web to initial out-of-plane induced manu-
facturing displacements. In spite of the lower-than-anticipated failure load, the web was
found to be significantly more weight efficient than the most efficient aluminum structures
or the titanium clad boron-epoxy web it was designed to replace.

d. Combined Loading
In spite of the fact that real structures are exposed to combined loading, stability and post-
buckling behavior studies on laminated composite plate elements subjected to combined
loading are quite scarce compared with the numerous uniaxial loading investigations (see
[14.5] and [12.89] and the many studies cited in them). Furthermore, hardly any experimental
investigations on the response of composite plates under combined loading are reported in
the literature.
Some theoretical aspects and results on buckling of composite plates under biaxial loading
are presented in Subsection 14.2.1 a of this section. Additional theoretical considerations and
results can be found in the surveys of [14.5] and [12.89], and in [14.13] and [14.58]-[14.67].
Since the beginning of the nineties, Romeo and Frulla of the Department of Aerospace
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Engineering, Turin Polytechnic, Italy, have become deeply involved in analytical and exper-
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1103

400

)00

I
~t T f
I
I
I
I

100
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

0~--~--~----~--~--~
0 ,I ,2 ,) .4 ,5

Dt$PU.CE.MENT • I N

Figure 14.57 Comparison of the observed experimemal dellections at midpoint of the web of the NASA
Langley large-scale shear beam with NASTRAN predictions (from (14.57])

imental studies on buckl ing and postbuckli ng behavior of laminated plates subjected to com-
bined axial compression and shear loading (see [ 14.67]-(14.70]). Ro meo and Frulla could
not fi nd any experimental resu lts tO compare their analytical results with and thus evaluate
the applicability of their analysis to the buckling and postbuckling behavior of laminated
plates subjected to simu ltaneously applied biaxial compression and shear loading. They there-
fore designed and constructed the testing setup depicted in Figure 14.58 and conducted the
required experimental program.
In their testing machine a maximum compression or tension load of 500 kN, a transverse
compressio n o r tension load of 200 kN and either a positive or negative shear load of 200
kN can be applied to panels with dimensions up to I ,000 x 700 mm. The loading and
supporting frames of the testing machine are made of steel alloy and 2024 aluminum alloy.
Two separately controlled servo-actuators introduce the longitudinal load. and a displacemenl
control is used to maintain the panel ends parallel to one another, within a controlled angle
of rotation of less than 0.00 I •.

Figure 14.58 The Politecnico di Torino combined loading testing machine (from [14.70))
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1104 Composite Structures

The transverse load application sy~tcm floats to avoid interference with the longitudinal
and shear loads. It consists of two separately controlled servo-actuators that mai ntai n the
sides of the plate parallel to each other. A servo-actuator applies the shear load to the bollom
end of the panel.
The frame test fixture employed for introducing load~ into the panel is composed of two
~tiff L-steel rails bolted to the Hat ends of the loading frame and to the four edges of the
panel (Figure 14.59). Hence. the boundaries of the panel arc nominally clamped.
The live servo-actuators introducing the load into the panel are separately controlled. This
allows application of various load combinat ions, including triangu lar and trapezoidal shape
loading alo ng the plate ends and sides.
The panels u~cd in the test program were fabricated from (M40/9 14V) graphite-epoxy.
They were vacuum bagged and autoclave cured. The in-plane strains were mea~ured by 50
strain gages. either linear or roselle~. bonded back to back to the panel 'urfaces (Figure
14.59a). Ten computer-controlled magnetic-resistive transducers were positioned on the cen-
ter cross-lines of the panel to monitor the out-of-plane displacement. Simultaneously. the
shadow moire technique was used to obtain a full-field qua litative represemation of the o ut-
of-plane dellection. Provisions were also available for mea~uri otg the initial geometrical im-
perfection of the panel.
Romeo and Frulla found good correlation between the predictions obtained by their POB-
UCK code (14.70] and their test re;ult~. This indicate> that apparently anisotropic stiffness
and boundary conditions were adequately considered in the analysis. A typical comparison
between experimental resu lts obtained with the testi ng machine of Romeo and Frulla and
predictions yielded by their analysis (for a panel loaded by biax ial loading and shear. where
the transverse and shear load arc respectively 40 percen t and 50 percent of the longitudinal
load) is shown in Figure 14.60. It is apparent from Figure 14.60a that the experimental
results corresponding to a plate with a maximum amplitude of initial imperfection-to-thick-
ness ratio of W0 / H = 0.75 arc in very good correlation with the analytical predictions, as
long as the magnitudes of the applied loads are smaller than 1.5 times the cnt ical load of
the plate. Romeo and Frulla suggested that the discrepancy between experimental observa-
tions and analytica l predictions. beyond this load level, might be due to changes in the
dcfom1ed postbucklcd shape of the panel d uring increase in load. Romeo and Frulla also
argued that discrepancies could stem from the con~ideration of either load-controlled or
displacement-controlled edges of the panel. They showed that in the initial range of post-
buckling (up to about 1.5 times the buckling load) very good agreement between test results
and analysis was experienced with the load-controlled edges. With further increase in load,
the displacement-controlled edges yielded better correlatio n with the test results.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(a) {D)

Fl~:ure 14.59 The Poli tecnico di Torino tCit panel connected with I.-steel rai l; to the loading frame
(from [14.701>
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1105

(a)

(b)

Fl~:ure 14.60 A lypical compari'>On belween 1es1 rcsuh~ obwined by lhe Turin group and 1heore1ical
prediclion~ (from ( 14.70)): (a) longiiUdinal load \Crsus oul-of-plane di<placernelll for a
panel loaded in biaxial loading and shear, (b) ou1-of-plane deOec1ion for a panel loaded
al P, = 175 ~N and CN, - OAN,_ N .. = O.SN )

Another experimental and theoretical investigation on the buckling behavior of laminated


rectangular plates subjected 10 biax ial loadi ng was presented in [8.158] (see also S ubsection
8.6.2 of Chapler 8). In this study a modified speci men, I he corners of which were truncaled,
wa~ employed. This truncated specimen was introduced in order 10 eliminate 1he interference
among the four supporting grips at each comer of 1he pa nel during the continuous biaxial
compression process. It wa~ pointed out in [8.158) 1hat when the cut comers were kept small
relathe to the plate dimen~ion~. the buckling response of the panel was close 10 1hat expe-
rienced by a complete ideal panel.

e. Crippling
The crippling failure phenomenon of co lumns and stiffener clements of thin-walled cross
seclions. in parlicu lar metallic ones, is discussed in Subscclion 2. 1.3 of Chap1cr 2, Seclion
6.2 of Chapter 6 and Sec1 ion 8.5 of Chapler 8. It is ind icaiCd 1hat following the loca l buckling
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

of lhc nange and web clements of the columns/stiffeners, funher increase in postbuckli ng
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1106 Composite Structures

loading is associated with load redistribution, such that the excessive applied load is mainly
being sustained by the junctures, or corners, between the flanges and webs of the columns/
stiffeners. This load redistribution results in a stress build-up in the corner regions leading
to local crippling failure that is initiated due to material yielding in the highly loaded corners/
junctures.
The thin-walled columns and stiffeners of efficiently designed semi-monocoque construc-
tions can therefore support compressive loads considerably larger than their local buckling
loads. Furthermore, comprehensive collapse of stiffened structures is precipitated by crippling
failure of the stiffeners.
The crippling failure mechanisms of columns and stiffeners made of laminated composite
materials (which provide superior strength-to-weight and stiffness-to-weight ratios) differ
from those experienced with metallic columns/ stiffeners; and unlike the crippling failure
mode of metallic thin-walled columns and stiffeners, they are not well understood. Experi-
ments demonstrate that composite thin-walled columns and stiffeners, due to their suscep-
tibility to delaminations and other brittle failure mechanisms, do not necessarily fail at junc-
tures or at corners between the flanges and webs.
It was pointed out in Subsection 2.1.3 and Sections 6.2 and 8.5 that in the absence of
satisfactory analytical solutions, the crippling stress has to be determined by semi-empirical
methods. Hence, studies dealing with crippling stress predictions had to rely solely on ex-
perimental studies and observations. Numerous experimental investigations on the crippling
behavior of metallic elements were performed, thus providing good insight and understanding
of the phenomenon. These investigations also provided semi-empirical formulae and design
curves for crippling predictions.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

A semi-empirical approach was also adopted in the study of the crippling behavior of
laminated composite columns and stiffeners. Most of these investigations were surveyed by
Bonanni, Johnson and Starnes in [14.71] and [14.72].
An intensive effort to develop tools for analyzing the crippling problem for stiffened panels
fabricated from AS/3501 and T300/5208 graphite-epoxy was conducted at General Dynam-
ics Convair Division, San Diego, California, by Spier and Klouman (see [6.24], [6.25] and
[14.73]-[14.76]. Spier and Klouman in their studies adopted the traditional methodology of
aircraft strength-of materials-analysis, used for predicting crippling of metallic stiffeners, in
conjunction with lamination theory. The predictions yielded by this hybrid approach corre-
lated well with their empirical observations, in spite of the fact that the panels failed by
delamination rather than by yielding (as was the case in metals).
Spier [14.73] studied analytically and experimentally the crippling and wide column be-
havior of panels stiffened by either rectangular or !-section shapes. The stiffened test panels
were fabricated from A-S /3501 graphite .. epoxy, their skins were pseudo-isotropic laminates
and several laminations were selected for the elements of their cocured stiffeners. Epoxy
potted end supports encapsulated in machined parallel aluminum channels, which provided
almost clamped-end conditions, were employed to introduce the axial load into the panels.
To achieve either crippling or wide-column behavior, various lengths of the panels were
carefully selected: for crippling a short, 12.7-in. (~323-mm) clear length between potted
ends, panel with !-section stiffeners; for wide-column behavior a moderate-length, 17.9-in.
(~455 mm) clear length !-section stiffened panel and a 21.7-in. (~551-mm) clear length
blade-stiffened panel.
The dimensions of the short and moderate-length !-stiffened panels are shown in Figure
14.61. Details and dimensions of the cross-section of the !-section stiffeners are presented
in Figure 14.62, and dimensions and details of the blade-stiffened panel are given in Figure
14.63.
The crippling failure pattern of the short panel is depicted in Figure 14.64, showing the
considerable delamination and fracture along the skin and through the stiffeners. The co-
curing process provided a real integral skin-stringer structure and thus failure did not occur
due to skin-stiffener separation. Concerning the wide-column failure of the moderate length
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1107

-j r- 1.0 TYP
~ 1 ALUMINUM
PANEL/~ CHANNEL

EPOXY POTTING / L PQT ENOS AS


SHGNN
MAKE PARALLEL

='l JL
~ c ~~J 50-~------- 3.50 ------1
·---- b - - - - - - - ---

Figure 14.61 Dimensions of the !-stiffened panels used in the Convair crippling tests (from [14.73])

!-section and blade-stiffened panels, it was reported in [14.73] that failure was associated
with slight skin buckling between the stiffeners and delamination both in the skin and stiff-
eners without any separation of stiffeners from the skin.
Good to excellent correlation between the test results and the predictions based on the
above-mentioned hybrid approach was obtained. It was, however, indicated that lack of data
on the compressive behavior, particularly on the nonlinear part of the compressive stress-
strain response, of graphite-epoxy laminates poses a problem in the analysis and design of
graphite-epoxy stiffened panels.
Spier and Klouman [6.25] attempted in their test program to determine the compression
behavior of A-S/3501 graphite-epoxy laminates and to provide complete compressive-strain
curves. In this study, estimates for characteristics of angle-ply laminates were furnished by
combining test results and lamination theory. The tests were performed employing either the
Celanese method (for details see ASTM Standard D-3410) or a crippling test fixture.
Further experimental studies to provide the necessary crippling data on no-edge-free and
one-edge-free A-S/3501 and T300/5208 graphite-epoxy plates were reported by Spier and
Klouman in [14.76] and by Spier in [6.24], respectively. The plate specimens had stacking

l"45zto 14 t; 45z 1T
0.110

i
0.060
li"Sotolsi 2;

Figure 14.62 Details of the !-stiffener cross-section of the !-stiffened panels employed in the Convair
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
crippling tests (from [14.73])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1108 Composite Structures

Figure 14.63 Dimensions and details of the bladC·SLiflcncd panels used in the Convair crippling tests
(from (14.73))

sequences of (±45AIORI+45A)r for BIA ply proportions of 2. 4 and 6. which represent the
common lay-ups for stiffener designs. The test program of [6.24] and ( 14.76] was described
in Subsection 6.2. I of Chapter 6. The typical no-edge-free crippling Jest loading fixture used
in these studies is depicted in Figure 14.65a and that of the one-edge-free crippling test is
shown in Figure 14 .65b. Spier and Klouman ( 14.76] employed a different approach for the
one-edge-free crippling tests, obtaining the empirical crippling curves indi rectly from tests
on channel secti ons (see Figure I4.66).
As mentioned above, a design methodology simi lar to that used for metallic stiffener
crippling was adopted by Spier and Klouman. Consequently, they proposed semi· empirical
crippling curves, based on Gerard's nondimensional crippling equation (6.7) of Chapter 6,

(b} { c)

Figure 14.64 Crippling failure mode of a short integral 1-stiflcned Convair panel (from [ 14.73]): (a)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

general view, (b) closeup of damage observed from left side, (c) c loseup of damage
Copyright Wiley observed fro m right side
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1109

(a ) (b)

Fi~:ure 14.65 Convair crippling tests loading fixture' ( from (6.2~]1: (a) no-edge-free. (b) one-edge-free

F" IF" - mF••f F" )" (14.16)


Plouing thi\ equation on a log-log scale provide~ linear graphs with (F" /F" ) as the ordinate
and (F•• fF" ) a~ the ab~cissa. The coefficient a is the intercept o n the ordinate at the absci~sa
value of 1.0 and 11 is the slo pe of the straight line on the log-log plot. The requi red data for
plo uing thi~ curve are the ultimate comprcs~ivc strength F"' and the theoretical linearly
clastic buckl ing strength F" , while F" represent~ the crippling stress.
To obtain the crippling st ress of the free-edge flange from the channel crippling test data
of (14.77], the following approximate equation was used:

Fj' "' CP, ;, - F ;,' tb,. )l2b/ ( 14. 17)


Here F;·
is the crippling slrcss of the channel flange. ~~ . is the total crippling load of the
channel. F~· represents the crippling strength of the web obtained from the no-edge free
empirical crippling curve. b. is the width of tlte channel web to the middle surfaces of
flanges. b1 repre~enls the width of the channel flange to the middle surface of the web and
t is the sl.in thickness of the channel cross-\ection.
It wa~ pointed out in (14.76] and (6.24( that Eqs. ( 14. 16) and (14.17) arc no t exact.
However. the potential advantage of these equation~ b that the precise stacking sequence of
the laminate needs not be specified. Hence, they provide the means to obtain crippling curves
for any new :mel as-yet not tested graphite-epoxy 'YMCm.
The crippling test results of (14.76] and 16.24( were used to establish the empirica l non-
dimcnsioml l crippling curves for the no-edge-free and one-edge- free speci mens (Figures
14.67a and b. respectively. reproduced from (6.24]). The parameters cr and 11 of Eq. ( 14. 16)
were se lected to minimize the squared difference between the test values F" and the ca l-
culated ones. F" . The following relation

or-t.•n:• IJ!Io_J .........



. ......
m 14.u. u
~t · .4
I

.' ..'"..
lt, lt,f4
(t. !!
1
q ..

Clo•,.,..,;~ ret~•·•"''"'

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 14.66 Spier and Klouman's channel ' llCCtmcn u•cd for one-edge-free crippling ICSI< (from
Copyright Wiley
[14.76))
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1110 Composite Structures

TEST POINTS
T-J00/5201 A-S/1501 -5

• l•storcst9Dilt ill (!:4s 2Jo 61,


11 l45to 2f.CS11J 4) 1 • l!415z1D 4 11
• (!4SIU 4 11 v (!45ziUzl 1
0 (45/D/·4510z/4SIOz
1-45(03/0J,

T-J00/5281

XS/:15-01-5
I 'znsr
0.1 'OINTS
u TEST POINTS
... T·lOI/5201 A·Sil$11-l
•145N-41/1tl 1, • Jt.45z1111,
!EF 1 • 145/llzl-4111"', •lt.45zll41,
,.)t.4Szllzl,
1.1 1.2 1.40.11.0 Zl4illlf ZO 41 11 41

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
{a)

Figure 14.67 Convair nondimensional empirical crippling versus ultimate compressive stress curves
for arbitrary graphite-epoxy laminates (from r14.74]): (a) no-edge-free, (b) one-edge-free

Test Value- Calculated Value)} 112


{( Calculated Value
SD% - - - - - - - - - - .- - X 100 (14.18)
Total Number of Test Pomts - 1

was used to obtain the standard deviation (SD) about the curves. It was argued in [6.24] that
the empirical crippling curves presented in this manner were general crippling curves, valid
for arbitrary symmetrical laminates. Hence, fewer crippling tests are required to determine
the crippling behavior of specific graphite-epoxy systems.
To establish criteria for short column buckling, the test program of [6.24] included tests
with graphite-epoxy square tubes and !-section columns as crippling and short-column test
specimens. The no-edge-free and one-edge-free crippling data for the graphite-epoxy plates
were used to predict the crippling strength of these specimens, which establish the upper
bound (at zero slenderness ratio) in the Johnson parabolic equation for inelastic buckling of
short metal columns (see for example [14.77] and Subsection 6.1.1 of Chapter 6):
2
(F"Y(L' I p,'")
(14.19)
47r 2 Ec
Here F'" is the short-column buckling stress, L' represents the effective length _LJVC:,
where C is the end-fixity coefficient (3.6 for realistic fixed-ended test specimens), L is the
total length of the specimen including the end supports, and Pmin represents the minimum
radius of gyration of the total cross-section. The tests with the crippling and short-column
square tube and !-section specimens substantiated the use of the Johnson formula (Figure
14.68) in conjunction with the crippling analysis methodology. (Note that the crippling test
results of the square tubes are presented in Figure 14.67a. It is apparent from this figure that
the results agree very well with those obtained from tests of plates with no edge free).
Spier [14.74] attempted to identify the deficiencies leading to the discrepancies between
the classical orthotropic plate buckling predictions and the test results obtained in his former
studies [6.24], [14.74] and [14.76]. Neglect of transverse shearing deformations in the clas-
sical theory, particularly in analyzing web and flange elements in intermediate-length col-
umns and stiffeners that are characterized by low plate width-to-thickness ratio (bit), was
suggested as one of these deficiencies. Spier indicated that this was best exemplified by
studying the plot of the applied load versus the end shortening of the specimen, in cases
where the (bit) ratio is sufficiently small, so that transverse effects considerably reduce the
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1111

900

BOO

700 100
U1
U1
w
0:: 600
"""'
U1
80
z
:::0: 500
:::J
cS MPa KSI Fcc(2 TESTS)
u 60
400
0
u
LL.
300

200 TEST POINTS


• SIDE SQUARE TUBES

100 e POST SQUARE TUBES

0
0 10 20 30 40 50 60 70 80
L'/p min (SLENDERNESS RATIO )

Figure 14.68 Johnson/Euler curves and square-tube crippling, including short-columns test results,
observed in the Convair tests (from [6.24])

critical load. In such cases, the orthotropic predicted buckling load P'' significantly exceeds
both the actual incipient buckling P:' and crippling P'' loads.
In order to resolve this discrepancy, Arnold and Mayers [14.78] presented a materially
nonlinear theory for the failure analysis of laminated plates subjected to compressive loading.
The theoretical model included transverse shear effects and material nonlinearity combined
with the maximum-strain failure criterion. Using this more refined analysis, Arnold and
Mayers demonstrated good correlation between their analytical predictions and the experi-
mental results of Spier et a!. [6.24] and [14.76] for initial buckling, postbuckling stiffness
and crippling failure.
Spier [14.74] pointed out that the use of V-grooves to provide simple supports along the
unloaded edges of the plate introduced another problem. The V-blocks were adjusted by
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

hand to the unloaded edges, allowing minimum clearances, so that the plate could barely
move under hand pressure. Hence, the tests were subject to some human error during their
setup. Furthermore, depending on the magnitude of these clearances, the simple-support
conditions could be satisfied up to the load at which incipient buckling occurs. Because of
the clearance, the unloaded edges were not completely restrained out-of-plane at incipient
buckling and were free to wave in postbuckling. Had the edges been held straight in the
postbuckling region, the number of waves could have been increased and consequently the
wave amplitudes would have been reduced together with the resulting bending stresses. Thus,
higher crippling stresses would have been obtained.
In these buckling tests of the no-edge-free plates another problem emerged. The loading
fixture did not adequately support the upper portion of the V-grooves, and thus buckled
specimens pried the V-grooves apart and falsified the test results. C-clamps near the top were
employed to eliminate this phenomenon.
The ultimate strength of a laminate is affected by its stacking sequence (see [14.79]-
[14.81]). Therefore, the crippling strength, which is associated with the ultimate strength, is
also influenced by the laminate stacking sequence. In laminates of (0. I± 45), configuration,
with all the 0 plies on the outside, and subjected to axial compression, interlaminar shear
and peeling tensile stresses develop in a boundary layer along the unloaded free edges. As
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1112 Composite Structures

pointed out by Spier [14.74], this combination of interlaminar stresses contributed to a deg-
radation in crippling strength. Also, with 0 plies on the outside, the transverse flexural stiff-
ness and strength are at a minimum, which again reduces the crippling strength.
In evaluating the results of his former studies ([6.24], [14.73] and [14.77]), Spier [14.74]
demonstrated that the log-log nondimensional plots (of [14.76]) can be improved and the
test results can be consolidated by replacing (F''/ F") and (Fe" IF'') by (Fe' IF;') and (Fe" I
F~'), respectively, i.e. replacing the theoretical elastic buckling load F'' by the incipient
buckling load F~'. This requires, however, a capability to predict adequately the value of this
incipient load (see Arnold and Mayers, [14.78]), or alternatively to obtain it from tests, as
Spier did.
Spier [14.75] reported the results of an extensive experimental program evaluating the
postbuckling behavior under static loading and the residual strengths after postbuckling fa-
tigue of short graphite-epoxy thin-walled compression members. A wide class of specimens
were included in the program (see Figure 14.69): one-edge-free plates; columns with channel
and Z (web buckling and flange buckling critical), "I"- and hat- (web buckling critical)
sections; and hat-stiffened panels (skin buckling critical). All of these compression members
were fabricated from A-SI3501-6 graphite-epoxy, and most of them had lay-ups of ( ±451
0 3 190)s. Some of the specimens were tested statically to failure, to determine virgin prop-
erties (such as crippling strength) and possible wavelength change in the postbuckling range,
as well as to serve as a reference for assessing the effects of postbuckling fatigue. Other
compression members were first exposed to 105 cycles of compression-compression (R =
0.1) postbuckling fatigue at two cycles per second, and if they survived they were then
statically tested to failure. It is worth noting here that failures during fatigue cycling occurred
in those specimens that experienced wavelength change in the postbuckling range. Hence, a
restriction must be imposed on the maximum fatigue load to avoid such a change in wave-
length, which increases the inner stresses.
The tests with the "successfully fatigued" specimens revealed that none of them experi-
enced any significant reduction in static (crippling) strength with respect to their statically
tested counterparts. Consequently, the crippling loads were essentially "insensitive" to the
compression-compression postbuckling fatigue. The fatigued channel and Z-section inter-
mediate columns, however, sustained lower buckling loads than the corresponding unfatigued
ones.
In addition to the above observations, the test results showed that postbuckling shifted the
neutral axis at midspan of compression members such as channels, and to a lesser extent
hat stiffeners. The much lower crippling stresses experienced by the channel cross-sections,
compared to those carried by the Z-sections, was attributed to the shift of the neutral axis,
which did not occur for a Z-section. Furthermore, this shift was associated with the
development of additional postbuckling bending stresses, which led to premature crippling
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

failure.
An attempt was made in [14.75] to establish a fatigue load criterion based on the following
expression:
pr = P" + (t::,.P1!11P")!:1PI' (14.20)
where P 1, t::,.pf and !::..PI' are the maximum fatigue load, differential postbuckling fatigue load
(Pf - per) and maximum differential postbuckling load (P'' - P"), respectively (see also
Figure 14.70). Using this criterion, the fatigue tests results are presented by plots of the
parameter t::..Pfl /::,pi' versus either the parameter P"l pee or P'' I P'', as shown in Figure 14. 71.
Because of limited test data, the criterion could be only established for the one-edge-free
plates and the Z-section columns, and only for those specimens that experienced no wave-
length change in the postbuckling region.
Crippling tests of channel and Z-section columns, fabricated from AS413502 graphite-
epoxy and having lay-ups of (±451+45190103 )s and (±4510190)"s' where n = 1,2 were
Copyright Wiley
carried out at Virginia Tech in the eighties (see [14.82]). It was observed in the tests that
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1113

~-bw , 3 0 '" -----+1


--.
Flar.ge cnt1cal

~-""CJ
r
br ""' 1 0 rn.
Flange CritiCal

l_
Channef resr sp«:1mrn configurations. Zet test sp«rmen configurations.

(a) (b)

Web
cnl1cal

Web critical

bf ""' 0 7 1n +-----1
1-SI!C(Ion ttsl sptc1men corrfilurarion: ( z. 4510JI90]5, o:cepr

(c ) as shown. (d )
Figure 14.69 Convair graphite-epoxy stiffeners for postbuckling fatigue tests (from [14.75]): (a) chan-
nel specimen, (b) Z-specimen, (c) hat specimen, (d) !-section specimen

just prior to reaching the maximum load, or crippling load, the readings of the axial strain
gages bonded to the surfaces of the columns indicated large compression strains in the
corners of the section that was nearly uniform through the thickness. The axial strain gage
data obtained from the free edges of the flanges indicated severe bending of the postbuckled
flanges. Such free-edge bending of the flanges is associated with through-the-thickness shear
and normal stresses, which may precipitate delamination. Based on these observations, it
seemed likely that failure would initiate either at the corners or at the free edges of the
flanges. In the tests, however, failure occurred so rapidly that the location of failure initiation
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

could be determined neither from visual observations nor from examination of the strain
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1114 Composite Structures

pee

pi
t,pP
1:;>
0
0
--'
per

End- shortening

Figure 14.70 Convair fatigue load criteria (from [14.75])

gage records. (Note that Spier and Klouman observed in their tests with channel sections in
[14.76] that failure originated at the corners.) Examination of the failed specimens revealed
a fracture across the section at midspan. together with free-edge delaminations in the flanges.

In the mid-eighties Reddy and Rehf[eld (see [14.83]-[14.85]) conducted tests with !-
section stiffener specimens, fabricated from C3000/5225 woven graphite-epoxy cloth with
the following stacking sequences: (45/0/ -45) 1, (0/90/0).,., (0/90/0/90)s, (45/0/90/ -45),
and (0/45/-45/90)s. Like Spier et a!. [6.24], [6.25] and [14.73]-[14.76], Reddy et a!.
adopted the metallic stiffener design charts and procedures for sizing their test specimens,
i.e. analyzing local buckling by examining the flange and web elements of the column as
individual one-edge-free and no-edge-free plate elements, respectively. The analysis focused
on identifying the buckling-critical element of the cross section. All specimens were loaded
to failure, and their buckling load, crippling strength, initial postbuckling stiffness and failure
strain data were recorded in the tests. It was shown that the majority of the specimens
validated the preliminary metal design methodology. Furthermore, it was found that failure
was initiated by flange free-edge delamination.
Rehfield and Reddy [14.85] substantiated, through the use of high-speed photography, the
observation that the critical failure mode was indeed flange edge delamination, followed by
simultaneous failure of the flange and the web. They emphasized that this finding implied
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

o.sr------------------,
0 Fail•d

"~~
e 100. 000 eye les on failure

0.6 Q5-F18,410cycles

a.
0. Tentative fatigue limitation
~ 0.4 6 F ·2
0. 7 F I / curve for one-edge-tree plates
<l

0.2

·~- F1 e

0.0 0.2 0.4 0.6 0. 8 1.0

5.0 2.5 1.67 I. 25 1.0

Copyright Wiley
Figure 14.71
Provided by IHS Markit under license with WILEY Convair one-edge-free postbuckling fatigue
Licensee=McDermott test data (from
Inc - India/8215328006, User=G, [14.75])
Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1115

that the usual material strength criteria might be misleading here and a criterion for free-
edge delamination would be required for specimen sizing. Furthermore, in order to predict
crippling reliably, multiple failure modes and criteria must be considered.
Naturally, the flange free-edge delamination failures experienced with the stiffeners in
[14.82]-[14.85] motivated a search for ways to inhibit this failure mode in order to yield
higher stiffener ultimate strength. This approach was taken by Causbie and Lagace at MIT
[14.86], who conducted tests with angle and channel sections fabricated from a thermoplastic
material, graphite/ APC-2 PEEK. They realized that since use of the thermoplastic matrix
did not change the overall component moduli that were responsible for initial buckling, the
buckling capacity of the stiffeners would not change significantly as compared to stiffeners
manufactured from standard thermosetting graphite-epoxy. However, they assumed that due
to the increased interlaminar toughness of a thermoplastic material, the stiffener sections
would delay crippling in a delamination mode and would exhibit higher crippling strength
and different failure modes. As a matter of fact, no apparent difference in crippling stress
and failure mode was found from a comparison of the test results of the PEEK specimens
with those obtained on similar sections made of T300/ 5208 graphite-epoxy. Again, crippling
failure was attributed to flange delamination.
Local buckling, postbuckling and crippling response of channels, square tubes and square
tubes attached to flat sections of skin (representing a portion of a stiffened panel), all fab-
ricated from AS4/3501-6 graphite-epoxy, were also investigated by experiment and finite-
element analysis at MIT (see [14.87]). It was observed that failure in the channels and tubes
was initiated by ply delamination in the corners. Delamination was again found to be the
mode of failure for the segments of tube-stiffened panels, but the location of damage initi-
ation was uncertain. Their analysis, which was based on classical lamination theory, predicted
the local buckling loads and modes reasonably well, but it failed to estimate the failure loads
from the postbuckling solutions.
A further experimental and analytical investigation of the local buckling, postbuckling and
crippling behavior of channel Z-, 1- and J-section stitieners manufactured from AS4/3502
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

graphite-epoxy unidirectional tape was carried out at NASA Langley in the late eighties (see
[14.71] and [14.72]). This study was a direct extension of the research at VIP (presented in
[14.82]). Being aware of the complexity and abruptness of the crippling process in laminated
composite thin-walled columns, the NASA Langley investigation aimed at providing further
insight into the local buckling and crippling of composite compression members by pursuing
a combined experimental and analytical investigation. As shown, the advantage of such a
combined approach was that the analytical model could be validated by the test results. Then
this verified analysis methodology could be employed to provide further insight into the
crippling problem, which was not directly provided by the experiments. Consequently, the
objectives of the NASA Langley study (see Refs. [14.71] and [14.72]) were to conduct
fundamental experiments for determination of the difference between composite and metallic
stiffener crippling; assess the effects of cross-section shape, stiffener lay-up, flange width,
wall thickness, comer radius and stiffener length on the local buckling, postbuckling and
crippling behavior of graphite-epoxy stiffeners; supplement the experimental study with com-
putations by the two-dimensional finite element shell analysis code STAGS C-1 [ 14. 88] to
create computational models for the full range of response (pre- and postbuckling) of selected
test specimens and develop a postprocessor to the STAGS C-1 code capable of estimating
the interlaminar stresses; and obtain the in-plane stress data from STAGS C-1 for each
analytical model, either at the load corresponding to occurrence of first damage event, ob-
served during testing of the physical specimen, or at the experienced crippling load if no
evidence of damage was found. These stress data were then used to evaluate for failure
according to the Tsai-Hill and Tsai-Wu (tensor polynomial) and the Hashin four-failure
criteria (see [14.89]).
The experimental observations indicated that damage initiated in a number of modes, all
of which were associated with either delamination in some part of the specimen or local
Copyright Wiley
Providedmaterial strength
by IHS Markit under failure
license with WILEY in a corner of the specimen. It was Inc
Licensee=McDermott found that theUser=G,
- India/8215328006, flange width-to-
Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1116 Composite Structures

thickness ratio affected the mode of damage initiation and that the inner comer radius, r,,
strongly affected the crippling strength in the I- and J-section specimens (see Figure 14.72).
Buckling predictions and experimental results were in excellent agreement, however the
correlation degraded with increase in postbuckling load (see Figure 14.73). This discrepancy
was attributed to neglect of transverse shearing deformations in the analysis by STAGS, the
occurrence of a damage event during the testing of the specimen at about 10,000 lb (44.5
kN) and inadequate modeling of the specimen corners. Use of in-plane stress data from the

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
STAGS C-1 analysis at the load corresponding to the first damage event revealed that
Hashin's compressive fiber mode failure criterion correlated reasonably well with the exper-
imental observations of damage initiation.

Frostig et al. [14.90] presented a joint Technion Aerospace Structures Laboratory and
Israel Aircraft Industries (IAI) intensive experimental study on postbuckling and crippling
behavior of selected laminated AS4/3502 graphite-epoxy I- and J-section stiffeners subjected
to static and cyclic compression loading. The I- and J-stitleners were chosen because they
are commonly used in the design of stiffened panels. Each group consisted of "long" spec-
imens for column-buckling studies and "short" ones for defining the crippling load capacity
of the stiffeners. Furthermore, the "long" stiffeners had an attached skin, 90 mm wide, to
evaluate the feasibility of predicting the behavior and failure characteristics of a stiffened
complex structure based on its constituent behavior, i.e. the equivalent section concept con-
sisting of a stiffener and equivalent-width skin. In the Technion-IAI tests [14.90], the attached
skin width equaled the typical effective width of the panels tested by Segal, Siton and Weller
within the framework of this program [14.91].
It was observed in the tests that the overall behavior usually involved initial buckling of
the attached skin and beam-column buckling of the stiffener section. Furthermore, cyclic
loading had favorable effects on failure loads, i.e. the cycled specimens always experienced

f--br bf --1 r-- bf -l


iS)
'• I'•
13)
T
b.,.
'•
@

...
'•
bt~
ID 1~bto 12)

bra
<D
__j
I-SECTION J-SECTION

Figure 14.72 Details of !-section and J-section stiffeners used in Bonanni et al.'s crippling tests (from
[14.72])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1117

12000

10000 o•
oo
0
0
0
8000 0
0
0
0

~ 6000
11.' 0 Experiment
--STAGS analysis
4000 • Failure

STAGS

Oil-~---L---L--L---L---U
0.00 0.01 0.02 O.Q3 0.04 0.05 0.06
U,IN

Figure 14.73 Typical comparison of load-end shortening data obtained in Bonanni et al.'s experiments
with STAGS prediction (from [14.72])

failure loads higher than those sustained by the statically loaded reference specimens. Failure
of all specimens, regardless of their shape and loading, static or cyclic succeeded by static,
was associated with cap delamination together with crack and fiber breakage in the attached
skin.
The experimental studies were accompanied by analytical predictions, based on beam-
column buckling of an equivalent cross-section, and an in-house-developed program
BPCOMP [14.92] and [14.93]. BPCOMP models any fiat, longitudinally stiffened panel
made of plate strips and connected transversely through equilibrium and compatibility con-

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
ditions and allows large deflections and small rotations. Hence, it provides a tool for analysis
of distorted cross-sections, which is the case when the applied load approaches failure. It
was shown that predictions of beam-column buckling based on equivalent sections are sat-
isfactory within engineering accuracy, regardless of section shape. The BPCOMP results
were found to be in good agreement with the experimental observations (see [14.90]). It was
also indicated that the predictions could be improved by allowing load redistribution during
the loading sequence, in compliance with a uniform end-shortening condition.

14.2.2 Stiffened Panels


Buckling of stiffened metallic panels and the complex modes associated with their buckling
are discussed in detail in Chapter 12. Buckling of laminated composite stiffened panels is,
however, a considerably more complicated problem. This is due to the presence of bending-
twisting and/ or bending-stretching coupling, inherent to various laminate lay-ups, which
were discussed earlier in Subsection 14.1.1 of this chapter. These couplings further compli-
cate the buckling modes of a laminated composite stiffened panel. As in metallic stiffened
panels, the modes are generally either local or overall, depending on the wavelength of the
bucking patterns.
Williams and Stein [14.94] showed that in open-section stiffened panels local, twisting
and column buckling modes should be anticipated (see Figure 14.74). These modes indicate
that either relatively independent or strongly coupled displacements of the plate skin sections
and web-stiffening elements are involved in the buckling process of the stiffened panel.
Methods for analysis of the local buckling between stiffeners and the overall general
instability of the panel were reviewed by Leissa [14.5]. Leissa indicated that numerous
studies, theoretical and experimental, and a large amount of data on the buckling and post-
buckling behavior of stiffened laminated composite panels are available, and he also cited
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1118 Composite Structures

--- --- =r----_-_-


___ I
T------- -
T
I 1
'
I 'I
I '
I I
I
J
LOCAL/SKIN LOCAU SKIN AND WEB TWISTING/SKIN AND WEB

TTT
TWISTING/WEB ROLLING

Figure 14.74
COLUMN/WEB ROLLING COLUMN

Buckling modes of open-section stiffened panels (from [14.94])

the most widely used computer codes for analyzing the behavior and buckling of stiffened
composite panels: BUCLASP, NASTRAN, STAGS, VIPASA and the relevant literature
sources. Leissa indicated that in many instances these codes are applied as parts of larger
optimization procedures for the efficient design of the stiffened panels. It should be noted
that the optimization analysis of stiffened laminated composite panels involves a considerably
larger number of design parameters than that of metal ones.
As mentioned earlier, weight-efficient structural designs can be achieved by taking advan-
tage of the postbuckled strength of stiffened panels. In addition to the computer codes enu-
merated above, others are available for the analysis of the buckling and postbuckling behavior
of stiffened panels. (See for example the evaluation of these codes and many others in
[14.95]). It appears, however, that they do not predict satisfactorily the behavior of the panels
in the "deep" postbuckling range. This is due to the strongly nonlinear response of the
panels and to the failure modes inherent to laminated composite stiffened panels. These
modes, such as delamination in the skin, stiffener/web separation evolving through the thick-
ness or interply complex three-dimensional stress field that exists at the source of failure and
compression failure in the skin, are not appropriately accounted for in the analytical model
on which the codes are based. Furthermore, even though today advanced analytical tools are
available, such as global finite element calculations, the question still arises whether calcu-
lation of the failure load can be adequately performed with the state-of-the-art finite-element
modeling combined with fracture and damage mechanisms.
To overcome these doubts, the design methodology for the postbuckled stiffened panels,
considering their failure characteristics, is therefore predominantly semi-empirical (see also
Subsection 14.2.le on crippling behavior). As such, its verification strongly relies on the
existence of a sizable database provided by numerous tests and requires its continuous up-
dating. Furthermore, the method of analysis has to be as general as possible. This requires
that the test programs be conducted with generic structural components that allow systematic
variation of their structural parameters. These tests should provide data on all phases of
structural behavior: skin buckling, overall buckling, deformation behavior, as well as other
characteristics, structural and material, peculiar to composite construction.
A discussion of typical experimental studies on buckling and postbuckling behavior of
laminated composite stiffened panels under axial compression, shear and combined loading
follows.

a. Axial Compression Tests


Further to Spier's earlier tests [14.73], discussed in Subsection 14.2.le, an additional inves-
tigation by Spier and Anderson was reported in [14.96], where a semi-empirical analysis
method for predicting the buckling and crippling loads of carbon I epoxy fabric blade stiffened
panels in compression was developed. The approach in this study was based on Spier's earlier
--`,`,`````,`,````,,``,

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1119

procedures (see [14.73]-[14.76] and [6.24]). In order to verify the proposed method, it was
employed to design a stiffened panel, which was fabricated and tested. The panel configu-
ration is shown in Figure 14.75. It consists of three blade stiffeners that were cocured to the
skin. The entire panel was fabricated from Hercules AS4/3501-5A carbon/epoxy fabric,
except for the C/Ep tow used at the flange-blade intersection (see Figure 14.75b), which
was introduced to provide structural integrity at the joint and add significant torsional stiff-
ness at the blades.
The panel, after being A- and C-scanned, was machined and assembled with potted alu-
minum end channels. The end surfaces were then ground parallel within 0.001 in. (0.025
mm). Split rigid steel tubes were attached to the unloaded edges to provide simple support
boundary conditions and simulate continuous supports to the stiffeners, as though they were
in a much wider stiffened panel (Figure 14.75a). Thus, analysis of only the middle stiffener
(and then application of the results to all stiffeners) was sufficient to determine the load
capacity of the whole panel. To monitor the response of the panel, strain gages were bonded
to its skin and middle stiffener (Figure 14.75a). These gages were used to assess the uni-
formity of axial strain prior to buckling as well as to detect incipient buckling and monitor
the postbuckling behavior up to failure.
It was found that the predictions yielded by the analysis were conservative (see [ 14.96]).
The buckling strength and crippling load predictions were 12 percent and 17 percent lower,
respectively, than those observed in the test.
Extensive studies on buckling under axial compression of laminated stiffened composite
panels were also carried out at NASA Langley. Detailed discussions of these investigations
can be found in [14.42], [14.94] and [14.97]-[14.105]. The studies reported by Williams
and Mikulas [14.97] and Mikulas, Bush and Rhodes [14.98] represent an effort to establish
an experimentally verified weight-strength database for various generic structural compo-
nents, which will provide the reference standards for assessing component performance.
To accomplish their objectives, Mikulas et al. [ 14.97] and [14.98], selected hat-stiffened
and open corrugated configurations (see Figure 14.76). Compared with metallic designs,
those with composite material involve a much larger number of design parameters, and
therefore the general approach taken in the program differed from that employed earlier by
NACA for metal panels [14.1 06]-[14.110]. Instead of thousands of panels with parametri-
cally varied dimensions being tested to develop design allowables and efficiency charts,
automated design was employed for the determination of efficient panel cross-sections for
the experimental program. To seek the minimum weight panel proportions (cross-sections),
the mathematical programming design method of [ 14.111] was used, subject to the following
constraints (not to be exceeded):

1. The local buckling load of the skin and stiffeners


2. The wide-column Euler buckling load
3. A prescribed allowable axial strain
4. Geometric proportions to lie within the limits prescribed by practical or manufacturing
considerations

For each cross-sectional geometry selected, short specimens (410 mm long), critical in
local buckling, and longer specimens (1520 mm long), critical in wide-column Euler buck-
ling, were built and tested.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

The panels, 23 short ones and 6 long ones, were fabricated from Thorne] 300 graphite/
Narmco 5208 in either tape or fabric form. The short panels were axially compressed in a
300 kips ( 1.33 MN) capacity hydraulic test machine and the long Euler buckling specimens
were loaded in a 1.2 million lb (5.33 MN) capacity test machine. Strain gages were used to
monitor axial and transverse strains of each of the hat-stiffener elements. Direct current
differential transformers were used to measure the cross-head movement and lateral displace-
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1120 Composite Structures

~ ,--- I :'

T-1, • II
I I Pott1ng I
I I I
;i I i r

i
191~
I 9N
I
I
I
I 10N I
I
I
I

,, I I I i
2N 1N

-~-..±--e-n '::l:
I I I
" I
'
.70
-r- I
Aluminum Channel (Typ)

Spht Steel Tube (Typ) ~


(Held on w1th Tape, --.............(()
Not Shown)
(a)

c(45". o·. 45". 0° 5 ~5·. o·. 45")~

1----~:~--
1
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

' 0", 0", 45")]

(b) o•

Figure 14.75 Spier and Anderson's blade-stiffened test panels (from [14.96]): (a) dimensions and
boundary details, (b) details of flange-blade intersection and blade and skin lay-ups

ment of the panel. The strains, displacements and applied compressive load were recorded
on magnetic tape. Selected measurements were monitored during the test on an oscilloscope.
Buckling patterns of the panel as they developed during loading were observed with the
moire method.
Specimen ends were potted into a l-in. (25.4-mm) thick block of epoxy and the ends were
ground flat and parallel. Unifonn loading of the panel was ensured by preloading the spec-
imen to a fraction of the ultimate load and adjusting the cross-head platen until all axial
strain gages on the stiffeners displayed approximately the same reading.

=±8PLIES
- 0 ° PLIES

~0
CONFIGURATION A CONFIGURATION B

~ CONFIGURATION C
Copyright Wiley
Provided by IHS Markit under license with WILEY Figure 14.76 Mikulas et al.'s compressionLicensee=McDermott
panel configuration options User=G,
Inc - India/8215328006, (fromBoopathi
[14.97])
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1121

The test results were compared with analytical predictions obtained with BUCLASP 2
[14.112] and [14.113] and BUCLAP 2 [14.114]. In [14.97] and [14.98] it was shown that
graphite-epoxy hat-stiffened compression panels theoretically present weight savings on the
order of 50 percent when compared with similar aluminum compression panels. In the ex-
periments, however, weight savings of only 32-42 percent were obtained. It was suggested
that insignificant alterations incorporated in the panels to improve fabricability or preliminary
design could significantly reduce the theoretical structural weight efficiency. Therefore, re-
alization of the full 50 percent weight -saving potential would require close attention to design
and fabrication details (laminate material properties, material thickness variations, panel ex-
tensional stiffness, fabrication of thermally induced initial curvatures and design details such
as overlapping). The open corrugation graphite-epoxy designs (configuration C in Figure
14.76) were found to be 20 percent lighter than the graphite-epoxy hat-stiffened designs.
The weight strength performances of the panels tested, together with the data points from
2000 NACA aluminum panels [14.106]-[14.110] are presented in Figure 14.77, which em-
phasizes the improved structural efficiency of composite panels.
Furthermore, it was pointed out that during the course of the investigations important
structural phenomena, which were not considered in the design synthesis method of [ 14.111],
were observed. The phenomena, which were revealed by either the experiments or by com-
parisons with the BUCLASP-2 and BUCLAP 2 predictions, were anisotropic effects, which
are important for relatively thin ply laminates; a column transverse shearing effect, which is
important for certain combinations of load ranges and vertical web thicknesses; local buck-
ling boundary conditions other than simple supports that offer more restraint; and residual

w
ALe'
LB --+-~+.2CL.:..--c__-./-

IN3 -----
'IJ>,I -
~--_.!/~"-/'"'"\___________ _
. ---· - - - -

I i i ~
:_I~,:C ---- I_;_'1--i ---~-----,--
L;_c_ _ _ ~--- PR[SEHJ THtORY ·UNCONSTRAINED ,,
, :':

' .
·+=+--··
..
' ---PRESENTTHEORYCONSTRAIN!St 1.t 1 .L,JO

-I--~ o
::::::
EXPERIMENTAL . GRAPHITE;EPOXY HIT STIFFENED DESIGN-
I I (!15' WEBI _.:::
A I 1 (!5t WEB]
b. A I (fABRIC WEBI
<> A· 5 (SEE liSLE 1111
o B · 2 (SEE TABLE 1111
o C · I (OPEN COiRUGITIDNI
''I I I "
I. ~ ~ I
lo' IL-~~~_u~~~~~+-~~·~:~1~6~;~·~a~·o~o~~~~,~.ct;~;~;~~~ooo·o
< ' • I I I\ I

10
Nx Le
Le IH2
Figure 14.77 NASA Langley comparison of structural efficiencies of graphite-epoxy and aluminum
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
compression panels (from [14.97])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1122 Composite Structures

thermal strains and panel warping stemming from the different coefficients of thermal ex-
pansion in adjacent elements due to different laminate lay-ups. These strains, which were
induced during the cool-down phase of the cure cycle, lead to panel warping and may have
both beneficial and detrimental effects. It was emphasized that smaller weight penalties
would result if these either degrading or enhancing effects were to be taken into account
earlier in the design cycle.
Additional experimental studies on composite hat-stiffened panels were carried out in the
late seventies to early nineties (see [14.115]-[14.119]). Vestergen and Knutsson [14.1151
studied experimentally the buckling and postbuckling behavior under axial compression of
the 5208/T300 graphite-epoxy hat stiffened panels and evaluated the dependence of the
buckling and postbuckling characteristics on the lay-up configuration. The experimental re-
sults were compared with theoretical predictions obtained by the STAGSC-1 code [14.88]
and reasonably good agreement was reported. It was pointed out (in [14.115], that the buck-
ling mode of the panels could be estimated with high precision with a bifurcation-type
analysis. However, for predicting the buckling load, a nonlinear-type analysis was preferable.
To estimate the load within reasonable tolerances, the results were found to be very much
dependent upon the model employed in the analysis and on the resemblance between the
actual boundary conditions and shape of imperfections of the test specimens and of those
introduced in the theoretical analysis. This also appeared to be the case with the failure load
when employing a failure criterion, such as the Tsai-failure criterion, in the analysis.
Extensive experimental studies on buckling and postbuckling of advanced composite-hat-
stiffened graphite-epoxy panels were conducted by Romeo at the Department of Aerospace
Engineering, Turin Polytechnic, Italy, in the mid-eighties. These tests provided data for cor-
relation with the theoretical analysis developed by Romeo (see [14.116] and [14.117]). The
theoretical model presented in these studies was based on a wide-column theory to predict
overall buckling and on the orthotropic buckling equations to predict local buckling. Ade-
quate correlation between analytical and experimental results was observed when the Euler
buckling mode was critical. When the local buckling mode was critical, buckling occurred

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
at lower strain values than the predicted ones.
The test program carried out by Falzon and Steven at the University of Sydney (see
[ 14.118]) represents an effort to study the effects of different manufacturing techniques of
hat-stiffened panels on their structural integrity in buckling and postbuckling when subjected
to axial compression. Five panels were manufactured using different manufacturing tech-
niques. Linear finite-element analysis was performed to compare the experimental results
with analytical predictions. The comparison emphasized the need for more accurate repre-
sentation of the edge conditions and local panel imperfections, as well as material data, in
the finite-element model for reliable prediction of the panel behavior.
In [ 14.119] the program PAN OPT [ 14.120] was employed to design CFRP panels with
different stringer concepts, "L", "J" and hat, and to investigate their response to compressive
load, in particular their buckling behavior. Good correlation between the predicted and the
experimental buckling load was reported. The fabrication techniques developed at NLR in
The Netherlands for manufacturing of the panels were also described in [14.119]. It was
shown that with these techniques CFRP compression panels of high quality could be pro-
duced.
Williams and Stein [14.94] at NASA Langley investigated the buckling behavior and
structural efficiency of open-section J and blade-stiffened graphite-epoxy compression panels
in order to assess how well experimental data could be correlated with buckling analysis for
these structural configurations. Furthermore, it was indicated in [14.98] that compared with
closed-section hat-stiffened panels (see [14.97]), the J- and blade-stiffened composite panels
offer attractive design alternatives for use in commercial aircraft wings. Therefore, the
study in [14.94] was also concerned with design and analysis procedures for the minimum
weight of open-section stiffened panels subject to a prescribed axial compression
load.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1123

The J- and blade-stiffened panels selected for the comparative study (sec Figure 14.78)
were constructed usi ng unidirectional fi lamentary graphite-epoxy tape and the integral as-
sembly was vacuum bagged and autoclave cured. The J-conliguration chosen fo r testing had
minimum weight panel dimensions. complying with buckling and strength con~trnints. weight
index WIAL = 5.70 kg/m ' and load index NIL= 0.689 MPa. The panels had two stH'feners,
which were 277.406 and 760 nm1long. with an axial qifTne'>'> of22.5 MN. A 760-mm long
panel. simply supponed along its loaded edges. was designed to buckle at an axial strain of
0.003. The blade con figuration chosen. though a minimum-weight design, incl uded additional
st iffness requirements typical of mediu m-sized commercial aircraft wing structures (WI AL
~ 11.3 kg/m 1) and (NIL = 2.07 MPa). T he panels had three stiffeners, which were 183.
277. 307 and 760 mm long with their axial stiffness designed to buckle at an axia l_;'!train of
0.0036 for a 760-mm long simply supponed specimen (note that here W, A, L and N denote
the weight. cross-section area. length and buckling stress rc~uhant of the panel. respectively).
To facilitate flat-end testing. the ends of the panel were potted in a 25-mm thick epoxy
compound that provided in essence clamped boundary conditions. The unloaded lateral edges
of the panels were unsupported. The response of the panel was ohtained fro m strain gage
measurements and from displacement readings of the ~ross-head platens and of selected
points of the surface. The moire technique was employed to display the deflectio n patterns
of the surface of the panel.
ll1e experimental re!>uhs for the two panel types were compared with two sets of analytical
predictions. one yielded by the STAGS C- l computer code [ 14.88] for the clamped loaded
end and the other obtai ned by the BUCLASP-2 code 114. 11 2] and [14.1131 fo r si mple
support boundary conditions.
From comparison of the analytical predictions wi th the experimental observations. buck-
ling strains and mode shapes, it was concluded in [ l-1.94] that in spite of some nonlinear
behavior observed during the tests. adequate correlation~ with analysis was obtained. This
juqified the u5e of the relatively simple linear. thin-plate buckling analysis in a mirtimum-
weight design synthe!>is program for J and blade configurations. Also. it was pointed out
that the buckling behavior of short-length panels wa~ affected by the flat -end boundary
conditions used in testing. Hence, to determi ne analytica lly the bu~kl ing behavior. the pre-
buckling stress field associated with clamped edges. including tero in-plane di spla~ements .
!>hould be included in the analysis.

L___!_
-12,7-
650 310
-119 -
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 14.78 William' and Stein's graphite-epoxy J- and blade-<ufrcncd u.-st specimen< (from [ 14.94])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1124 Composite Structures

Studies on composite blade-stiffened panels were also included in [14.116], [14.117] and
[14.119], which have already been discussed. In [14.119] investigations on composite J-
stiffened panels were reported as well. In general, the discussions about the experimental
observations and their correlation with the theoretical predictions, pertinent to the hat-
stiffened panels, apply as well to the test and analytical results of the J and blade-stiffened
panels investigated. It should, however, be noted that in the case of the blade-stiffened panels
studied in [14.116] and [14.117], adequate correlation with test results was obtained when
the torsional buckling mode was critical. Furthermore, it was pointed out in [14.119] that
the actual failure loads sustained by the J and blade-stiffened panels were higher than their
design loads.

Further studies on composite blade-stiffened panels were reported in [14.121], [14.126],


[8.165] and [8.166]; and additional investigations on J-stitiened panels were discussed in
[8.125], [14.91] and [14.126]. Frentin, Alesi and Scott [14.121] and Scott and Rees [14.122]
presented the results of test programs conducted at the Cooperative Research Center for
Aerospace Structures (CRC-AS), Melbourne, Australia. Here, testing approaches different
from that employed in all of the aforementioned axial compression tests (where the loaded
ends of the panels were potted in a relatively soft material, machined parallel and loaded
between the loading platens of the test machine) were employed. In the tests of [14.121] the
rig presented in Figure 14.79 was used to introduce the compression load into the test panels.
In this apparatus, the experiments with the blade-stiffened panels of Figure 14.80 were
performed by utilizing a four-point bending rig consisting of an aluminum three-bay box
beam structure (used as the bend specimen) supported in a steel rig. The load jacks and
reaction points were attached to the rig, thus providing a self-reacting system. A detachable
skin in the middle of the bay of the box beam allowed the test panels to be interchanged
via a mechanical fastening system.
It is apparent from Figure 14.80 that the test panels in the test program of [ 14.121] ditier
from those tested in the earlier studies. Consequently, they were exposed to a different type
of loading. In the earlier investigations [14.94], [14.96], [14.97]-[14.105] and [14.115]-
[14.119]), the skin and stringers of the test panels were uniformly loaded, whereas due to
the stiffener run-out of the test panels of [14.121], they experienced a bending-buckling
response. As a matter of fact, these types of stiffeners are common in real structural elements
but did not receive enough attention because of their complicated bending-buckling behavior.
The Australian CRC-AS test panels of [14.121] were cocured from unidirectional Ciba
Geigy T300/914C carbon fiber-epoxy preimpregnated tape. Their laminate design was con-
strained by the criterion that the stacking sequence for the stiffener be symmetric. Since the

_s_

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
AI( dimensions in milimeters

Figure 14.79 Trentin et al.'s box beam test rig (from [14.121])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1125

SIClH SECTION I'Lr 0/IJEHTATIOH


sa (4S/-451'}()(0~
Ul 4S/-4~4ji-4S
114 112 S9 4S/-4~-4ji4S
• Ul • Ul • ~

•• U9
SIO
4S/-4S/9010v'!J0(4ji-4S
(4"-4.5IJOIO,)c
UIO 4S/-4.519010.19(Y4S/-4S
mFFENE/t SEcrtOH
Sl6 I (4S/-4j/'9()/Q.Js
Sll6 l HS/45190/!),)s

HtU«ri41 propcrfi11: "'- • IJO GPa


E,-•4.6S GPa
GLr • 4.6S GPa
~~~--- IILr•O.JS

Figure 14.80 Trentin et al.'s blade-stiffened test panels-dimensions and lay-ups (from [14.121])

stringers were manufactured by extending the top three skin plies of the sides of the blade
stiffener, to allow them to become efficiently integral and comply with this symmetry re-
quirement, their construction departed from conventional symmetrical construction used in
most of the studies. This type of construction led to skin lay-ups that alternated between
being symmetric and asymmetric across the panel between each pair of stringers (see Figure
14.80).
The investigation presented in [14.121] evaluated the capabilities of the advanced finite-
element codes MSC/NASTRAN and COSMIC/STAGS to cope with the complicated bend-
ing-buckling response associated with this type of test panel. On the basis of comparing the
test results with the analytical predictions, it was concluded that the onset of buckling, as
well as out-of-plane buckling and postbuckling of the stiffened panels, could be accurately
determined with the two-dimensional finite element programs (see Figure 14.81), though
there was a lack of correlation in the deep postbuckling regime. It was suggested that this

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
disagreement could be attributed to the fact that the four-point bending test introduced cur-
vature, as well as compression, on the unloaded edges of the panel.
In the other program carried out at the CRC-AS by Scott and Rees [14.122] the com-
pression test rig depicted in Figure 14.82 was developed for studying the performance of
postbuckling cocured, blade and !-stiffened composite panels. In this compression rig the
load was introduced into the panel by two end fittings bolted to the short edges of the panel
(instead of the common practice of potting the loaded edges with a relatively soft material).
The unloaded long edges of the panels were supported by edge members, which were guided

200 ---{}-- EXPERIMENT


- - NASTRAN

E1so
E
z
"0 100
.3
50

0~-~~~~~~-~-~--L-~
-6000 -5000 -4000 -3000 -2000 -1000 1000 2000
Strain (~e)

Figure 14.81 Trentin et al.'s typical comparison of test data with NASTRAN predictions (from
[14.121])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1126 Composite Structures

end fixture for axial test machine


bolted to panel short edge

side frame guided in


-------- end fiXture

side frame bolted to


panel edge
- bo~s finger tight
4
holes in panel overs1ze

-~- 5 blade test panel

Figure 14.82 CRC-AS compression test rig (from [14.122])

in end fittings. To allow uninterrupted compressive deformation of the panel, these edge
members were lightly clamped to the panel long edges and the holes along the panel long
edges, through which the edge members were fastened, were drilled 1.5 mm oversize.
The objective of this Australian study (see [14.122]) was to compare the performance of
the thin-skinned, postbuckling stiffened composite panels with that of conventional non-
buckling sandwich-type composite designs. The postbuckling panels were therefore designed
to the same criteria and load cases as the sandwich panel, with the exception that local skin
buckling was permitted. The configuration of the blade-stiffened panels tested in [14.122] is
shown in Figure 14.83. This five-blade stiffener configuration was selected with a detailed
finite-element analysis using MSC/NASTRAN. It is seen in Figure 14.83 that, as in the test
panels of [14.121], the specimens in these investigations were characterized by a run-out of
the stiffeners at the panel loaded edges. Hence, a bending-buckling response was also ex-
perienced with these panels.

- . . - . . l Exploded sectfon view


: h1 M ~ : of top ply IQyup
L~_J
20 ply end
laminate '125mm

79mm

(j) SKIN (-4514519010 )5


(D SKIN (-4514519010101901-45145)
G) STIFFENER (-4514519010101451--4510) 5
G) STIFFENER (451-45190/0101-45145/0)s
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 14.83 CRC-AS blade-stiffened test panels-ply configuration (from [14.122])


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1127

Furthermore, as in the other CRC-AS tests [14.121], it was decided that in the case of the
blade-stiffened panels the stiffener lay-up should be symmetrical because in postbuckling
designs the stability of the stiffeners is more important than that of the skin. Consequently,
this led also here to the design of the panel shown in Figure 14.83, in which the stiffener
laminate is symmetric and consists of 16 plies (sec also Figure 14.84a), whereas the skin
alternates between symmetric and asymmetric eight-ply quasi-isotropic lay-ups. It should be
noted that in Figure 14.83 the loaded edges laminate was 20 plies and that of the unloaded
side edges was I 2 plies. These additional plies terminated in pairs in the 20o stiffener run-
outs. Furthermore, the design of the stiffener flange in Figure 14.84a is integral to the skin.
The blade-stiffened panel of [14.122] was modified to form the !-stiffened test panel.
Recognizing in this case that the top flange of the stiffener was stability critical in postbuck-
Iing resulted in a symmetric flange and asymmetric web lay-up of the !-stiffener (Figure
14.84b).
Both blade and !-stiffened panel designs in [14.122] sustained higher specific strength than
their sandwich counterparts. However, the blade-stiffened panels achieved the design ultimate
load in compression, whereas the !-stiffened panels did not carry their designed ultimate
compression load. As in the earlier test series [14.121], the onset of buckling and out-of-
plane displacement patterns was accurately predicted using the finite-element method. Cor-
relation between the test results and analytical predictions, however, was not as good in the
deep postbuckling regime. Furthermore, use of the maximum strain initial ply failure criteria
in the analysis failed to predict adequately the strength of the panels. It was suggested that
this was due to interlaminar shear stresses and through thickness normal stresses, which

45 45 9g41~1 \4~1~151 F.l50190 45 45


matenal axrs system \ \ I II II 1/ /

;~===:~ IL=~~
o
%~
"
" "" _o
~%
45 /~ tZ mote-nal ax•s for sKm
~"
45
45 ..-J
y
(a)

45

~~~
o-..:::::::~~§~gl1J
45 45
45 45 45 45
45
~0---- 45
45-
45 ........-- Y ....J' Z material axis system for skin and top flange.
(b)

Figure 14.84 Lay-up detail of integral stiffeners of CRC-AS test panels (from [14.122]): (a) blade
stiffener, (b) !-stiffener
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1128 Composite Structures

might have been significant and were not accoumed for in the analysis. Also, the postbuck-
li ng behavior was found to be sensi tive to the panel edge conditions. thus requiring their
actual modeling in the finite-element analysis.

Addi tio nal tests with graphite-epoxy blade-stiffened panels were carried ou t by Romeo
and Frulla at the Departmer.t of Aerospace E ngineering. Turin Polytechnic, Italy [ 14. 123].
In this investigation. in contrast to the previous Torino studies [ 14.116] and [ 14.117]. the
combi ned loading testing machine used by the Torino group to load unstiffcned panels under
combined loading (see Figure 14.58 and 11 4.67]- [ 14.70]) was employed to load the blade-
stiffened panels (see Figure 14.85), bo th in axial compression and under combined loading.
As can be seen in the figure. the test panels had fo ur blade sti ffeners. The compression test
was conducted with a panel. the thicknesses of ski n and blade-stiffeners of which were

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
relatively thick as compared to those of the earlier test programs. and which had a 26-ply
ski n lay-up (+45~10~1 +45/90'10,), and a 38-ply stiffener lay-up (+45/0.~,1 +45 2 ). The
loaded ends of the panel were potted (in epoxy) and encased in an aluminum frame, thus
providi ng nominal clamped boundaJ)' conditions. while its un loaded edges were simply sup-
ported.
The test of this program was conducted in order to verify the adequacy of the analysis
and associated computer code POBUCK [14.70], and to evaluate the postbuckJi ng behavior
of graphite-epoxy stiffened panels contai ni ng initial geometric imperfections, and in the
presence of load eccentricity. Good correlation between the test results and analytical pre-
dictions was reported in [ 14. 123]. It was also shown that the test panel was capable of
sustaini ng loads significantly in excess of the local skin buckling load.

figure 14.85 Blade-stiffened panel installed in the Turin combined loading testing machine (from
Copyright Wiley (14. 123))
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1129

Wiggenraad [ 14.124] presented the results of an experimental and analytical investigation


on the postbuckling behavior in axial compression of blade-stiffened carbon-epoxy (T300
carbon fiber tape preimpregnated with 120oc cure Hexcel Fl55 thermosetting epoxy resin)
that was carried out in the mid-eighties at the Fokker Aircraft Company in cooperation with
the National Aerospace Laboratory NRL, The Netherlands. Tests with full-scale panels de-
signed to buckle in a global mode and with short columns designed to buckle in a local
mode were performed. The design approach of the specimens was identical to that adopted
by the Australian CRC-AS tests [14.121] and [14.122], i.e. they had to be integrally stiffened
and cocured and the lay-up of their stiffeners (blades) had to be symmetric. Again, to comply
with this latter requirement, every other skin lay-up between stiffeners inevitably became
asymmetric. The panel, stiffener and skin geometry and the skin and stiffener details are
shown in Figure. 14.86. Note that Figure 14.86b represents two designs, one for a root section
and one for a tip section of a stabilizer. To perform the local buckling mode tests, full-scale
panels were cut into short columns containing either three or six stiffeners. The loaded ends
of the panels were potted in an epoxy-resin material and machined parallel to ensure uniform
loading in testing and avoid brooming of the carbon fibers. The unloaded side edges of the
specimens were free.
The panels were loaded in a 900 kN capacity servohydraulic Wolpert-Amsler testing ma-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

chine under displacement control, in steps of 0.02 or 0.04 mm. Back-to-back bonded strain
gages and linear variable differential transformers applied to selected points were used to
record strains and out-of-plane and longitudinal in-plane displacements. Out-of-plane dis-
placements were also visualized with the moire shadow technique and with sliding trans-
ducers passing over the panel surface. The short columns were all tested to failure, whereas
the tests with the full-scale panels were terminated, to save the specimens for further testing,
once the end shortening started increasing rapidly without adding any load to the panel.
Analytical predictions of the buckling behavior of the panels were obtained with BUS-
CLASP [14.112], while their nonlinear postbuckling response was analyzed with STAGS C-
1 [ 14.88].

..........
, ..,. ..
S~J.n and stiffener laalnar ..

. b

fl.\ •• •• .. I.
•••• 10
" "
II
" 12
"
D••1&n
sy--tric•l Non-sy. . . trl~;ll St1Henu I Stifhn.r
tyJ~e
Sltin (:!:) Skin (NS) tlP• • , 111

root (i·45/0 /H~)


3
(t45/0l/!4'5) <••Ho,,:;:..,l I,..,,,o,i•Ul
tip ,._,,,/!'I) (t45/0 /.t4S) (t4S/0
6
t+45) (i-4SI0 1/t4S)

IV ... ETRICAL IICIN NOH-IY...ITfUCAL SIC IN

(a) (b)

Figure 14.86 Wiggenraad's blade-stiffened test panels (from [14.124]): (a) panel geometry and stiffener
Copyright Wiley
Provided by IHS Markit under license with WILEY
geometry, (b) panel lay-up procedure Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1130 Composite Structures

On concluding the test results and comparing the predictions with the experimental results,
it was pointed out in [14.124] that the full-scale panels buckled in a global mode without
showing any further postbuckling load-carrying capacity. On the other hand, the short col-
umns, which buckled in a local mode, sustained 1.8 times their buckling capacity before
failing. Their failure probably initiated at the free stiffener ends and resulted in a complete
fracture of the cross-section along a nodal line. In the case of global buckling, BUCLASP
overestimated the experimentally observed buckling loads by about 15 percent. It was sug-
gested that this was due to the inability of the code to account for out-of-plane buckling
deformations. Good agreement between BUCLASP predictions and the experimental results
was obtained in the case of local buckling. The STAGS predictions also showed good agree-
ment with test results. It is worth noting that since the skin of the panels was relatively thin,
bending-torsional coupling existed. Consequently, the local buckling mode showed skew
buckles patterns.
In another joint study between the NLR and Fokker Space Division [14.125], fiat stiffened
panels were tested in compression to verify selected design details of the Ariane 4 interstage
2/3 CFRP blade-stiffened cylinder. The panel design was constrained by the requirement
that interaction between the general and local buckling modes should be avoided, and there-
fore the panel was designed to buckle in a general instability mode lower than the local
buckling loads. This resulted in a buckling-resistant design (Figure 14.87a) of the following
characteristics: skin laminate ( ± 45/04 / ± 45/0 4 / ± 45); stiffener laminate ( ± 4510) ±451
0 5 / ±45); flat panel local buckling load = 742 N/mm and flat panel Euler bucking load =
532 N/mm. The test program included three different types of subcomponent panels to
(1) determine the local buckling behavior, (2) determine the strength of the panel-to-end-
ring connections and (3) establish the global panel buckling behavior between ring frames.
Tasks (2) and (3) of the test program appear to be of particular interest and warrant detailed
discussion.
The strength of the panel to end-ring connections was evaluated by testing short panels,
featuring the proper connection to the ring at one loaded end, potted epoxy blocks at the
other loaded end and free unloaded edges (see Figure 14.87). It is apparent from Figure
14.87b that since the load is introduced from the ring into the skin through the bolts, the
skin laminate must have sufficient bearing strength to carry the bolt loads. The load con-
centrations caused by the adjacent structure require additional strength. Consequently, skin
reinforcements depicted in Figure 14.87c were introduced. The 0 plies in these reinforce-
ments enhanced the axial strength, and the ± 45 plies provided the bearing strength and
augmented the distribution of the load concentration. In the tests with these types of panels,
failure was observed at loads 5-8 percent smaller than the expected ones. Failure was as-
sociated with consecutive separation of two stiffener ends from the skin over approximately
35 mm. This was due to severe bending caused by load eccentricity induced by the stiffener
run-out, the edge laminate reinforcement and the one-sided connection to the end ring. Long
panels (see Figure 14.88) were tested to establish the global buckling behavior. Like the
short panels, these panels were connected by bolts to an end ring at one loaded end, the
other loaded end was potted in epoxy blocks and the unloaded edges were unsupported. The

, ....,
L~HINO
MAli"IAl (b)

==I~=======
I '\.__ :t: -PlYotfron

"'
.--.r-.,
~7
,;-.:;-+;- ..
• • + • • • ~ • • • • +
(c) _ _ _ _ _ _
(a)

Figure 14.87 Blass and Wiggenraad's short blade-stiffened test panel (from [14.125]): (a) panel ge-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
ometry, (b) end ring configuration, (c)
Provided by IHS Markit under license with WILEY
skin reinforcement at end ring connection
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 113 1

....
...-:.t l."J lr.-. r:J r.-J

---- ---- --- ---- ---- ...


!
1-.-.· ,,
.. .. .. .. ..: T1 n T1

Figure 14.88 Bins> und Wigg~nrnad's long blade-stiffened test panel (from [ 14. 125])

panel was also provided with an alum inum "angled" plate to simulate an intermediate ring
frame and obtain a buckl ing mode sim ilar to the critical one. corresponding to the complete
cyli nder. In the tests the panels failed in the expected global buckling mode. at loads well
above the calculated ultimate load. Again. the fai lure mechanism consisted of skin stiffener
separation; however, as was not the case in the short panels, failure of the long panels was
observed at the location of maximum panel dcncction. This failure was followed by sideways
buckling of the loose stiffeners. leading to more extensive damage (sec Figure 14.89).
It was pointed out in [ 14.125] that due to the nonlinear behavior stemming from the
stiffener eccentricity. which was rcnectcd by the displacement patterns. obtained with a
sliding traveling transducer during the loading of the panels. it wa~ difficult to detenninc the

f igure 14.89 Blass and Wiggcnraud 's long l>tlncl failure mode associated with sideways buckling of
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley loose s1iffcncr' (from ( 14. 125))


Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1132 Composite Structures

actual buckling load of the panels. These buckling loads were determined from the records
obtained by the strain gages located at the buckling crest of the panel.
1 as well as I stiffened panels were included in an extensive test program on the post-
buckling and durability of graphite-epoxy panels under cyclic compression loading that was
conducted in the late eighties at the Aerospace Structures Laboratory (ASL), Faculty of
Aerospace Engineering, Technion-Israel Institute of Technology, Haifa, Israel. This inves-
tigation was reported in [8.125], [14.91] and [14.126]. The program aimed at assessing the
capability of the flat stiffened panels to sustain repeated axial compression loads far in excess
of their initial buckling loads (4.4-5.5 times).
The panels were fabricated from AS4/3502 unidirectional graphite-epoxy tape material in
a cocure process. Their geometry and dimensions are shown in Figure 14.90 and the lay-
ups of their skin and stiffeners are presented in Table 14.1. Three panels of each stiffener
type were manufactured. One specimen of each type was statically tested to failure. The two
remaining panels were loaded in cyclic compression to a life goal of 250 k cycles. Cyclic
loading was applied in blocks of constant amplitude compression load at load levels ranging
from 150 percent of the panel initial buckling to about two-thirds of its ultimate strength
(limit load). Preceding each block of cyclic loading, a static loading test one ton above the
last block level was performed to assess the predicted effects of repeated loading (see Figs.
14.91, 14.92 and 14.93) as well as to verify that the panel was capable of carrying the load
amplitude of the last prescribed block. It is apparent from these figures that the repeated
buckling barely affected the static behavior of the panel. After completion of the cyclic
loading, the residual strength of the panel was determined. It was observed that failure of
all of the panels initiated at the caps of the two central stiffeners, almost midway between
the panel loaded ends, and of the outer stiffeners at approximately one-third and two-thirds

DETAIL B

4
I I (p I
D.O-I 1-160.0-t-160.0~ 160.0-ol ~o-40 ' 0
SECTIOII A-A

DETAIL A

..[ _[ (f) ..[


1-160.0-160.0 ...... 160.0·~ .....
40 40 0
:!'.t •
SECTIOII A-A

ALL DIMENSIONS IN MILLIMETERS


1 - - - - 560.0 - - - 1

Cl,:(t4S;D4 ;90;0z0*45;0)
1

1--- ll.' ---t

o) I - SUtteuer b) J - SL lr I oDor
OETAI L 8 DETAIL A
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Figure 14.90 Technion ASL axial compression stiffened panels-geometry, dimensions and lay-ups
(from [8.125])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1133

Table 14.1 Panel lay-up orientation and sequence (from [8.125])


Configuration Skin Stiffener web Stiffener cap
!-stiffener (±45/0/ ±45/90), (±45/90/0/ ±45), (±45104/9010,1 ±45/0),
]-stiffener (±45/0/ ±45/90), ( ±45/90/0/ ±45), (±45/04/90/0,/ ±45/0),

of the length of the panel, where the broken 4SO fibers coming off the inner stiffeners met
the external stiffeners (see Figure 14.94). No separation failure mode at the skin-stiffener
interface occurred.
In another NASA Langley study, by Stein and Williams [14.99], a minimum-mass structural
efficiency curve was determined for sandwich blade-stiffened composite compression panels
subjected to buckling and strength constraints. The study included buckling experiments on
seven moderately and lightly loaded designs of clamped one-bay layered graphite-epoxy
panels having two sandwich blade stiffeners with aluminum honeycomb core. Good agree-
ment among the theory of [14.99], the experiments and finite-element (NASTRAN) calcu-
lations was reported. It should be noted that the analysis developed in [14.99] assumed that
the ends of the panels were simply supported but accounted for transverse shear deforma-
tions. As a consequence of the good correlation, it was suggested that detailed boundary
conditions are not important when predicting the buckling of sandwich blade-stiffened panels
where twisting occurs and through-the-thickness transverse shearing is important.
A substantial experimental study on the postbuckling behavior and failure characteristics

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
of selected fiat stiffened graphite-epoxy panels in compression loading was conducted at
NASA Langley Research Center by Starnes et al. in the eighties [14.42] and [14.102]. The
specimens tested in this investigation were fabricated from commercially available unidirec-
tional Thorne! 300 graphite fiber tapes preimpregnated with 450 K cure Narmco 5208 ther-
mosetting epoxy resin. Each panel had four equally spaced identical T-shaped stiffeners. A
relatively heavy stiffener design was selected for the panels so that the nonlinear response
of the panels would be primarily due to the buckling of the skins. In order to evaluate the
effect of skin postbuckling behavior on panel performance, both the skin thickness and
stiffener spacing were varied, 16- or 24-ply quasi-isotropic skins with either 102-, 140- or
178-mm stiffener spacing. The panel lengths were selected so that in compliance with the
stiffener spacing at least four longitudinal half waves would be contained in the initial buck-
ling mode (500, 660 and 813 mm). The panel and stiffener geometry and details are presented
in Figure 14.95. A typical test specimen is depicted in Figure 14.96. To determine the effects
of the stiffeners on panel behavior, the lay-up of the stiffener cap (Figure 14.95b) of the 16-
ply skin panel with the largest stiffener spacing ( 178 mm) was varied. The combinations of
skin and stiffener laminates employed in the tests are listed in Table 14.2. In the program

24 - - BEFORE CYCLI HG 16 --BEFORE CYCLING \


-·-AFTER lOOK CYCLES -·-·AFTER 100P; CYCLES 1/
---··AFTER lOOK CYCLES '% 14 -----AFTER 150K CYCLES :
e :'
'0:'12 ~s.a. II
g. 10 tml1~ I·
aa
~
UlllJ ,\\\
g; 6
5u •
BUCKLING
...
~

x 1

OL-~~~~~~~~-L~
"'
o 200 •oo 60n eoo 1000
(a) COMPRESSION STRAIN ( 10'6 HH/HM} (b) COMPRESSION STPAIN (10. 6 HH/HM]

Figure 14.91 Technion ASL typical test results of axial compression strains in center of the skin of a
stiffened panel under repeated axial compression buckling (from [8.125]): (a) !-stiffened
panel, (b) ]-stiffened panel
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1134 Composite Structures

z::!
24
22
--BEFORE CYCLING
-·-AFTER lOOK CYClES /
z
2
16

14
- - BEFORE CYCLING
-·-·AFTER lOOK CYCLES
----AFTER 150K CYCLES
I /

~1·~
20

~ :: t001 &..1
----·AfTER lOOK CYCLES

..f''
.•/
//
1 ·' 0..,12
0

~10 ~m~~
~-
~SG·

~ 12 l1l:UJ ,/ ;:;; 8
~
~ 10 /'/' "'~ 6
g: ..v
i!j I ~· '-'
~ 4
..., ' -;:::~?P·" <
;( 4 .'/f X
x-< 2
< 2

100 1600 2400 llOO 4000 4600 4 00 1200 1000 2800 1600
CURVATlffiE,h·t ( 10' 6 HHIHH) (b) CURVATURE,~· t ( 10· 6 HHIHH]

Figure 14.92 Technion ASL typical test results of longitudinal curvature of the skin of a stiffened
panel under repeated axial compression buckling (from [14.91]): (a) !-stiffened panel,
(b) ]-stiffened panel

three panels of design "l ", one for each stitlener spacing, three panels of design "2" (50-
ply cap laminate), one for each stiffener spacing, and two panels with a stiffener spacing of
178 mm, one of design "3" (38-ply cap laminate) and one of design "4" (30-ply cap
laminate) were tested.
All specimens were cured in an autoclave. The stiffeners of all of the panels with 24-ply
skins were cocured with the skins. The stiffeners of all of the 16-ply skin panels were
secondarily bonded to the skins. Following cure, the specimens were ultrasonically inspected
for defects. Some small delamination defects, up to 4 mm in diameter, were detected in the
skins directly under the stiffeners in the 24-ply skin specimens. No defects were found in
the specimens with 16-ply skins. The loaded edges of each specimen were potted in an
epoxy resin material to prevent "brooming" of the graphite fibers during handling. The
potting material was encased in a steel frame for protection (Figure 14.96). The loaded edges
were machined fiat and parallel to permit uniform compressive loading. The unstiffened side
of each panel was painted white for application of the moire technique to monitor out-of-
plane deflections during testing.
A 4.45 MN capacity hydraulic testing machine was employed to load the panels in com-
pression. The panels were fiat -end tested without any lateral supporting of the unloaded
edges. Strain gages were used to monitor the strains (see Figure 14.96), and direct-current

2200r - - BEFORE CYCLING


~ I ~ -----AFTER 2 OOK CYCLES . .
~ 2000 ·;-:~~ .. --·AFTER lOOK CYCLES ;,;--,~
l: \\ •I •
X: • f,
>;16oo I ~· 1

g \ . -- '

~
~1100
1600

1400
~

\ g 1.~1
b

"'
~ 1000
\
~
.,.~.
. _)

~ 800 \..._,__../
<
6QOL-~~.~~-L-L~-J~L-~
0 0·2 Q.l, o.6 0·6 1-0
yl ii
Figure 14.93 Typical influence of repeated axial compression buckling on the axial membrane strain
distribution in the skin of an !-stiffened panel-Technion ASL tests (from [8.125])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1135

Figu re 14.94 Typical failure modes of axial compression panels obsen•ed in the Technion ASL tests
(from [8.125]): (a) 1-stiflcncd panel, (b) J-stiffened panel

differential transformers were used to monitor longitudinal in-plane and out-of-plane dis·
placements at selected locations. All strains, displacements and corresponding applied loads
were recorded on magnet ic tape. The out-of-plane deflections obtained by the moire fringe
method were recorded photographically.
All the panels tested in the program buckled into a stiffened-panel buckling mode, in
which both the stiffeners and the skin deformed (see [ 14 . 102)). The buckling loads measured
in the tests were lower than the predicted ones. S ince a different val ue of PIP, (where P
and P. , are the applied and predicted buckling load, respectively) was observed for each
panel, it was suspected that the panels had varying amounts of initial geometric imperfec·
tions, which was responsible for the degradation in buckl ing load capacity. It was also
observed that failure of all spec imens initiated at a skin-stiffener interface and that the failure

).•., r ., ' 1- , t-u


AMAlYTICAl 1 1 I 1
tcuiOARY

~ t--w ' - - ! '·'


n)
u•(()I'SlA!fl
, . .,... o ~ T

..
"·• · "·r. ·f 0
l
. IS R
,.,

(a)
'
0


00

POlliNG MAftltl.-( .1
AI() S:UU rAM\(
J

(C)
I
LUNIDIRECTIONAL
fAP£ fillER
LsKtN
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

All OIMENSI ONS IN CM.

(b)

Figure 14.95 Starnes eta i.'s flat 1-stiffe ncd test panels (from [ 14.102)): (a) panel geometry, (b) stiffener
geometry, (c) stiffener auachrnent flange details
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1136 Composite Structures

Figure 14.96 Starnes et al.'s typical test specimen with poned


ends encased in a steel frame (from [1 4.102))

mode of all panels was the same, skin and stitTener separation from one another in the
interior of the specimen. ll wa$ therefore concluded that the high membrane strains that
existed at the stiffeners after buckling were sufficient to induce local skin-stiffener separation,
leading to comprehensive panel failure. A typical failed panel is depicted in Figure 14.97.
As pointed out in [14.102). the postbuckling test results correlated well with results from
STAGS C-1 nonlinear analysis 114.88]. Typical comparisons of test results with STAGS
C- 1 predictions; end shortening versus applied load. out-of-plane deflecti on near a point of
maximum buckl ing mode amplitude versus applied load and membrane strain distributions
across the center bay of a panel for various applied load levels are shown in Figure 14.98a-
c. respectively. Very good agreement between test resulL~ and analytical predictions is ap-
parent from these figures. One concluding remark from 114.102) is worth noting: " Since the
analytical buckling mode had a considerable amount of deformation in the stiffener webs
and attachment flanges. it was necessary to model these stiffener components with plate
elements having the appropriate stiffness to obtain satisfactory correlation with the postbuck-
ling test results." --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Additional tests on graphite-epoxy hat-, blade- and Y-stiffened panels in axial compression
were carried out at NASA Langley at the end of the eighties [14.103]- (14.105).
Posrbuckling failure rests of composite /-stiffened compression panels, performed by Ste-
vens. R. Ricci and G.A.O. Davi s at Imperial College of Science, Technology and Medicine,
were recently reponed in [8. 1651 and [8.166]. These tests represent an important attempt at
gaining better insight into failure mechanisms of stiffened composite panels and at provid-
ing a tool to overcome the present inadequacy of global finite-clement analysis models. no
matter how complex, for prediction of posthuckling failure of stiffened panels. To achieve
this goal, tests on stiffened panels were conducted in conjunction with an effon of testing
two-dimensional component strips cut from similar panels so that the strips wou ld simulate
the failure mode of the panel correctly. without the panel compressive load bei ng applied.
The results of these two-dimensional component tests would be used as a calibrator for the
global finite-element prediction.
The !-stiffened panels used in the program (sec Figure 14.99) were fabricated from T300
914C unidirectional prepreg. The stiffeners were either cocured or bonded to the skin with
a layer of B.S.L. 322 film adhesive in each ca.5e. It is apparent from Figure 14.99 that the
proponion of the panels, skins and robust stiffeners was again chosen so that considerable
postbuckling would be experienced (about four times the buckling load).
The panels were axially loaded under displacement control in the special purpose 250 ton
panel test machine depicted in Figure 8.113 of Chapter 8 (described in Subsection 8.7.3).
The unloaded edges of the panel were unsupported.
In tha tests it was observed that all panels buckled initially into six axial half-waves with
the stiffener caps remaining essentially straight. whereas the stiffener web and flange de-
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Table 14.2 Skin and stiffener laminates (from [14.102])
Design Skin Stiffener web Stiffener cap
(±4510,1+45190,1 ±45/0/90), (90/0/ ±451+4510,1 ±45/0/90), (90/0/ ±451+4510,1 ± 1019010,1 ±45/06/+45/0,),
2 ( ±45/0,/+45/90,), 0/ ±451+4510,1 ±45/0,)1 (0/ ±45/+45/0,/ ±4510,19010,190,104190,10,),
3 ( ± 4510,1 +45/90,), (0/ ±451+4510,1 ±4510,), (0/ ± 45/ +4510,1 ± 4510,19010,19010,),
4 (±4510,1 ±45/90,).1 (0/ ±45/+45/0,/ ±45/0,), (0/ ±451+4510,1 ±4510,19010,)1
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

.....
.....
...,.
(,)

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1138 Composite Structures

Figure 14.97 'J)Ipical skio·stiffener separation failure mode of 1-stiffened panels observed in the NASA
Langley tests (from (J4.102))

f. t,

'
f;,
l

w
(a) i

mru·1-
...
'u
u

..
.l . ...
(b)

Figure 14.98 Comparisons of Starnes et al.'s test results with STAGS C· l predictions of postbuckling
response (from 114.102]): (a) end sho11Cning versus applied load. (b) membrane strain
distributions across the center bay of the panel (c) out-of-plane dellection near a point
of maximum buckling-mode amplitude versus applied load
--`,`,`````,`,````,,``,``,,`-`

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1139

Stiffener cap: b 6.25 mm


c 4S/-45/0I451-4510 21451-451o2 t90I0 3 1902 Jo 4 tso 2 10zl 5

Tapereod stiffener flange:


[45/-4 5/0/451-45/02145/-45/02 Js

-1
I
I 2.75
-1--

1~43.2---1
Figure 14.99 Stevens et al.'s I-stiffened test panels-dimensions, stiffener details and lay-ups (from
[8.165])

fleeted as shown in Figure 14.100. With further increase in load, the longitudinal mode
suddenly changed from six to seven half-waves without change of this deflected pattern of
the stiffener. This phenomenon of mode switching and its implications for nonlinear com-
putations was discussed in detail by Bushnell (see [13.135], [13.136] and [14.127]), as well
for metal plates in Subsection 8.1.3 of Chapter 8.
It was further reported in [8.165] and [8.166] that initial failure of the panel was experi-
enced at three locations indicated on the buckled mode of the panel in Figure 14.101 a. These
locations, which were detected using ultrasonics, were anticipated since the unsupported
edges of the panel consisted of a narrow width of skin together with the stiffener flange,
which stabilized the narrow edge region. Also, due to some overall bowing, the skin away
from the panel ends was more heavily buckled. Each failure was accompanied by an audible
crack, running in the first plies of the skin from the base of the stiffener web towards a
location under the flange (see typical failure in Figure 14.10lb), on the tension side in
relation to the web bending moment.
It was further observed that the cracks extended along the panel for a distance equal to
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

about one half of the buckling half wavelength. With further loading no propagation length-
wise of the panel was experienced, but the cracks could travel to the free edge of the flange.
Bending moments in the web, measured during the test by strain gages bonded to both web

Figure 14.100 Deformation at an antimode of !-stiffener web and


flange (from [8.165])

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1140 Composite Structures

FSl

Buc:kl rg mod* cr~d foilllft ~ot.K


( o) (Yt~d lnrougl\ sk,,.l

Figure 14.101 lniual faolurc mode obsened on Stevens et at:~ buckled 1-stiJJener pane), (from [8.165)):
(a) failure sites viewed through skin. (b) faolurc in 3 cocured panel

surfaces, suddenly dropped when failure ini tiated. The moments were further reduced as the
crack propagated out under the skin.
Buckling of the panels tested occurred at a load of :1bout I I tons. and they collapsed
under loads greater than four times their buckling load . Cracks initiated at a load that was
close to the final failure load. In panels experiencing initial failure of the type shown in
Figure 14.10Ia and slowly taken up to the collapse load. failure was ~o sudden that the
change from local failure to total collapse was not discernible. An observed pattem of a
collapsed panel showi ng the separation of sti ffeners is depicted in Figure 14. 102.
Based on the above test observations of panel failure. potential failu re mechanisms were
proposed in 18. 135]. For the failure ~hown in Figure 14.101b. bending moment in the web
was suggested as the cause for failure. In this case the thin skin. which buckled first, inter-
acted with the stiffener tending to twist it in a clockwise direction (see Figure 14. 100). Since
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Fi~:ure 14.102 Collap..e panem showing >tiiTener-skin separation ob,crved in Steven> ct nl."; tests (from
[8. 165))
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1141

the torsional rigidity of the stiffener was almost entirely provided by the torsion-bending
stiffness of the stiffener cap, the web experienced the bending moment distribution shown
in Figure 14.100, reaching its maximum at its base. This moment can be shown to have a
tendency to peel away the right-hand flange of the stiffener. The threshold at which this
peeling starts depends very much upon the highly complex stress field inside the triangle
contained between the stiffener flanges and the plate skin to which they are bonded (PANDA
2, [13.135], [13.136] and [14.127], calculates this effect from T-peel test data).
To estimate the web bending moment associated with the above-described failure mech-
anisms, related to disbonding of the stiffener flange, a transverse strip from the panel, 40
mm wide, was loaded in the special rig shown in Figure 14.103. This rig deformed the strip
in a similar manner to that in the postbuckled panel and introduced in the web a bending
moment distribution representative of the panel test (see Figure 14.1 00). In the rig, the test
strip was bolted through the stiffener cap to a fixed base E and the skin was clamped at
points C and D to a rigid beam, which pivoted about a fixed axis at B, by a displacement
applied at C. The bending moment variation in the web obtained in the rig in terms of the
force applied at Cis presented in Figure 14.103.
A comparison of micrographs of failure obtained from tests of cocured and bonded panels
with those observed in the component tests (reproduced from [8.165]) is depicted in Figure
14.104. The similarity is indeed striking, but it should be noted that the component tests
overestimated the web moment at which cracking occurred in the panel. Nevertheless, it was
pointed out in [8.165] that, in spite of this disagreement and the limitations of the component
test as a true representation of the panel, due to its two-dimensional nature and the omission
of axial loading, this test was much more likely than the full-scale test to provide information
about the nature of the damage mechanism before the unstable crack appeared.
Based on the above observations it was concluded in [8.165] that the proposed component
test method, in conjunction with a nonlinear finite-element calculation, suggested a method
of determining the postbuckling failure load of a stiffened panel without performing a full-
scale test.
In a recent Cooperative Technology Development Program on Advanced Composites of
the Douglas Aircraft Company in California and the ALENIA Company in Italy [14.128]
for development of new manufacturing technologies and for studying structural solutions
adequate for an unpressurized aft fuselage of a medium-sized transport aircraft, bead-
stiffened compression panels emerged as the preferred compromise (Figure 14.1 05). This
configuration was chosen after investigation of a typical panel stiffened by several types of
stiffeners and comparison of manufacturing costs, weight savings and structural risks char-
acteristics.
The panels tested in the program (see Figure 14.106) were designed to fail in buckling.
The main goal of the test program was to investigate their buckling behavior when subjected

Componem skm
clamped 10 ~am.

WcbB.M.

Figure 14.103 Stevens et al.'s loading rig for studying !-stiffener disbanding failure due to stiffener
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

web bending (from [8.165])


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1142 Composite Structures

~~ ·
~- ·--
(_·ool"Uf\- d (_'O!IIIpo!ll~·l 'l

Figure 14.104 Comparison of failure modes observed in the lmpeti al College tests on bonded and
cocurcd panels with fail ure modes obtained in tests on bonded and cocured strips loaded
in the test rig of Figure 14.103 (from [8.165 ])

to compression and to veri fy the theoretical predictions o btained by the BEADBUCK code
developed by Douglas (Boeing) Aircraft Company.
The panels were manufactured in a cocuring process and fabricated from stiff fibers and
thickened thermosetting resin system CFRP tape and fabric (see Table 14.3). Two Z-shaped
frames were connected to the panel skin by shear ties. The junction frame/shear tic and
shear tie/skin was obtai ned by locks. The two frames were introduced to provide a stabi lizing
effect o n the stiffeners, as much as realistically possible. To allow a realistic comparison
between theory and experiments. the height of the panel was defined in such a manner as
to have its instability occur under the same load as that calculated in the analysis of the
fuselage compression panels (using end fix ity coefficients equal to 2).
Buck ling loads of the panels were determined from the responses of the panels (load
versus strain) obtained from strain gages bo nded to the skin of the panel at various locations
(see Fig ure 14.1 06). The experimental resu lts correlated well with the analytical predictions
yielded by the specially dedicated BEADBUCK code.
Buckli ng of open-section bead-stiffened composite panels was studied analytically and
experimentally in [ 14.129]. In contrast to the bead-stiffened panels of [ 14.1 28). the open-
section beads have relatively low torsional rigidity. However. because of their manufacnuing
methods, such as thermoformi ng. which make integral formatio n of the stiffeners in the panel
as one piece feasible, they provide an economically allractive alternative to conventional-
type stiffened structures.
The investigation focused on a parametric study to determi ne the effects of bead spacing
and bead cross-section (see Figure 14. I07) on the initial buckl ing capacity of a bead-stiffened
thermoplastic composite panel. Part of this effort was devoted to evaluating the effective

Fi~u rc 14.105 Alenia bead-stiffened compression panel concept (from [ 14.128])


--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1143

f TYPII
r+c

r:r
1(11 319

1"-

~ Vr11a
722.4 508
B 0/ 12-----
L 2
.
5 Sf 488.
I j
B

,~~1
I...J
/
• 234.
25.4 107.2 -<::::::" I '7
cc I /4110
f 0 LAT &. CURVED PANELJ
L..
c

A-A
[ FLAT PANEL l
BEAD (<0190)1(!45),(0/90))

L '*' '*'
I~
. 761.0
B-B
[SKIN (901451-45/-4514510)

'*'1
[FLAT PANEL]

ALUMINUM FRAME POTTING


I COMPOUND
1

jl cp cp~
901.7
D-D
I FLAT PANEL]

Figure 14.106 Alenia bead-stiffened compression panel-dimensions and details (from [14.128])

Table 14.3 Unidirectional tape and fabric materials used to fabricate


Alenia bead-stiffened compression panels (from [14.128])
Supplier Fiber Matrix Nominal cured
ply thickness
(mm)
Tape: Ciba T800H Fibredux 0.147
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

6376C
Hercules IM7 8551-7
Fabric: Ciba T800H Fibredux 0.315
6K-5H 6376C
E2743
Hercules IM7G Mag amite 0.356
12K-5H XSW370-SH/X855l-7
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1144 Composite Structures

I
I

- - r-
j./
c~'\
t l I

L PW CW FW RW
J
-
I

t'\.
J tL+ L - -
t

View/\

Figure 14.107 Laananen's open-section bead-stiffened compression panel-geometry and dimensions


(from [14.129])

boundary conditions provided by the bead stiffener to the flat panel enclosed within the bead.
The parametric study was performed using the NASTRAN code. In order to verify the critical
buckling loads predicted by this finite-element analysis, two bead-stiffened panels of AS4/
APC2 thermoplastic composite were manufactured and tested. The panels had (0/90/45/
-45) 5 skins with 25-mm high bead. The panels were produced by double diaphragm forming
at 380°C and 690-760 k Pa.
Each test specimen was instrumented with strain gages bonded face-to-face at the center
of the panel to monitor the state of stress·-strain in the panel and enable determination of the
onset of buckling by the Southwell technique. The panel was clamped around its perimeter
by the loading fixture. This resulted in a 480 X 480 mm2 test area inside the fixture that
was modeled as a panel with clamped boundaries on all edges in the finite-element analysis.
[14.129] indicated that the test results were 4-12 percent lower than the finite-element pre-
dictions.

Buckling tests of transversely stiffened FRP panels- Fiber-reinforced plastics are very ex-
tensively used in the boat-building industry and in larger ships, including military patrol craft
and minesweepers. The need for general robustness and high strength under transverse loads
often leads to the adoption of transversely framed hull structures in these boats (Figure
14.108), containing few, if any, longitudinal stiffeners. Clearly, panels designed by this ap-
proach (see Figure 5.28a) are "weak" in resisting longitudinal compression associated with
hull bending. Unlike longitudinally stiffened panels, whose inherent elastic stability is con-
siderably higher and which can sustain substantial postbuckling strength, these transversely
stiffened structures are likely to exhibit catastrophic failure at loads approximately equal to
the theoretical critical values and hence must be designed with extreme care against elastic
buckling. This is particularly true for glass-reinforced plastic (GRP) panel designs, which
are very susceptible to instability because of their low material stiffness. Furthermore, the
anisotropic nature of the material and the common practice of using hat stiffeners in GRP
hull construction complicate the buckling behavior. Consequently, buckling failure of lon-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1145

Figure 14.108 Typical transversely framed hull 'tructurc (from C.S. Smith, "Compressive Strength of
Transversely StifTcned FRP Panels,'' on: Axpects '!{the Anlllysis of Plate Struclllres, D.J.
Dawc, R.J. Horsington, A.G. Kamtckar and G. H. Liule, eds., copyright 1985, by per·
mission of Oxford University Pre>,.)

gitudinally compressed. tran~vcrsely stiffened panels (Figure 5.28a) is of significant impor·


tance in the design of FRP \hips.
This buckling problem. leading to failure under axial compression of transversely framed
bottom and deck panels of a mi nesweeper hu ll. and the co rresponding test program were
di scussed in Subsection 5.9.3 of Chapter 5. Smith [14.1301 made an cll'ort to assemble
available experimental data on 'mall model testing. as well as o n full-scale model,, together
with new data. He also extended the theoretical analysis by considering nonlinear buckling
behavior and imperfection •en,iti vity and correlated the test data with alternative theoretical
estimates of buckling strength, and consequently drew up recommendations for improved
design of transversely stiffened GRP panels subjected to axia l compression .
It wa.' pointed out in [ 14. 130[ that if simple support condition' along the unloaded edges
of the panel can be assumed. the initial buci.Jing loads for longitudinally compressed, tran\·
versely stiffened GRP panel~ could be evaluated by employing (I) folded-plate analysis.
which establishes the lowest init ial buckli ng stress together with the correspo nding instability
mode. one of which is presented in Figure 5.27. (2) data curves for approxi mate evaluation
of interframe buckling and (3) a finite-elemcm analysis. the real merit of which lies in it>
ability to deal with irregu lar ~tnoctural geometry and boundary condition~. Furthennore.
finite-element analysis may be applied incrementally to 'tudy the nonlinear buckling and
postbuckling behavior. as well a' the imperfection sen,itivi ty of stiffened panels. It was
indicated in [14. 130) that the finite-clement rrcclictions were found to correspo nd quite
closely with the folded-plate solutions ( I) .
The assembly of tests su111mariLed by Smith in [ 14.130] includes ( I) tests on small Perspex
models that were reported in )5.60] and di<;CU\\ed in Subsection 5.9.3 of Chapter 5. as well
a~ (2) tests on large-scale panel~ reported in [5.61). together with unreported tests on two
additional identical large-scale panels, de,ignated panels "4" and "5" in [ 14. 135]. whi ch
correspo nd to the bottom shell structure in a typical shi p design. The Iauer panels had an
overall lengt h of 6096 111111 and e ffective trathverse span of 3048 mm . They were each
~tiffened by eight equally !.paced transver.,e hat-section fr:unes (see Figure 14. 109). The
panels were fabricated from E-glass rcinforccrncm embedded in a ship-type isophtalic pol·
yester resin. The panel skin wa' laminated from 37 plies of 800 gm 2 woven rovings (WR).
wh ile the frames incorporated 14 plies of 800 gm 2 WR with 12 plies of 600 gm 2 unidi·
rect ional glass in the tables.
The panels were tested in the large test frame (LTF) at the Adm iralty Marine Technology
--`,`,`````,`,````,,``,``,,`-`-`

Establishment. Dumferline (AMTE), which wa., described in Subsection 12.3.5 (see Figure
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1146 Composite Structures

(a)

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
(b)

figure 14.109 Smith's large-scale GRP transverscly-Miffcncd tc>t panel (a) definition of panel dimen-
sions, (b) stiffened panel assembled with cnu bay> and side reinforccmcm< (from C.S.
Smith, "Comprcs>ivc Strength of TranWCI'l'cly Stiffened FRP Panels." in : Aspects of
the A11alysi< of Plate Strucrures, D.J . Dawc. R.J . Hor,ington. A.G. Kamtckar and G.H.
Liulc, cds., copyright 1985. by penni,.ion of Oxford University Press.)

12.1 4). Ln order to load the test panels uniformly a~ well a;, to avoid premature failure close
to their loaded ends, 1he end bays of the panels were reinforced by increa~ing the laminate
thicl..ness and fitting steel sandwich plates. boiled and bonded to the GRP (~ee Figure
l4. 109b). The unloaded side; of the panels were al•o reinforced by fiLLing short. discontin-
uous steel sandwich plales between transverse s1iffeners. The ends of the hat-section :.tiff-
encrs were streng1hened by sho rt built-in steel formers through-bolted to external steel pads.
P:mel "5" was tested to collapse under compressive load alo ne, whereas panel "4'' was
subjected to combined compressive and uniform hydro>Latic pressure. Latera l pressure was
applied by a water-filled rubber bag contained between the Jested panel and support platform
(sec Figure 12.14).
Prior to testing of each panel. its inilial deformation' were surveyed-the ma~imum in-
terframe displacement w 0 of the shell laminate. mea~ured in a gage length equal to the frame
spaci ng. as well as overall displacements relative 10 1hc ends of the panels mca~ured o n the
longitudinal centerplane. II was noted in [1 4.130] that the local distortions. significant in
relation to interframe buck ling. were generally very small and barely disti ng uishable from
surface irregularities and variat ion in laminate 1hickness. During tests. displacements were
recorded at closely spaced positions alo ng the ccntcrplane of the panels to check the uni-
fonnity of compressive load and appropriate vertical alignment of loading jacks. Strains were
measured on bolh sides of the panel skin laminme.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1147

Panel "5" was incrementally loaded in compression, with frequent removal of load to
allow inspection for damage and provide a check on permanent set and loss of stiffness. No
significant damage, permanent distortion, or loss of stiffness was observed up to an average
compressive stress of 60N · mm- 2 , when cracking noises commenced. With further increase
in applied load, "whitening" of the panel skin was observed at the base of the frame webs
and large buckling deformations developed. Collapse finally occurred at an average stress of
67.0N · mm - 2 , taking the form of a fracture over the full width of the panel in the central
bay. Failure of the skin laminate involved delamination with shear crimping and local buck-
ling of the delaminated plies. The deformation of the panel during the test, together with
the theoretical buckling modes of panel "5 ", are depicted in Figure 14.110. It is apparent
from this figure that the buckling stresses can be accurately estimated by folded plate anal-
ysis; however, the buckling modes are different.
Panel "4" was exposed first to a lateral pressure of 103 kNm- 2 • Then the lateral load was
held constant and compression load was applied incrementally. No damage was observed at
compressive stresses up to 60 Nmm- 2 • Collapse in a similar fashion to that experienced with
Panel "5" occurred at an average stress of 65.8 Nmm- 2 • It was shown in [14.130] that this
test result also correlated well with the folded-plate analysis prediction. From evaluation of
the comparison study of the test results of panels "4" and "5" of [14.130] and of those
experienced with the small test models and large-scale models of [5.60] and [5.61], which
are also presented in [14.130] with folded-plate analysis predictions, it was concluded in
[14.130] that initial buckling stresses of longitudinally compressed, transversely stiffened
panels can be accurately estimated by folded-plate analysis. It was cautioned, though, that
care should be taken to examine each of the buckling modes shown in Figure 5.27.

b. Shear Loading Tests


Buckling under shear loading of relatively thin composite webs surrounded by sparsely
spaced heavy stiffeners was discussed in Subsection 14.2.1c. That section focused on studies
of relatively complex deeply postbuckled or semi-tension-field shear panels as a means to
achieve minimum weight designs. As pointed out there, this design concept is most suitable,
as well as most structurally efficient, for comparatively lightly loaded areas of aircraft struc-

Figure 14.110 Comparison of observed large-scale panel experimental deformations and theoretically
predicted buckling modes (from C.S. Smith, "Compressive Strength of Transversely
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Stifl'ened FRP Panels," in: Aspects of the Analysis of Plate Structures, D.J. Dawe, R.J.
Copyright Wiley
Horsington, A.G. Kamtekar and G.H. Little, eds., copyright 1985, by permission of
Oxford
Provided by IHS Markit under license University Press.)
with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1148 Composite Structures

tures, e.g. aft and forward fuselage shell sections. It should, however, be noted that the use
of postbuckled tension-field designs for shear loaded panels requires that spacing of their
stiffeners be fairly close.
Experimental studies on buckling behavior and postbuckling of composite stiffened shear
panels are rather scarce. Some documented test programs can be found in [14.49], [14.42]
and [14.131]-[14.136]. The aim of the tests conducted at the Vought Corporation in the
United States [14.49], was the evaluation of a new design concept of tension-field stiffened
shear loaded panels that produce significant weight saving. Instead of the more common
design practice of tension-field shear panels depicted in Figure 14.111, a new shear panel
design with honeycomb stiffeners, shown in Figure 14.112, was proposed. This design con-
cept provided continuous load paths and stiffener-to-edge member joints. It also had excellent
potential for significantly reducing tooling costs and recurring fabrication labor costs. There
was some concern about panel warpage that could result from curing. However, sample single
web panels fabricated according to this new design concept turned out to be very fiat, dis-
tortion-free and structurally sound.
Following verification of the fabrication process, a tension-field cantilever test beam was
manufactured (Figure 14.113) and tested. The beam failed at 67 percent of design ultimate
load. It was suggested that this premature failure was due to curing and bonding deficiencies,
which were detected by C-scanning prior to the test, in the stiffener and edge member caps
and the stiffener-to-web bond. Buckling of the web of the panel, with three shear buckles
per bay, occurred at a load level equal to about 4 percent of the experimentally observed
ultimate strength.

Experimental investigations on stiffened graphite-epoxy shear panels supported by theo-


retical calculations were carried out at the NASA Langley Research Center in the eighties
(see [14.104], [14.131] and [14.132]). The picture frame concept (see Subsection 8.4.2 of
Chapter 8) was employed in these studies for introducing shear loads into the panels. In
[14.131] a graphite-epoxy shear panel with bonded-on }-stiffeners (Figure 14.114) was in-
vestigated. The design of this test specimen allowed local elastic skin buckling to take place,
while the stiffeners enforced a general instability failure mode at a load level much higher
than that of the skin buckling. In the test, a rectangular picture frame was employed to load
the specimen (see Figure 14.115). The details of load introduction from the frame into the
test panel are depicted in Figure 14.116. In this figure one can note that in order to ensure
that neither bending nor loads normal to the in-plane shear load would be applied to the
edges of the panel, these edges were attached to heavy steel frame bars by means of thin-
gage aluminum load introduction strips. Furthermore, to force the long edges of the panel
to remain parallel during the test, I 0 equally spaced aluminum cross-braces, 25.4 X 1.90
mm2 in cross-section, were placed between the long edges of the panel (see Figure 14.115).
To evaluate the anisotropic effects of the six-ply skin lay-up (Figure 14.114b) on the shear
buckling load, the panel was first pulled in the tensile loading machine along one of its

HAT SECTION ALL MEMBERS ARE COCURED


EDGE MEMBER""
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 14.111 Typical joint used in composite tension-field shear panels with conventional cocured J-
and hat-stiffeners (from [ 14.99])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1149

SICTION A • A
Xl\lt.A./11 'AIIt'IC
TtHSION rllLO
WEI · ~ 46•
{t a t.O#")

TV"CAL TI.HS~ N PIUD TY.-.CA_L INfliiSlCHON 0 ,


UilA. PANIR sn,llr'U SrtF'INtl AND MAJ0C \ONGUO N

Fi~:ure 14.1 12 Vought concept of tension field panel with integral stiffeners and longcrons ( from
[14.49])

diagonals and its behavior under this loading condition (designated as positive shear loading)
was studied. Then I he panel was un loaded and removed from the lesting machine, the loadi ng
yokes of the piclure frame were placed in lhe olher two corners and the panel was retested
under negative shear loading. The test results clearly demonstrated a marked anisotropic
effect, a ski n buckl ing load of 26.71 k.N under posilive loadi ng as compared 10 a buckl ing
load of 21.00 kJ\1 under negative loading, i.e. a ratio of 1.27. To assess the repeatabilily of
the observed lest resul ts. the panel was twice subjected to each loading direction.
To monitor the shear behavior of the panel. it was instrumented with IS back-to-back
strain gage rosettes, 5 rosettes along each bay of 1he panel. S ince at buckling. roseues on
one side of the buckled skin showed an increase in the calculated shear strain y, while the
opposi te rosette experienced a decrease in y with increasing load, the strai n reversal point
on the response plot, applied force versus shear strain (Figure 14. 117), was used to dclermine
the skin buckl ing load (the strain reversal method).
The tests performed on lhe stiffened panel, together with the stress analysis lhat accom -
panied it. revealed that use of the bonded doubler material under the sti ffeners and at the
panel edges (sec Figure l4.114b). in conjunction wi th the test fixture loading (see details in
Figure 14. 11 5), resulted in a highly non-uniform shear stress distribution in the panel. The
stress analysis was performed with the SPAR fin ite-element code of (14.133], where 1he
model used in the analysis included the complete details of the bonded doubler material
under the st iffeners and the panel edges. It shou ld, however. be noted thai to reduce com-
pulational costs. the stiffener web and its oulside !lange were not included in the model.
because they were not considered to significamly contribute 10 the shear stiffness of the
panel.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

T
16.2:5 IN.

l
1--- - - 32.0 IN. - - - . J
TEST SECTION

Figure 14.JJ3 Vought composite tension field test beam (from [14.49])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1150 Composite Structures

•--,

.
L
.
.....J Jn.tt

L~. _J
J ALL DIMENSIONS IN CM.
\.IJ+t-
J~It=::J ·f.L
~ I·"·" I .... I
Cjrectlor~of
l-1~t11191RC!II8\d

1+'·"
'"[1., 1
(a) (b)

Figure 14.114 NASA Langley J-stiffened shear panels (from [14.131]): (a) details of panel geometry,
(b) details of lay-ups of }-stiffeners

Comparisons of the finite-element results with test results are presented in Figures 14.118
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

and 14.119, which show excellent correlation between test results and the finite-element
predictions (which accounted for the influence of the test frame and the bonded doubler
material). It is apparent from these figures that predictions ignoring this influence will result
in significant errors.
In later NASA Langley shear panel studies [14.42] and [14.132] the picture frame of
Figure 14.39 [14.43] was used. Fifteen panels fabricated from unidirectional tapes of Her-
cules AS4/3502, with 19-, 22-, 25- or 33-ply skins, were tested in this study (see stacking
sequences in Table 2 of [14.132]). In the tests of [14.42] panels with 22-ply skin were not
included. The 19- and 22-ply skin specimens were stiffened with four equally spaced !-
shaped stiffeners, while the 25- and 33-ply skin specimens had two equally spaced !-shaped
stiffeners (see Figure 14.120). Both investigations studied the postbuckling response and
failure characteristics of the shear panels, as induced by the attachment concept of the stiff-
ener to the skin of the panel. To assess the attachment effect, the stiffeners were attached
with two different types of mechanical fasteners and/ or adhesive bonding. The investigations
of [14.132] were accompanied by finite-element analyses with STAGS C-1 [14.88].
Except for the 33-ply skin specimens, which failed near the buckling load, all other panels
failed under loads higher than the theoretical buckling load. The initial mode shape of the
buckled panels was one in which both the skin and stiffeners deformed out-of-plane and no
change in the buckling mode occurred as the panels were further loaded into the postbuckling
range. It was observed (see [14.132]) that failure of the panels was caused by debonding
of the stiffener from the skin and/ or pulling of the mechanical fasteners through the skin
and/ or attachment flange of the stiffener. It was shown that postbuckling performance de-
pended upon the stacking sequence of the skin laminate, and in the case of the 19-ply skin
specimens also on the mechanical fastener concept. Similar observations were reported in
[14.42].
Recently a series of tests on stiffened composite shear panels was conducted at the Co-
operative Research Center for Aerospace Structures Limited, Victoria, Australia (see [14.122]
and [14.134]). A picture frame was also used in these investigations for loading the panels.
The tests of [14.122] dealt with the design and postbuckling performance under edgewise
shear loading of the blade and !-integrally stiffened panels presented in Figures 14.83 and
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1151

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
(0 )

( D)

figu re 14.115 NASA Langley poclurc frJme for loading of J·\loffened shear paneh (from 114.131]):
(a) fronl view, (b) hack view

14.84, which were also used for axial compress io n posthuckli ng behavior studies (sec dis-
cussion in Subsection 14.2.2a). The experiments were accompanied by MSC/NASTRAN
fi nite-clement analysis. and good agreement between the result~ of the finite-element analysis
and experimental observation' was reported. In particular. it was emphasiLed that the po~t­
buckling behavior was well predicted (see (14.122]). This concl usion wa~ drawn from com-
parison of the shadow moire fringe panem of the buckled surface with the out-of-plane
displacement contours yielded by the finite-element analysis. The panel strengths, however,
were not wel l predicted using a maximum strai n. initial failure criteria. It was suggested that
this could be due to effects of interlaminar shear and throug h thickness no rm al stresses,
which were not considered in the ana lysis. Therefoo·c, the measured failure loads were sig-
nificant ly lower than the predi cted ones. It was also emphasized in [ 14.122)that ideal edge
restraints cannot be assumed for the design of postbuck ling structures. The postbuckling
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1152 Composite Structures

ALL DIMENSIONS IN CM.

Figure 14.116 Details of load introduction method into NASA Langley J-stiffened shear panels (from
[14.131])

response was again found to be very "sensitive" to the imposed edge conditions, indicating
that they had to be accurately introduced in the finite element modeling.
The second test series [ 14.134] involved experimental and theoretical investigations of the
postbuckling characteristics of the three-blade-stiffened composite shear panels shown in
Figure 14.121. These three-blade-stiffened panels were manufactured from T300/914C car-
bon fiber-epoxy unidirectional, pre-impregnated, zero-bleed tape. The stiffeners were sym-
metric laminates consisting of 16 plies (total thickness ~2 mm) with a ( -45/ +45/90/0 2 /
-451+45/0)s stacking sequence. It can be seen in Figure 14.121 that the top four layers of
the skin form part of the stiffener, with the purpose of improving the interface strength.

15

5trafn ri!'>'Hsal
buckling 1oid

20

15

..- 8uckl ing strain


10 1

.OOOJ .0006 .0009 .0012 .0015 .0018


S"e•r strain, y

Figure 14.117 Determination of shear buckling load of a Davis's J-stiffened shear panel by the strain
reversal method (from [14.131])
--`,`,`````,`,````,,`

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1153

2.0 p
Lw.d distr1b~tion
shown along this line Y

1.5 -

' Appllededge ('-

1.0
' ,',\ displacement\

(_-:.-==--=--=--"""·--~=---::_-_-_-:_
\ £'"''"' Unif~rm edge

_-_-::: _-_-_-_-:_ -::::_-:..-:.. -=~~~ -~~'-"")


,I
1 1

',

Test J
p "Z_ 0 0
Specimen and
test fixture model

.5

.2 ·4 y/:P&Ml length ·6 .6

Figure 14.118 Comparison of shear load distribution along the center section of a shear panel observed
in the Davis' shear tests with SPAR finite-element predictions (from [14.131])

Hence, as in the panels of [14.120] and [14.121], alternate bays of the panel had skin
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

laminates that were about 1 mm thick, either eight plies ( -45/ +45 190/0)s symmetric or
eight plies ( +45/ -45/90/02 /90/ +45 I -45) unsymmetric. In the latter bays extensional
bending coupling occurs. Therefore, in-plane loading results in in-plane bending of the un-
symmetric bays. The clamping regions of the panels were built up to 1.5 mm thickness,
consisting of 12 plies with (-45/+45/90/0/-45/+45), stacking sequences. To assist the
load transfer under the stiffeners a ( +45/ -45) doubler was positioned under the stiffeners
(see Figure 14.12lb).
A point worth noting in the test procedure of the panels is the insertion of antislip strips
between the panel edges and the picture frame edge members to achieve effective load
transfer from the rig into the panel through frictional contact. Also, in all of the tests the
panels were placed in the picture frame in a manner that applied the uniaxial tensile load
along the direction of the -4SO fibers.

2,0

1.5
-
'

< 1,0

0 0

~
c•
Tc;l/ ~Specirec fixtJre
,5

,2 .4 ,E .8 1.0
x/Panel 1-1idth

Figure 14.119 Comparison of shear load distribution across the middle of a shear panel observed in
the Davis' shear tests with SPAR finite-element predictions (from [14.131))
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1154 Composite Structures

I ±45/+4:i!Oz1 ±45io /90/0 1 5 cap


6 2
r ±45/+4'i/0/±451 s web

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
• 2-stiffener web has cap
width of 3.05
ALL DIMENSIONS IN CM.

Figure 14.120 NASA Langley stiffened graphite-epoxy shear web-stiffener geometry and stacking
sequences (from [14.132])

The tests were accompanied by finite-element analysis with the MSC/NASTRAN code.
Very good prediction of the buckling load and of the buckling mode was reported in [ 14.134].
However, the postbuckling strains and out-of-plane displacements calculated by the finite-
element analysis tended to underestimate the experimental results. A noteworthy observation
not predicted by the finite-element analysis was the gradual progression from one buckled
mode to another without "snap-through", which was reflected by the continuous decrease
in half-wavelength with increase in applied load. It was argued that this was due to a strong
influence of the geometric nonlinearities. Therefore, it was suggested that to predict the fully
developed tension field behavior of shear panels, analytical methods that use incremental
loading and take into account the effect of geometric nonlinearities should be employed.
Insofar as failure of the panels was concerned, accurate prediction of the first ply failure
load and location, using the maximum strain failure criteria, was also demonstrated in
[14.134]. Failure initiated on a buckle peak, on the stiffener side of the skin in a -4SO ply
of an unsymmetric bay. The failure mode was longitudinal tension. Ultimate failure of the

...I ..,.

~~
------+--+~--+
l:·I ~~
...r• ~~' i ~

- - - - - - - .;_.o..l_______r.
I• ;

/
r-
(a)
Figure 14.121 CRC-AS stiffened composite shear panels (from [14.134]): (a) three-blade-stiffened
panel dimensions. (b) details of integral blade-stiffener lay-up
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Flat Panels 1155

panels was through fiber failure, fiber pull-out and delamination. It was pointed out that the
ply buildups along the clamping areas of the panel to the test rig prevented failure initiation
in these regions.
Additional investigations comparing experimental and analytical results of the post-
buckling behavior of stiffened composite shear panels were reported in [ 14.135] and
[14.136]. In the former a first-level, simplified mathematical treatise based on strip analysis
technique to predict the shear postbuckling behavior was presented. It was shown that the
interlaminar and substructure (stiffener) separation forces can be adequately assessed once
the shape of the buckled panel surface and the induced surface strains were established by
the strip type analysis.
In the latter study the STAGS C-1 code [14.88] was used to analyze the behavior of
stiffened shear panels tested at the Grumman Corporation. The study aimed at providing
detailed analysis information that could be used to develop simplified analysis tools needed
to design postbuckled shear panels, which operate at loads several times their skin buckling
load. In particular, the study focused on the failure of such panels, resulting from disbonding
of the stiffeners (stiffener pull-off) due to diagonal tension effects. It was shown in [14.136]
that the STAGS C-1 predictions were validated by the test results and agreed well with the
results for the out-of-plane displacements and strains observed in the tests throughout the
load range till 15 times the skin buckling load applied in the test. It was therefore suggested
that the physical insight into the postbuckling behavior gained by the results of the STAGS
C-1 analysis could be used to assist in the development of simplified analysis methods for
the postbuckling behavior and stiffener disbonding failure mode of composite shear panels.

c. Combined Loading
The buckling and postbuckling behavior of unstiffened panels subjected to multiple loading,
and in particular to biaxial loading, was discussed in Subsections 8.6.2 of Chapter 8 and
14.2.1 d. In studying the buckling response of stiffened composite panels under biaxial load-
ing, additional aspects have to be considered (see for example [14.123]):
1. The load applied in the direction of the stiffeners is introduced along the centroidal axis
of the stiffened panel transverse cross-section (perpendicular to the loading and stiffeners
direction), whereas the transverse load (perpendicular to the stiffeners), applied or re-
active, is applied to the skin of the panel, i.e. eccentrically with respect to the load
aligned with the stiffeners. This produces couples along the sides of the panel parallel
to its stiffeners, leading to a particular behavior.
2. The postbuckling mode of the skin between the stiffeners is significantly affected by
introduction of the transverse load perpendicular to the load along the stiffeners. Whereas
under uniaxial loading buckling is associated with the development of several half-wave
buckles along the skin between the stiffeners, only one half-wave buckle develops along
the skin when the panel is subjected to biaxial compression loading.
As in the case of biaxial loading of unstiffened panels, experimental studies of the post-
buckling behavior of stiffened panels are rare. In fact, apparently no such tests have been
published. Only recently has an analytical and experimental investigation on the postbuckling
behavior of graphite-epoxy blade-stiffened panels subjected to eccentric biaxial compression
been carried out, by Romeo and Frulla in Torino. The unpublished results of this study
appear in [14.123].
Two blade-stiffened panels, N.l and N.2, fabricated from different prepregs M90/914
Vicotex and T300/F263 Hexcel, respectively, were employed in the test program. The di-
mensions of the panels and the laminate stacking sequences of their skin and blade stiffeners
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`

are presented in Table 14.4. (Note that this table also provides the data on the blade-stiffened
panel N.l, which was also tested uniaxially within the framework of the investigation pre-
sented in [14.123] and which was discussed in Subsection 14.2.2a.) The panels were loaded
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1156 Composite Structures

Table 14.4 Panel and cross-section dimensions of the blade-stiffened panels (from [14.123])
Panel L *B Lay-up b,., t\k Lay-up h,., t"
mm mm skin mm mm stiffener mm mm
N.l 817 620 (+45,/0) +45/90,10,), 155 3.35 (+45,/01(1/ +45,) 35 5.09
N.2 925 622 (+45,/0)901/0), 155.6 3.35 (+451/0,h/+45,) 37 5.62

L = panel length between potted ends


B = panel width between bolt lines (see Figure 14.85)
b,, = stiffener spacing
t,, = skin thickness
h" = stiffener height
t" = sti!Jener thickness
*= total panel width-700 mm

under various ratios of compression biaxial loading in the special multiloading test rig of
Figure 14.85.
The buckling loads observed in the biaxial tests correlated well with the predictions yielded
by the POBUCK code, which compiled the analytical formulations developed in [14.70].
Plots of out-of-plane deflections of a biaxially loaded panel, corresponding to initial buckling
and postbuckling load levels, are shown in Figure 14.122. These plots represent (a) the
displacements measured along the centerline of the panel, parallel to the stiffeners, and (b)
across the centerline transverse to the above direction. Moire patterns of the buckled surface
of the panel are depicted in Figure 14.123. It is apparent from Figures 14.122 and 14.123a
that buckling was associated with one half-wave along the stiffeners. With further increase
of the load level, the number of halfwaves of the buckled surface changed to three, a rela-
tively long dominant one and two short ones close to the loaded edges of the panel (see
Figure 123b). This buckling behavior differs from that experienced with uniaxially loaded

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
panels, where many short, equal-length half-waves develop at buckling along the skin of the
panel in the stiffener direction.
The panels failed under loads that were about twice the magnitude of their buckling
capacity (sec [14.123]). Failure was characterized by stiffener-skin separation due to very
high peeling stresses developed at the corners of the stiffeners (sec Figure 14.124). These
high stresses developed because the stiffeners were too stiff to comply with the large de-
flections of the buckled skin of the panel. It was also observed that after stiffener separation
took place, explosive stiffener crippling occurred.

6.00...,
Si>XIAL (-1,-0.15)

= 368.0
u....u.;. 519.9
(kN)
4.00

E'
..§.

"'
0.00
()
-2.00 00000 2t.5.25 (k'J)
-2.00 343.35
/>.J:.lt./>.b
¢¢000 4E5.98
V"'tt~5i9.93

- 4 - ~ ao
-4.00 ,~.~~.,., 0
0.00 0.20 C.40 0.60 0.80 . 00 0.20 o.40 0.60 c.Ba· 1.00
X/L Y/8

(a) (b)

Figure 14.122 Romeo and Frulla's test results of a graphite-epoxy stiffened panel subjected to eccentric
biaxial compression loading--transverse load/longitudinal load (NJ NJ = 0.15 (from
[14.123], courtesy of Professor Romeo): (a) out-of-plane deflection along the centerline
parallel to stiffeners, (b) out-of-plane deflection across the centerline transverse to the
stiffeners
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Wing Box Structures 1157

(OI {b l

Fit;ure 14.123 Buckling pauems under different load levels observed in Romeo and Frulla's tests on
graphite-epoxy stiiTencd panels under eccentric biaxial compression loading (from
[14. 123). courtesy of Professor Romeo)

14.3 Wing Box Structures


T he previous sections of thi s chapter as well as most of the other chapters deal primarily
with structural models. real or scaled. typical of elements of a whole structure, which are
subjected to prescribed " ideal " in-plane loading and boundary condi tions. In the tests wi th
such isolated elements. attempts are usually made to sim ul ate as closely as possible nominal
simply supported, clamped and free boundary conditions. which in many cases are unrealistic
for the actual structural element that is simulated. Under these circumstances, the tests on
the elements serve mainly as basic experiments with the purposes of ( I) verifying the validity
of theoreti cal tools; (2) revealing phenomena unaccounted for in the development of the
theoretical tools. and (3) gaining better physical insight into the role of the parameters. wh ich
dominate the behavior of the loaded structure. These tests. therefore, provided the means
and data for improvement of the theoretical tools, for development of semi-empi rical for-
mulae and design charts for complex type structures and for ca libration of undete~mi ned
parameters. pruticularly in nonl inear analyses.

Figure 14.124 Stiffener-skin separation failure observed in Romeo and Fruna·s tests on postbuckling
behavior of graphite-epoxy stiffened panels under eccentric biaxial compression load-
ing- transverse over longitudinal load ratio (NJ N,) = 0.30 (from [ 14.123]. courtesy
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

of Professor Romeo)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1158 Composite Structures

When real structures, such as wing box beams, are exposed even to simple types of loading
like bending or pure torsion, the ideal type boundary conditions and in-plane loading cease
to exist. In a wing box beam subjected to bending, the boundary conditions along the edges
of either upper or lower skin panels are not well known, because the other substructures of
the beam can induce deformations along the edges of these panels. Furthermore, curving of
the beam due to application of the moment and the subsequent development of longitudinal
in-plane forces in the upper and lower skin panels introduce loads perpendicular to these
panels, resulting in a crushing pressure (see Figure 14.125). This crushing pressure can
significantly affect the deformation field of the panels and (as is apparent in the figure) lead
to in-plane warping of the cross-section of the wing box. The warping of the cross-section
may atiect the compression skin panel, which is susceptible to buckling. Due to its becoming
a curved panel (see Figure 14.125), its local buckling capacity is increased as compared to
the unloaded flat panels. However, the in-plane warping and the subsequent reduction of the
wing box cross-section reduce the cross-sectional moment of inertia and subsequently the
moment-carrying capacity of the wing box beam.
A distributed load perpendicular to the upper and lower skin panels of a wing box beam,
which is associated with torsion curvature, also develops when subjecting the beam to pure
torsion. This results in subjecting the panels to a lateral pressure and in-plane warping of
the wing box cross-section in a manner similar to Figure 14.125. Furthermore, when the
panels of the wing box beam are in a postbuckling state, their effective shear modulus
diminishes due to the development of the diagonal tension field. Consequently, this reduction
in stiffness, together with the in-plane warping of the wing box cross-section, causes con-
siderable nonlinear reduction in the wing box torsional stiffness when the beam is subjected
to increasing torsion.
It should further be noted that the use of laminated composite materials introduces cou-
plings, e.g. bending-twisting, or extension-twisting couplings [see Eq. (14.1)]. The significant
shearing and warping of the beam, out-of-plane and in-plane, which are inherent to laminated
materials (due to the high ratio of extensional modulus to the shearing modulus), and the
presence of couplings considerably influence the overall behavior of the beam, as well as
the stress-strain distributions.

The Turin bending tests - Extensive theoretical and experimental studies on buckling and
postbuckling behavior of wing box beams subjected to bending and torsion loadings were
carried out by Romeo and his colleagues at the Department of Aerospace Engineering, Turin
Polytechnic, Italy (see [14.116], [14.137] and [14.140]). Their impressive test programs
aimed at validating their analyses and the applicability of their computer codes for prediction
of the buckling behavior and response of the wing box beams.
The Turin bending tests were performed in the specially designed bending testing machine
depicted in Figure 14.126. This apparatus consisted of a frame A, two spar clamps B, hinged

~ ~
~I ~
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`--- b

Figure 14.125 Crushing pressure in wing box structures under pure bending (from [14.138], courtesy
Copyright Wiley
Provided by IHS Markit under license withof Professor Romeo)
WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Wing Box Structures 1159

Fl~:u rc 14.126 The Turin wing box 1->cnding testing machine (fmm [ 14.138], cout'tcsy of Professor
Romeo)

through bolts to four swinging link rods C. and two hydraulic jacks D. To fix the specimens
in the bending machine. suitable linings were bolted to their ends, and then they were
anachcd through these linings to the clamps B. The hydraulic jacks were hinged to the other
side of the clamps. so that the bending moment could be applied step by step.
Deflection data along and across the surface of the compression panel. at different load
levels. were monitored by a dial deflectometer that moved approximately 430 mm along a
spec ial gui de independent of the box structure. The clellections were recorded by an X- Y
recorder connected to the dellectomctcr. Longi tudi nal struins were measured by back-to·back
strain gages bonded to the compression and tension panels at panel half-length .
Two wing box specimens wi th cellular cross-section panel~ (sec Figure 14.127a), which
>imu latcd an integral ribless wing. were tested in the 1986 test program (14.1161. These
specimens were fabricated from T300/F155 graphite-epoxy material prepregged (made hy
Excel Belgium). Each panel was reinforced at its end~ by 140-mm wide glass-epoxy exten-
sions (~ce Figure 14.127a). A third wing box specimen that was tested within the framework
of the 1986 program. as well as of the tests reported in [ 14.137]. had unstiffened (0:90)
cross-ply orthotropic skin punels (sec Figure 14.127b). A graphite-epoxy wing box beam

.........

H (a)
I
l..c·•i";t"O. ~

t!"'i'-r'!.,,..,.., ••-

...'
'"~

I
·l ...
).tJ

(c )

·~

ALL DIMENSIONS IN MM.


--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,

Fig ure 14.127 The Turin bending tests wing box specimens: (a) cellular cross-section dimcn,io n and
lay-ups (fro m ( 14.116]), (b) unstiffened ct·oss-ply 011hotropic skin cross-sectio n-
stacking sequences. dimensions and strain gage locat ion ( from [ 14.1371. courtc<y of
Professor Romeo). (c) blade-stiffened cross-sectio n dimensions. laminate configura-
tions and strain gage and dcncction measurement locations (from [14. 138). counc>y of
Copyright Wiley
Professor
Provided by IHS Markit under license with WILEYRomeo) Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1160 Composite Structures

with blade-stiffened skin panels (see Figure 14.1 27c) was investigated in the more recent
tests (see ( 14. I 37] and [ 14. I 38)). Its manufacturing details are depicted in Figure 14.128.
One may note in this figure the short rectangular profi les, which were bonded to the stiffeners
and skin to achieve beuer clamping of the panel ends to the 30-mm thick steel fittings.
The test results of the studies conducted by Romeo and his colleagues demonstrated the
effect of the bending curvature that induced a lateral pressure leading to in-plane warping
of the wing box cross-section. as shown in Figure I 4. I 25. This is apparent from Figure
14. I 29, which shows the longitudinal deflection along the middle section and side section
of the wi ng box compression panel. As observed in Figures 14. 130 and 14. 131. which
represent the maxi mum deflection of the compression panel and the stmins. respectively (see
Figure 14.127c), the test results (which were also repon ed in [ 14.1 37] and [ 14. I 38]) correlate
well with the theoretical predictions by the POBUCK code [1 4 .70], when considering the
effects of the lateral pressure. Note that in this figure ETB designates engineering theory of
bending, which does not account for the lateral pressure. Figures 14. I 30 and J4. I 31 also
re flect the influence of the induced lateral pressure on the hehavior of the panel. Due to its
existence, the deflection at the center of the panel was significantly larger than those along
the sides of the panel (Figure 14.130), and it exhibited strong nonli nearity. Figure 14 .131
reveals that with increase in bendi ng moment , load in the stiffeners of the compression panels
changed from compression into tension, thus drastically altering the hypothesis that the panel
was subjected to uniform longitudi nal compression loading. The argument is emphasized in
Figure 14. 132, which represents the transverse strain distribution experienced by the stiff-
eners of the compression panel and side webs of the wing box (see Figure 14. 127c for gage
location).

The Turin torsion experimems - Romeo and Frul la's torsion tests (sec [ 14. I 39] and [ 14.140])
were carried out in the torsion machine depicted in Figure 14.133. T he machi ne consists of
a basement A, at the four corners of which two co lumns B. one connecting rod C and a
hydraulic jack I are placed. The external load applied by the jack is converted into a twisting
moment by the two stiff plates £. The wing box is attached to these plates by thin. flat

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 14.128 Blade-stiffe ned wing boxes with proOics bQnded to the ends for improving clamping
Copyright Wiley
behavior of panels tested in the Turin
Provided by IHS Markit under license with WILEY
bending tests (c(mrtesy of Professor Romeo)
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Wing Box Structures 1161

15

Bending Moment (kNm)


GeeeEI 0.0
..-7.18
E' 10
..s
~14.36
-17.95
+-+-+-+-+ 2, .5-4
-25.13
z -28.72
0 .......,. 32.31
B -35.9
~
w 5
a

Is

~a 0~"o0o
0 200 400 600
LONG. LATERAL SECTION (mm)

Mb=35.9 ( k Nrnl

Figure 14.129 Longitudinal deflections along the middle and side sections measured in the tests of
Romeo and Frulla on a blade-stiffened wing box (from [14.138], courtesy of Professor
Romeo)

aluminum plates, which have high in-plane stiffness and very low out-of-plane stiffness, and
which are bolted to the ends of the wing box beam. The plates E are connected to the
loading frame by ball bearings F and floating rods G, allowing the box to be free of warping.
Two types of wing boxes were used in the Torino tests. The first type (depicted in Figure
14.134b), which was employed in the tests of [14.139], was fabricated from graphite-epoxy
prepreg M40/940 VICOTEX. It was manufactured in two parts, each one composed of a

40

30

10 ........., PRESENT THEORY (middle)


..........,. EXPERIMENTAL (middle)
~ ETB (lateral)
- EXPERU.1ENTAL (lateral)

10 15
w (mm)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 14.130 Comparison of the maximum deflections observed in the Turin tests on a blade-stiffened
Copyright Wiley
wing box with POBUCK predictions (from
Provided by IHS Markit under license with WILEY [14.138],Inccourtesy
Licensee=McDermott of Professor
- India/8215328006, Romeo)
User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1162 Composite Structures

~0

30

20
E'
z
-=-

-s.G.1
-S.G.2
........., PRESENT THEORY

-2 00-2000 -1500 -1000 -500


STRAIN (E -6)
(a) (b)
Figure 14.131 Comparison of longitudinal strains measured in the Turin tests on a blade-stiffened wing
box with PO BUCK predictions (from [ 14.138], courtesy Professor Romeo): (a) along
side blade-gages 1 and 2 of Figure 14.127c, (b) along center blade-gages 5 and 6
of Figure 14.127c

skin panel and two webs that were cured in an autoclave under a controlled pressure cycle,
and bolted along the webs to one another. This wing box had the following dimensions: L
= 720 mm, a = 360 mm, B = 396 mm, H = 133 mm, t, = 2 mm and t .. = 4 mm (see
Figures 14.l34a and b). The second type of wing box (Figures 14.134a and c) was used in
the tests of both [ 14.139] and [ 14.140]. Tests with one wing box of this type fabricated of
M40/940 VICOTEX with the dimensions L = 768 mm, a = 384, B = 400 mm, H = 135
mm, t, = 2 mm, t" = 4 mm, tF = 3 mm and tv = 2 mm were reported in [14.139], while
experiments with two additional wing boxes of this type, one made of M40/940 VICOTEX
and identical to that of [14.139] and one fabricated from graphite-epoxy T300/F263 HEX-
CEL, with a lay-up of ( -45 2 / 45 2103!90)s where t.1 = 2.35 mm, tw = 4.62 mm and tF = 5
mm, were reported in [14.140]. These wing boxes consisted of four panels that, after their
autoclave curing, were bolted along their entire length to the corner L 30 X 30 X tF mm 3
shaped aluminum flanges (tF = 3 mm for the VICOTEX and tr = 5 mm for the HEXCEL).
The geometry of both types of wing boxes is defined in Figure 14.134a. Nate in Figures
14.134b and c that the thicknesses of the webs of the wing boxes were twice the thicknesses
of their panels.
Both types of wing box specimens were first tested free of ribs. After completion of the
experiments with these specimens, ribs were bolted on at half-length of the wing box beam
(Figure 14.134a) and the beams were retested.

1000

U)
I 0
~
z
~ -1000
tii
;;/,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

:;r; -2000
0
1:!
G
i5...J -3000

Figure 14.132 Transverse strain distribution under various moment levels observed in the blades of
Copyright Wiley
the compression panel and side webs of the blade-stiffened wing box tested in the
Provided by IHS Markit under license with WILEY Torino bending test (from [14.138], courtesy ofIncProfessor
Licensee=McDermott Romeo)
- India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Wing Box Structures 1163

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Fi~:ure 1-1.133 Tile Torino tor--ion rnachone ror wing bo•e' Umm )14.139]. counesy or Pnofe"nr Ro·
meoJ

A twl\Ung moment w;l\ apphcd to the beams step by step. Deflection data of the lower
box panel were obtained at 48 stations under various load levels by an electric transducer
mounted on a controlled carriage (Figure 14. 133). Strain> were monitored by li near and
rm.cttc strain gages bonded hack to hack o n the upper and lower panels at a quarter, hal f
and three-quancrs of the beam length.
Very good correlation was obtained between the huckhng loads experienced in the tests
114 U9) and [1-U-WI and those detennined by the ALPATAR code developed at the De-
panment of Aerospace Engonecnng. Turin Polytechmc as \~ell as those yielded h) POBUCK
114.701. respective!). Vef) good <~greement bet\\een the e•perimentall) obscr.cd response
or the wong boxe>. as rellectcd hy thetr angle oft\\ ist. str<lins and deflection pattem, and the
theoretical predictions was abo reported. This is apparent from Figure 14. 135, which depicts
a compari~on between the angle of twist measured in the tests and the theoretical pred ictions.
Note that this comparison corresponds to a wing box <lf" the lype shown in Figure 14. 134c
without a rib (see Figure 14.135a) and to one including a rib (see Figure 14. 13511).
Furthennore. it is apparent frmn Figure 1-t 135 that the buckling of the panels dra,tically
reduce' the effective torstonal ,llffne\~ of the structure as a whole when operating under an
incomplete diagonal shear 'trc" field and due to the sigmficant warping of the wing box
cru"·'cctton caused hy the induced lateral crushing prc"urc
Typtcal deflection pattern' of the lower panel of the wmg box. under differcm levels of
the arplicd torsion moment that were observed in the te't of a wing box of the second type,
containing a rib. are shown in Figure 14.136. The inll ucncc of the buckling pattern on the
defl ect ion of the panel is apparent from this figure .
Additimwl studies 011 fiOWimcklillfi of composite bo1 .If rue //Ires can be found in 114.141]
and 114.142]. The first of these paper~ deals with the devclopmcm of design and con,truction
Copyrighttcchnoque
Wiley of a postbuckled carbon fiber reinforced pJa,tlc wing box for an acrobatic light
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1164 Composite Structures

(b)

(c)

figure 14.134 Wing boxes used in the tests of Romeo ctul. (from 114.139) and with intermediate rib,
dimension' 114.1401. counesy of Pmfc'""' Romeo): (a) general view of IO"'ion wing
box. (b) configuration and geomctncal data. of first type of wing box 114.139]. (c)
10rsion wing bo• "ith "L" comer>. configurmion nod geomelrical data

12
o•
~ (•~~tol

::::: ~;:::nt lii• O'Y - t.o--rlf"'~tol

::::: ~!.~ttl Tl!.ory

• (b)
6
TORSION MOM(NT ~ (~)
10 12

(a) fO!itSION' tro~O~o~ tNT t.l, (1o1tm)

figure 14.135 Comparison between the angle of twist mcn~urcd in Romeo ct al.'> wing box torsion
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

tests and theoretical predictions (from [ 14.1391. counesy of Profe,~or Romeo): (a) wing
Copyright Wiley box free of rib. (b) wing box with Licensee=McDermott
Provided by IHS Markit under license with WILEY
a rib bolted toIncthe center of the beam
- India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1165

lr.4t=1 0.08 k.Nm

Figure 14.136 Typical deflection patterns of the lower panel of a torsion wing box observed in the
tests of Romeo et al. (from [14.139], courtesy of Professor Romeo)

aircraft. The concept lead there to a single cell torsion and bending box, with stiffened skin
panels to sustain the major loads. The second paper [ 14.142] studied the postbuckling be-
havior of a box-like substructure, of a carbon-epoxy composite rudder for a fighter aircraft,
by means of a finite-element analysis. The test program reported in that investigation aimed
at verifying the prediction capabilities of such analysis. Good correlation between theory
and experiment was obtained.

14.4 Curved Panels and Shells


14.4. 1 Unstiffened Panels and Shells
a. Theoretical Considerations
As for isotropic shells, shell theory of laminated composite shells is governed by a set of
eighth-order equations. In laminated composites, however, the relatively large number of
stiffness coefficients involved makes these equations algebraically more complicated than
those associated with isotropic shell theory, in which there are only two coefficients, but
which are still quite complicated by themselves. Furthermore, the definition and solution of
a buckling problem with these equations, irrespective of the shell material, requires that four
boundary conditions be prescribed along each edge of the shell.
In laminated composite shells the above-mentioned intricacies associated with shell theory,
as well as other secondary parameters that have to be considered, such as material anisotropy,
radial inhomogeneity, ditierences in tension and compression elastic moduli inherent to lam-
inated composite shell skins, various types of loading, end constraints, geometric shape
imperfections and material nonlinear stress-strain behavior, significantly complicate the de-
velopment of an appropriate shell theory. Hence, the calculation of stability criteria of lam-
inated composite shells becomes a formidable task and a great challenge. The design of
laminated composite shells with these computational tools therefore requires a sound exper-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
imental database, obtained from carefully and sophistically conducted tests. These experi-
ments have not only to adequately validate the analysis tools but also to account for the
Copyright Wiley
Providedinfluences oflicense
by IHS Markit under thewithabove
WILEY mentioned secondary parameters.
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1166 Composite Structures

Derivation of laminated composite shell buckling equations is thus a complicated proce-


dure. Depending upon the assumptions made, various forms of final equations (shell theories)
are obtained, and the results for critical loads may significantly differ according to the theory
applied. An example of a "classical" -type theory (for which relationships between stress-
strain and strain/displacements are assumed and a membrane prebuckled shape is considered
to satisfy the boundary conditions, partially or completely), a set of perfect cylindrical shell
buckling equations was presented by Wiswanathan, Temekuni and Baker (see the survey of
[14.5]). Theoretical and experimental investigations on buckling of laminated cylinders, ad-
dressing the issues of material and surface geometry imperfections and boundary conditions,
and their influence on the buckling load capacity of the shell, as well as on the correlation
with theory, were reviewed and evaluated by Tennyson in [14.6]. Detailed derivations of the
equations governing the equilibrium and buckling behavior of perfect laminated composite
cylinders, within the framework of classical buckling theories, can be found, for example,
in [14.143]-[14.150].
Following Vinson and Sierakowski [14.151], the buckling loads of a composite circular
cylindrical shell (of radius R, length L and wall thickness h) are given below for various
types of loading, in compliance with the following assumptions:
1. Special anisotropy (A 16 = A26 = 0)
2. Prebuckling deformations are not taken into account
3. Ends of the cylindrical shell are supported by rings rigid in their planes, but no resistance
to rotation or bending out of their plane

Axial compression- General case, no mid-plane symmetry, and n > 4:


ell el2 el3
en
r
e21 e22
e,l e,2 Cn
N-
( -
L ( 14.21)
.\n m1T
I ell el21
e21 e22
where
N," = critical compressive load per unit circumference
L = cylinder length
R = cylinder radius
m = number of buckle half-waves in the axial direction
n = number of buckle waves in the circumferential direction

ell= A11 (mL1rr + Ar,r, (~)" (14.22)


--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

e22 =An (~r + A66 (mL1rr (14.23)

(14.24)

(14.25)

(14.26)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1167

(14.27)

and A,i, Bu and Du are the stiffness coefficients of Eq. (14.2). The buckling load N," is
determined from the combination of integer values m and n, which make N," a minimum.
Special case, mid-plane symmetry (B 11 vanishes):

N~, U
1T-DII
= m2 (1 + 2Dl2
Dll
{32 + D22 {34)
Dll (14.28)

where
nL
{3=- (14.29)
TTRm
(14.30)

(14.31)

and y is an empirical (knock-down) factor that ensures that the calculated buckling load will
be conservative with respect to all the available experimental data. The buckling load is
again obtained by employing the combination of m and n, which make N," a minimum.

Buckling in bending (Bu = 0)- In this case Eqs. (14.28), (14.29) and (14.31) are used, but
with y given now by:
(14.32)

The procedure of determining N," is identical with that employed for the previous loadings.

External lateral pressure and hydrostatic pressure- The critical lateral external pressure is
determined by:
ell cl2 Cn
c21 c22 en
c,l c,2 c,,
R
Per = --; (14.33)
n-
Iell
c21
cl21
Cn
To find Per' set m = 1 and vary n(n ::::: 2) to obtain the minimum value for Per·
For long cylinders subjected to lateral pressure, the critical buckling pressure is given by

(v - Bi2)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

3
22 An
Per (14.34)
R'
In the case where B 11 = B 22 = B 12 = Boo = 0, Per is obtained from
114
= s.5.13 [m2(A11A22 - Af2)] (14.35)
Pa LR'I2 An
which is valid only when
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1168 Composite Structures

12
D22)'
(Dll (A11An-
12422DII
AT2)
112
(U) >
R
500 (14.36)

For hydrostatic pressure, Eq. (14.33) is again employed, but with n 2 replaced by

n2 + ~ (m~Rr (14.37)

in the Cu terms Eqs. (14.22)-(14.27) and using the combination of m and n that yields the
lowest buckling pressure, Per·

Torsional load (Bu = 0) - The critical torque is given by:


A A A2 )3/H RS/4
T = 21 75(D )"/H II 22 12 -- (14.38)
(T • 22 ( A22 LI
12

provided that

(DDll ) (A 12A22DII

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
516 112
22 A 22 - A'l2 )
II L'
_..:... 2: 500 (14.39)
R

Postbuckling behavior and imperfection sensitivity- As with isotropic shells, classical shell
theory usually appears to be adequate for treating torsional and external pressure buckling
behavior for practical end constraint conditions (see for example [14.6]). However, as in the
case of isotropic shells, geometric shape imperfections can significantly degrade the axial
buckling strength of laminated shells and affect the correlation between theory and experi-
ment. Consequently, postbuckling behavior shell theories are required to assess the imper-
fection sensitivity of a laminated composite shell and to determine the effect of imperfection
amplitude and distribution on its buckling behavior and load capacity. In developing such
imperfect shell theories, a major difficulty in the estimation of the load reduction due to
imperfection is the lack of reliable imperfection data that could also provide the basis for a
statistical approach (see discussion in Sections 10.10 and 10.11 of Chapter 10). Obviously,
unless such extensive imperfection data are available, empirical knock-down factors have to
be resorted to.
While numerous studies are available on the postbuckling of isotropic shells and the effects
of geometric imperfections on their buckling, relatively few investigations have been under-
taken to study these in laminated composite shells. Postbuckling and shape imperfections in
shells and shell panels were addressed in the survey of [14.5] and the relevant investigations
reviewed there, while buckling of geometrically imperfect composite cylinders was reviewed
by Tennyson [14.6]. Most of the studies discussed there dealt with the application of Koiter's
imperfection shell theory (see Card, [14.152], who was the first to apply the theory as a tool
for assessment of imperfection sensitivity of laminated composite shells subjected to axial
compression). It was concluded in [14.6] that for axial compression and bending loadings,
the application of Koiter's approach provides reasonably accurate estimates of the buckling
loads. Additional details on postbuckling behavior and imperfection sensitivity of laminated
composite shells can be found, for example, in [14.153]-[14.155].

Transverse shear effects- In studying the buckling of laminated shells, the thickness of shell
skin relative to shell radii requires special consideration. Simitses, in his review on buckling
of moderately thick laminated cylindrical shells [ 14.156], states that the assumptions made
in the development of classical lamination shell theories (relatively thin laminates), or at
least in some of them, are not valid and have to be changed. Because present-day advanced
composites have low transverse shear moduli, transverse shear deformation plays a much
more important role in the kinematics of composite laminated shells than in homogeneous
metallic ones. Also, in some applications thicker walls are used to achieve the strength
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1169

equivalent to metallic ones, leading to the reduction of the ratio of radius of curvature to
total wall thickness.
Based on the studies and results reviewed in [14.156], as well as other ones, Simitses
concluded:
1. For thin (R/ h > 30) and medium-length (L/ R > 3) cylinders, classical lamination shell
theory (CL) yields accurate estimates.
2. For moderately thick (1 0 < R/ h ::; 30) cylinders, classical theory (CL) is not applicable.
For these geometries, regardless of the construction material, first-order deformation
theory (FOSD) with a modest shear correction factor (SCF) yields accurate estimates.
3. The stacking sequence that yields the strongest configuration (largest critical load) is
both R I h and L/ R dependent.
4. From the limited results of nonlinear analysis, it is observed that moderately thick cy-
lindrical shells are imperfection insensitive for axial compression.
He recommended that:
1. In calculating the critical load for a moderately thick laminated shell, the possible oc-
currence of a strength failure prior to buckling should be checked.
2. Study of delamination initiation and growth and their influence on the buckling load
should be included.
Another important aspect in composite laminated shells is the great influence of the lay-
up of the shell wall on the buckling loads and mode shapes of shells subjected to axial
compression (see Geier, [14.157]). It was shown there that the anisotropy of the bending
stiffness, while the in-plane stiffness was orthotropic, rendered interaction of axial compres-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

sion and torque in circular cylinders with balanced, but nonsymmetric skins. Furthermore,
it could affect the buckling behavior of panels subjected to axial compression and shear in
a nonconservative manner. It was concluded that in this situation a design that tries to exploit
the potential of composite materials requires careful analysis.
In the next sections experimental studies performed to validate shell theories and evaluate
the influences of various parameters on cylindrical shell panels and complete shells subjected
to various loading types will be discussed. First, tests on panels and cylinders subjected to
axial compression will be presented, followed by experiments on panels and cylinders under
torsion loading and then by tests on shells subjected to combined loading.

b. Axial Compression Experiments

Curved panels- Curved composite laminated panels constitute an important group of struc-
tural elements; they are also used as substitutes for complete shells in many test programs
because of the high cost and complexities involved in fabrication of complete composite
shells as well as in construction of their test facilities, and when the dimensions and load
capacity of the test apparatus are too small to meet the specifications for full-scale testing
of complete shells. The tests with panels can, however, simulate tests with complete shells,
provided that the panel geometry is appropriately chosen and the boundary conditions along
its edges are correctly prescribed.
Theoretical and experimental studies on curved laminated composite panels were discussed
by Leissa in his 1985 survey [14.5], in which the effects of curvature, boundary conditions,
axial wavelength and direction of laminae on the buckling under various loadings were
considered. Another noteworthy 1985 study is that of Hui [14.158], who investigated the
asymmetric postbuckling of symmetrically laminated cross-ply, short panels. It was shown
in this study that by introducing geometric imperfection in the shape that initially bulges
out, a cylindrical composite laminated panel could be designed to carry loads above the
classical buckling load, provided the imperfection amplitude was sufficiently small.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1170 Composite Structures

Wilkins [ /4./59] reported the rewlts of ll comprehensit•c 1e~1 program aimed at providing
basic experimental data on the buc~ling behavior of laminated graphite-epoxy curved panels
subjected to compres>ivc loading. a~ well a; ~pccific design data to verify orientation~ and
thicknesses chosen for a fuselage component. The test variable~ were thm, lamina oriem:uions
and lam inate thickness, whereas p~me l length . width, radius and material were held constant.
In the test program . curved panel~ 13.0 in. (330 nun) long by 9.0 in . (229 mm) wide (arc
length). having an out\ide radiu~ of 12 in. 005 mm), and of 23 laminate lay-ups [(0/90);,,;
( ± -t5),: (0/90)s: ( ± 45),: ( 45)4 ,; ( ±45),,: C- 45 h,: ( 45),,,: ( +30),: C+30)4 ,: C+ 30),,,;
(0)~: (0)"; (0),,,: (0;90).\S; (±30) ,; (±30b: (±30).,: (01451901 45),; (0/ ± 60),;
(0/ ± 60)": (0/ ± 45;0),: and (0/ ±45),] were em ployed. The panels were hand laid from
Morganitc n /4617 by laying up multiple specimens on a tahle. dropping them into a conca\e
steel tool. and then having them bagged and cured. Following this procedure. the !.traight
edges of the panels were trimmed on a spec ially jigged table saw and the curved edges were
trimmed with an end-mill. II is noteworthy thai no a11emp1 was made to fabricate "perfect"
panels for the test~. so that the specimens would exhibit the thickness variations common to
production pans made of the above material 'Y'Iem.
The panels were tested in the loading fixture 'hown in Figure 14.137. Th1;. fixture provided
clamped-clamped boundary conditions along the curved edges and ei ther si mple-simple or
clamped-clamped bou ndary condi tions along the straight edges (Figure 14. 138a and b. re-
spectively). Clamping bars accommodated the 'ariations in thicknesse\ of the panel~ c~ce
Figure 14. 138a and b). Variations in curvature and warpage of the panel~ were slight and
thus were corrected when installing the panel in the rigid loading rig. Parallelism of the
loaded edges of the panel was determ ined upon instal lati on (less than 0.003 in. [O.OH mm]
over the edge length) and corrected when necc~~ary.
To reduce shear loads stemming from friction generated between the ~uppons and the
specimen edges. each panel. before being 3.\\Cmblcd in the 1cq fixture. was bordered with
0.003-i n. (0.08-mm) thick metal. To obtain reliable and repeatable te~l results that would
clearly distinguish between the responses of one panel from another. a common procedure
for install ing the panels and aligni ng the setup for test run' wa~ employed. The bolls on the
unloaded edge suppons were hand tightened when ~imple \uppon condition., were used. and
60 in.-lb (6.78 kN-mm) wrench tightened for clamped suppons. ln each test the boiL~ were
checkeu ~1f1er applyi ng two low-load excursions, which were used to scat the panel and
remove most of the hysteresis. Followi ng the completion of the pane l in~tal l ation. it was
subjected 10 an axial load in a 120.000 lb (533.4 kN) Baldwin Universal testing machine.
Before testing the composite panels. aluminum pancb were tested m the rig of Figure
14.137 10 evaluate the actual edge restraint~ of the fixture. h was observed from the test
results that the clamping action on the speci mens' loaded edges was very ncar the cla~~ical
value; however, the si mple support~ provided sl ightl y more than classical rc,lraint. an exec~~

Figure 14.137
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`--- Wil~ins's comprc,~Jon loading fiAture ofcuf\OO panch !from [14.159)l
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1171

(a)

(b )

Figure 14.138 End-fixtures used to simulate boundary conditions along the side edges of the curved
panels tested by Wilkins (rrom [14.159]): (a) simple-simple. (b) clamped-clamped

of 10 in.-lb ( 1.1 3 k.N-mm) edge moment per radian per in. of length. This value was deter-
mined to be within acceptable lim its, and thus the tests proceeded without further alteration
in the setting-up procedures.
The responses of the panel, o ut-of-plane deflection versus axial load that were required
to dete>mi ne the buckling load and panel behavior were monitored by two methods, by a
linear differential transformer whose output was recorded on a machine-mounted x-y drum
recorder, and by the moire shadow technique.
It is worth noting that the panels were fabricated lO provide one test data point each.
However, before testing began it was rea lized that the nondestructive Southwell plot method
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

was applicable to shell-type structures as well, prov ided knowledge of where the first buck le
would form was furnished in parallel (see also Section 9.6 of Chapter 9). This led to intro-
d uctio n of the moire g rid shadow technique for full-field displacement monitoring. Both
methods, the moire technique prov iding a lower-bound buckling criterion and the Southwell
method providi ng an upper-bound one. were applied to furnish nondestructive buckling es-
timates for each panel. Some panels were allowed to snap throug h. Both methods were
simu ltaneously applied, and whi le the moire pallerns were observed. a plot was made of the
out-of-plane deflection measured at a point on the opposite face of the panel (see Figure
14.137). This plot was used to obtain the Southwell plots (an example of such a plot,
reproduced from [ 14.159). is shown in Figure 14.139). The Southwell plot used in this study
was a modification of that suggested by Tenerelli and Horton for metal shells (see S ubsectio n
9.6.2 in Chapter 9. and [9.202)).
It was reported in [ 14.159] that the moire procedure. used to determine the lower buckli ng
loads of the panels, proved to be effect ive in saving the majority of the panels for further
testing. When loading of the panel was termi nated at initial evidence of buckling, subsequent
loading cycles produced repeated results. Due to the success of the moire technique in saving
the panels, 14 panels out of the 72 tested in the program of [14. 159] were tested with clamped
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1172 Composite Structures

Nott: eu, ....,&ppH~td .....u E.ltobllthtd ot


"llO lbs, t:fl.rtfon, Oato RtMtlofl
Stortt d ot this Po"!.

!~ 4000 10 20 30
..
.i! JOOO
Dote f« Southwfll P\ot ·

I PI 161 (t) IPc~t t Molrt


• 11,300 tb5
• 6,300 lbs
!Pelt) Snop tood • 6,3~ tbs
• so so
S22S 10 ' 0. 00)6
0 .0019 (PcJt ) Southwt tl ~ 11000 lbs

'"'
5725 ""20 0 .0026
o.oo :u
SS IO
6<)10
6110 ,." 0 . 001'
0,0040
o.oo•s
Boundar y COt'lds: cess
&280
6 JOO "" O.OOSJ
o.oos•
20 lO •o so 16 32
"
(a) Ori\t¢1lon, OMtiMI (b ) 6, 01¥ "
Figure 14.139 Typical load versus deftection response observed in Wilkins's compression tests of
curved panels and application of the Southwell method (from [14.159])

sides, after all the panels had fi rst been tested with simply supported sides . Obviously, this
capability of repeated testing represented a major contribution of the test program of
[14.159]. As stated by Wilkins, it established a nondestructive test method for testing ques-
tionable production parts whose subsequent performance could be crucial to the life of the
structure. The test results of the panels, simply supported along their sides, were correlated
with theoretical classical buckling loads calculated, using an in-house program developed at
General Dynamics, Fort Worth, Texas (see [ 14. 159]). An imperfection sensitivity analysis
was also perfonned, and two types of knock-down factors were presented: an experimental
one, the ratio of the snap-through load to the class ical buck li ng load or the moire load. in
case no snap-through occurred: a theoretical one based on a modification of the imperfection
sensitivity analysis of Tennyson et al. [ 14.160] and [ 14.161). In the imperfection sensitivity
analysis of 114.159], because of practical considerati ons. the standard deviation of the panel
thicknesses over the shell (see Figure 14. 140), rather than a precise axisymmetric shape
imperfection, employed by Tennyson and hi s coll eagues, was used as a measure of imper-

.
0. 08 .

. ..
~
~
~
~

~ -
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

0.04
$
8
0,(12

0
0

Figure 14.140 Thickness coefficient of variation as a function of panel radius to thickness ratio Rl t
of the curved compression panels tested by Wilkins (from [14. 159])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1173

fection. Furthermore, the knock-down factor was assumed to be applicable to any partial
cylinder, regardless of the boundary conditions. It was indicated in [14.159] that though the
knock-down factor based on the standard deviation of the panel thickness was not always
conservative, it did indicate trends of behavior fairly well. It is observed in Figure 14.140
that the range and trend of the coefficient of variation in thickness strongly depend on the
ratio R/t. Since the decrease in buckling strength in the tests was also Rlt dependent, it may
have partially stemmed from the increase in thickness variation. The applicability of this
thickness-based knock-down factor is depicted in Figure 14.141, where the analytical knock-
down factors are compared with the experimental ones. It is apparent from this figure that
considering the thickness coefficient variation as the sole measure of imperfection was in-
complete. Based on the results of the imperfection sensitivity study, it was concluded that
composite shell elements subjected to compressive loading suffer from imperfection sensi-
tivity to the same extent as do metal shells.
Experimental studies on buckling behavior of composite panels, supplemented by numer-
ical studies, were conducted in the eighties at the Wright Aeronautical Laboratory, Wright
Patterson AFB (see Bauld and Khot [14.162], Becker, Palazotto and Khot [14.163] and Khot
and Bauld [14.164]). Some of these investigations performed by Bauld and Khot [14.162]
and [14.164] complement one another. The objective of the study in [14.162] was to deter-
mine experimentally the stability characteristics of fiber-reinforced graphite-epoxy identical
eight-ply (0/90b laminated circular cylindrical panels subjected to uniform axial end dis-
placements, to compare the test results with analytical predictions obtained from a computer
program CLAPP [14.165] and thus to assess the capability of CLAPP to predict the stability
characteristics of composite laminated panels for various boundary conditions.
Four panels, 16 in. (406 mm) long, 8 in. (203 mm) wide (plane-form), 0.038 in. (0.965
mm) overall average thickness and 12 in. (305 mm) internal radius of curvature, were tested.
The loading fixture used in the experiments was a modified version of the test rig depicted
in Figure 14.137 and used by Wilkins [14.159]. The test fixture provided clamped end
conditions along the curved edges of the panels, which were achieved by employing the
loading head and base plates with auxiliary pressure blocks and clamping blades depicted
in Figure 14.142. The pressure blocks that provided the clamping, together with the clamping
plates that restrained the pressure blocks from altering their position, are shown in this figure.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

LAMINA ITS
o I0111:llc 0 lt]()lc
o It451c 0 10145/lll/·451,
o H51c o IOittillc
c. l+]()lc 0 IO/t45101 5
o IO/t451 5

v
to. IOic
1.0
t:.. (
t:.. (TO
2u .8
1~
0
::z
'i9
"'g .6 ()
8:
0

If.
u<D
"'
u
0
0 0
z 00
"'
~
0
~ .4

v~
z
~

"'
v
0.
~ .2

0
0 .2 .4 .6 .8 1.0
ANALYTICAL KNOCKDOWN FACTOR

Figure 14.141 Correlation of experimental knock-down factors obtained in Wilkins's compression tests
with the analytical thickness-based knock-down factors
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1174 Composite Structures

Figure 14.142 Head and ba.o;e plates wilh auxiliat) pn:§sure blocks and clamping plates employed for
clamping !he curved edges of the compres"on panels of Bauld and Khot (from (14. 162))

Axial di;placemenL~ were prevented at the upper curved edge by the upper platen of the
loading rig. while uniform prescribed axial displacements were applied a t the lower curved
edge. 1\vo distinctly different simply supported boundary conditio ns along the straight edges
o f the pa nels were provided by the test apparatus: zero circumferential displaceme nt or 1cro
normal force. The common bo undary condi tio ns alo ng the unloaded strai ght edges were zero
transverse displacement and zero mome nt. The tests were performed on two specimens with
unsupported un loaded side edges and o n two specime ns with si mply supported side edges.
The panels were tested in a 120.000 lb (533.4 kN) Tinus-Oisen hydraulic testing machine
and loaded through the surfaces of the platen and the cross-head of the testing machine
(Figure 14. 143). Tilting of the cross-head of the te>ting machine wa.s prevented by u~ing
two large aluminum nuts positioned o n the vertical ~rews of the loading machine beneath
the cross-head (Figure 14.143). These nul!> were tightened against the lower surface of the
cross-head. locking it in place.
Bauld and Khot's study (14.162) included a ,cnsitivity evaluation of initial geometric
imperfections of the specimens. To accomplish this task. the device shown in Figure 14. 144
was used to measure the deviations of the panel surface from a perfect cyl indrical form.
S ince proper lo;ading and shape measurements of a single c urved panel rely stro ngly on the
parallelism of the loading pla tes and o n the grooves in these pla tes being concent ric so that
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

t' lgurc 14.143 Bnuld and Kl•ot's compression tc>ting fixture of curved panels (from [14.162))
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1175

Figure 14.144 Geometric imperfec tio n measuring device anached to the baseplate of the testing fix ture
used by Bau ld and Khot (from [ 14. 162])

they both lie in the same imagi nary cyli ndrical surface, the device of Figure 14. 144 was
used as well to align the head- and haseplates to meet these requirements.
Bauld and Khot ( 14.1 62] described in detai l the carefu l appl ication of the procedures
pract iced by them for selli ng up the test apparatus, installation of the test panel in the loading
rig and measuremem of the surface imperfections. Since these represent a good example of
consistenL procedures for proper conduct of tesL~ on curved panels they are presented next.
The platfo rm of the imperfection measuring device was positioned so that its circu lar
groove was concentr ic with the circular groove in the load ing basepl ate (see Figure 14.144).
This was obtained by the dial indicator mechanism (shown in Figure 14. 144). which fitted
snugly in the circu lar groove of the platform and cou ld be smooth ly rotated along the groove.
T he d ial gage mechanism moved alo ng the platfotm, with its tip resting against the external
vertical wall of the circu lar g roove in the load ing baseplate, and the positio n of the imper-
fection device was adjusted until no change in the reference of the dial indicator was ob-
served for a complete scan along the platform groove. The measuring device was secured
by bolts to the baseplate of the test rig. To avoid accidental changes in the position of the
measuring device. position pi ns were installed. The above installation procedure yielded a
variation of less tJ1an 0.00 I in. (0.025 mm) for a complete transverse scan along the platform
groove.
To allow imperfection measurements at various he ig hts of the panel, the side supports of
the imperfect io n measuring device contained slo ts that permitted the positioni ng of the plat-
form. Positioning pinholes I in. (25.4 mm) apart in the side suppo rts ensured the posi tion
of the platform groove relative to the g roove in the loading baseplate for repeated measure-
ments at appropriate levels. Once the platform was located in a des ired position by the
positioning pins. it was secured in that positi on by cap screws.
The required parallelism of the base and head loading plates was achieved by a device
consisting of a rigid base that fitted snugly into the g roove of the baseplate and a dial
indicator auached to a stiff steel rod connected at its lower edge to the rigid base. This
device was moved along the groove in the baseplate while the plunger of the di al indicator
rested against the surface of the groove in the headplate. Once both plates did not deviate
from one ano ther by more than 0.002 in. (0.011 mm), the plates were clamped to the cross-
beam and to the platen of the Tin ius-Oisen testi ng machine by specially desig ned clamps
(sec Figure 14.143).
Alignment of the headplate relative to the baseplate, so that the corresponding sides of
their ci rcular grooves wou ld lie in the same circular cylindrical surface, was achieved by
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

posi tioni ng the movable platform of the imperfection measurement device at an appropriate
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1176 Composite Structures

level, setting the tip of the dial indicator extension against the vertical wall of the circular
groove in the headplate and adjusting the position of the hcadplate in accordance with the
change in the reference reading on the dial.
The procedure used to install a panel in the test fixture was as follows. The panel was
centered in the base plate and its sides were inserted in the slots of the vertical edge supports,
referred to in [14.162] as the "bookends" (see Figure 14.143), and the pressure blocks in
the baseplate and at each "bookend" were adjusted to a finger-tight position. The platen of
the test machine was raised to insert the upper edge of the panel into about three-quarters
of the depth of the headplate. The pressure blocks in the headplate were adjusted to a finger-
tight position and the machine platen was further raised until a compressive preload of 25
lb (111 N) was experienced by the panel. The pressure block in both plates were then
adjusted to their final position by a 40 in.-lb (4.5 kN-mm) torque, whereas the pressure
blocks in the "bookends" were only adjusted to the finger-tight position, to comply with the
requirements of simple supports.
Two types of measurements were recorded during the test of each panel: the end-
shortening as a function of the applied load level (i.e. the equilibrium path of the panel),
measured with a dial indicator that determined the relative displacement between the platen
and cross-head of the testing machine; and the axial strain distribution along an arc located
1 in. (25.4 mm) below the upper edge of the panel. Strain gages bonded face to face along
the centerline of the panel and at two other equally spaced locations on both sides of the
centerline were used to measure these axial strains.
The equilibrium paths obtained in the tests of [14.162] with the two panels, having un-
supported sides are shown in Figure l4.145a, while the other two specimens simply sup-
ported along their straight edges are depicted in Figure 14.145b. Figure 14.145a also includes
the equilibrium paths computed with CLAAP in the presence of an initial imperfection of
the form
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

W0 = Wl (1 + cos 0.393x) ( 14.40)


with WI = 0.0, 0.005 and 0.010 in. (0.0, 0.127 and 0.254 mm), respectively. Figure 14.145
includes the CLAAP [ 14.165] computed responses of a perfect panel with simply supported
straight edges, when either the circumferential displacements are not restrained or are pre-
vented. The measured initial imperfections in [14.162] were slightly larger than those char-
acterized by W1 = 0.005. Furthermore, Figure 14.145a indicates that the experimental buck-
ling loads experienced in the tests with the unsupported sides (1090 lb and 1065 lb, ~ 4.85
kN and ~ 4.74 kN) are in good agreement with those computed by CLAAP in the case of
Wl = 0.005 (1170 lb, 5.21 kN). Figure 14.145b reveals that the simply supported panels
failed under buckling loads in the range between the extreme buckling limits when circum-
ferential displacements were unrestrained and when they were entirely prevented. This good
agreement between tests and theory in [14.162] was attributed to circumferential slippage in
the "bookends" that was observed in the tests.
It should be noted that deviations of the panel surface from a perfect one at nodal points
of a 9 X 9 in? (228.6 X 228.6 mm 2 ) rectangular grid marked on the inner surface were
measured for each of the tested panels. The reference point for these measurements was
located on the axial centerline of the panel. The recorded imperfection data were, however,
only relative to the reference point, since this point might not necessarily lie on a surface
of a perfect circular cylindrical panel. These measurements, however, provided a qualitative
assessment of the surface deviation from 1~he perfect cylindrical geometry. The imperfection
data were not presented in [ 14.162], but they were used to establish a realistic imperfection
amplitude for an assumed analytical distribution of imperfections.
The Wright Patterson tests [ 14.162] revealed the acute sensitivity of the buckling loads of
the panels to the circumferential displacements imposed along the straight edges of the
panels. To assess this sensitivity further, Khot and Bauld conducted additional tests with five
(0/ ±45/YO)s panels simply supported along their straight edges and five (0/90) 2 s panels
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1177

12901b
1300
1170 lb
1200

1000

900
~
aoo

~ 700

.. ""'
""'
Panel 16 In 8

J
400 1 1n

F•ber pCJttem: (0/901 2,


Ed91 concl1hons: Clamptd curved WI• 0.010
edqu,
Uf'l~upported
stra1ql'lt

--o:ooz~~'''"
VI 1
o.006 o ooe
~~o.o1e o u2o
End shortero1~, In
(a)

4311 lb
Bud, ling limit wlk:n .::iTL"lnnli:rentml
Oive~enct at 001975 in.
di1<pilh.:.:n~;m:Pf1--.:ludc:d

,..,
3600 Anoly~u;OI eQuilibrium pO'ti\S
_/

"""' 3240 Itt buckled


"""
lOOO

'"""
21100
- 2~0 lb
~
"""'
,;mo
~
~ """'

~
I~

1<00

I"""
"""
100 P~;~nel16 ,, 1 e '"
F'ibu pattern: C0/'90) 1 ,
000
Ed9t condUIO'It: Clamped curved ~u
Somply-k!pporttd strooqht
tdQtl

0 60 S I 0 I 2 1... 1 6 I 8 2 0 2 l 2.-4 2.6 2 II 30


End- shorttnu•.q 11 10- 11 I 1n 1
(b)

Figure 14.145 Equilibrium paths of the compression panels tested by Bauld and Khot (from [ 14.162]):
(a) sides of panels unsupported, (b) side of panels simply supported

unsupported along their sides (see [14.164]). The material used for fabrication of the addi-
tional test specimens was identical with that used in their earlier tests. The new specimens
of [14.162] had plane form dimensions of 16 X 12 in." (406 X 305 mm2 ) and a radius of
curvature of 12 in. (305 mm) and were 0.038 in. (0.97 mm) thick. Before the panels were
tested, their initial imperfections were recorded and used later in the numerical computations
carried out with CLAAP.
A comparison of the test results of the simply supported (0/ ±45/90) 5 panels with nu-
merical predictions of CLAAP is presented in Figure 14.146a and b. The calculations pre-
sented in these figures correspond to two extremes: fully restrained circumferential displace-
ments and free ones along the sides of the panel. The computations were based on the initial
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
geometric imperfections measured on the panel, whose buckling load was 5775 lb (25.7 kN).
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
....
....
ex!

7000 . - - - - - - - - - - - - - - - - - - . . . . ,
.......-.. o.-. -RHtrw.I-
_,.
C - I o l ~~~.. Dlopl<lc-nlo 6000

1000 -~ IIOitb /
Ooto- ,,..
C-lol EO,._.,,. / ' 1711tb cfl
000 Exp.rii'M'ntal Dota ~ 0 1510 lb 0

D D 0 ....,_.t Ooto
•/
0
0
~
80"
0

:e
5000

410$~·
,I 0 °
IIIOtb :e
r / ....-- '\105 •
~ '0 00
oo ~
.....
g 4000 ,/o oo
iI.U
• 0 0
i... 0 oo .....

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
00 4(

~ 3000 0 ~

L'D ~
)( (JI 1-- 11 In --j .....
4(
;!
..... l2 2000
4( ·oooo

E- c-
~ 2000 D
ftblr Pattwn -(0f!.4!i/to)•
'IMr llattem""'(OI!'SJto)a
1000 ~ c-.oono-<:......0 Eo,oo Clompod.
_ . , . E.... Simjlty S..-'od
Total Ptalwl Thdcne.- 0.0)1 in.
flinle OWk,..,.. Grid - th 20
50ralllh•---
- - , ~<..-•
Total Jlawl Thc:knw•- 0.031 In
I'Wt• Dlftefwc.• Grtlf- f1 • 20

25 SO 75 100 125 ISO 175 200 225 250 275 300 325
0
o_L.-:so=--100~~1S0~72oo=--:250~-=3~oo::-:3~so=-4:-:!oo'='"""-"4so=-7so~o=--=s~so=--=600~
END SHORTEN~G In !x10- 4 J (b) TRANSVERSE DISI'LACEMENT PARAMETER.~•
(a)
Figure 14.146 Comparison of Khot and Bauld's compression panel test results with CLAAP predictions (from [14.164]): (a) end shortening versus load. (b) transverse
displacement parameter versus load

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1179

Figure 14.146 shows that as in the earlier tests of [14.162], the panels buckled here in the
range between the two extremes corresponding to the free and restrained circumferential
edge displacements along their straight edges. It was therefore concluded that the scatter in
the test results stemmed from the inability to control these displacements, in conjunction
with the sensitivity of the panels to initial geometric imperfection.
Similar results were experienced with the (0/90) 2s unsupported panels. There, however,
the different buckling values observed in the tests were attributed to the sensitivity to initial
shape imperfection and initial edge stresses arising during installation of the panel in the
test device. When the panels were removed from the mandrel after curving, they distorted
into noncylindrical surfaces. Upon their installation in the head- and baseplates of the loading
rig, their curved edges were forced to become circular. Consequently, their generators ran-
domly distorted, the distortion being largest along the side edges of the panels. It was as-
sumed that this distortion induced small unknown stresses along the edges.
Becker, Palazotto and Khot's investigation [ 14.163] concentrated on the fabrication process
and, like the other two studies [ 14.162] and [14.164], on test methods and analytical pro-
cedures for evaluation of the buckling of composite curved panels.
Six panel configurations were included in the test program of [14.164]: three were 12 X
12 in. 2 (305 X 305 mm 2 ) (chord length X height), simply supported along their unloaded
sides and of ( ± 45) 2s, (90/ ± 45 /0) 5 and (90/0) 2s eight-ply skin configuration; the other three
were 16 X 12 in. 2 ( 460 X 305 mm 2 ) with unsupported sides and of the same three-ply
orientations. Two samples of each configuration were tested.
The eight-ply laminated panels were fabricated from unidirectional Narmco T300/5208
graphite-epoxy tape hand laid into 17 X 36 in. 2 (423 X 914 mm 2 ) panels in a curved steel
mold.
To quote Becker et al. [14.163] the manufacturing process was as follows:
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

First, a nonporous separator material was laid into the steel mold. Then the eight plies of graphite/
epoxy tape were laid in followed by a TX 1040 porous separator. Two plies of Mochburg (one
ply for each four prepreg layers) were added as a bleeder for the resin, and then a layer of Mylar
(with small slits) was added to separate the bleeder material from a fiberglass vent cloth. The entire
lay-up was sealed in a nylon vacuum bag and the following cure cycle was used.
(I) Apply full vacuum pressure (12 psi minimum)
(2) Heal to 27SOF at 3°F per min
(3) Hold at 275°F for one hour
(4) Pressurize to 85 psi and vent the vacuum
(5) Heat to 350oF at 3°F per min
(6) Hold at 350°F for one hour
(7) Cool to 150°F in 45 min or more and remove pressure
(8) Cool to room temperature (usually overnight).
After curing, the panels were ultrasonically inspected to insure that no voids or delaminations
were present (note that none were found). The 17 X 36 in. 2 (431.8 X 914.4 mm 2 ) panels were
then cut into smaller test panels (12 X 12 and 16 X 12 in. 2 ), [304.8 X 304.8 and 406.4 X 304.8
mm'] using a specially jigged radial-arm saw with a diamond-tip blade. The sequence of cuts and
the cutting process itself were designed to yield panels with the opposite edges parallel. Two length,
two width, and five thickness measurements were made and recorded for each panel. This technique
of lay-up. curing and processing resulted in high-quality curved panels with minimal geometric
imperfections (such as warping, thickness variations and nonparallel edges). The maximum length
or width variation within any panel was 0.15 percent and the maximum thickness variation was
1.5 percent. The material properties were determined experimentally for a given ply[:]

£1 = 20.5 X 106 psi; G 12 = 0.75 X 106 psi


E 2 = 1.3 X 106 psi; v 12 = 0.335
The test fixture used in the program was similar
Copyright Wiley
Provided by IHS Markit under license with WILEY
to that in [14.159] and [14.162]. The
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1180 Composite Structures

boundary conditions along the unloaded edges were provided by fixture side supports (see
Figure 14.147). The straight edges of the test panel were inserted between the two inner
bars clamping the specimen with thin, knife-like edges that allowed the specimen to rotate
in the W,. direction (out-of-plane rotation). To obtain simply supported boundary conditions
(with the panel free to move in the in-plane u and v directions), the inner bars were advanced
by the tightening bolts until they just touched the specimen surface. To provide free-edge
boundary conditions, the side supports were removed. In installing the panels in the test rig,
a 20,000 lb (89 kN) Instron testing machine, the baseplate of the loading fixture was placed
on the load cell. The panels were clamped to the baseplate first, and then the side supports
were adjusted to their proper position. The headplate was attached to the upper edge of the
panel, making sure that it was aligned with the baseplate and that the headplate remained
horizontal to eliminate any bowing of the test panel. At this stage the loading ram was
lowered until contact was established with the test fixture. The loading ram and headplate
of the fixture were then clamped together to avoid tilting during the loading cycle. At this
point the instrumentation, consisting of back-to-back strain gages in two locations and four
LVDT's, was set to zero. The strain gages were bonded 2 in. (50.8 mm) below the top edge
and 3 in. (76.2 mm) to each side of the vertical centerline. Three LVDT's, used to measure
out-of-plane displacements, were spaced equidistantly along the vertical centerline. One
LVDT was placed between the head- and baseplates to record end shortening.
As the test began, load was introduced at a rate of 0.05 in. (1.27 mm) per minute. When
the first buckle was observed, the location was noted, and loading of the panel continued
into the postbuckling region. Loading was terminated when the headplate contacted the side
supports, when failure of the panel occurred or when sufficient postbuckling had taken place.
Equilibrium paths obtained in the tests of two panel configurations of [14.163] are depicted
in Figure 14.148. At buckling, in the case of the simply supported panel, a drop inN, (Figure
14.148a) was associated with the formation of buckles, whereas in the case of free-edge
panels, a drop in N, resulted from the vertical edge snapping into a new deformation pattern.
The STAGS C-1 [14.88] computer code was used here for comparison studies between
theory and experiments. The analysis with STAGS C-1 included bifurcation analysis for five
sets of boundary conditions along the unloaded edges, SSl, SS4, CCI, CC4 and "free," and
a nonlinear analysis. The bifurcation results were found to be 23-45 percent higher than the
test results, whereas the nonlinear collapse loads compared better and were only 12-23
percent higher than the experimental loads. Becker et al. [ 14.163] argued that the above
discrepancies were due to imperfection sensitivity. A comparison study at the Lockheed Palo
Alto Research Laboratory, using the STAGS C-1 code and introducing an imperfection in
terms of nondimensional initial lateral displacements ~ (where ~ is the ratio between the
lateral initial displacement and the panel skin thickness), yielded Figure 14.149. It is apparent
from this figure that the panels were indeed sensitive to imperfections. Hence, though im-
perfection studies were not performed in the program of [14.163], Lockheed's results indi-

,---~,':--,---.-,~--TIGHTENING BOLTS
,,
''

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 14.147 Fixture side support used in Becker et al.'s compression tests of cylindrical panels (from
Copyright Wiley
[14.163])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1181

so
T\10 12 x 12 PAIID.S
40
(90,0)25

/
,'",.... --""'\.. ... _,..,-' --- SIMPLE SUPPORT B.C.

I
/
, I

.01 ,02 .OJ ,04 .as .06

(a) End Shortening (inches)

TWO 16 x 12 P AJIELS
10 (!:45) 2s
FRtt B.C.

\
'-----

.01 .02 .OJ .04 .05 .06

(b) End Shortening (inches)

Figure 14.148 Equilibrium paths obtained in the compression tests of Becker eta!. (from [14.163]):
(a) two ( ± 45) 2 s panel with free boundary conditions along their sides, (b) two (90/
Oh,, panel with simple support boundary conditions along their sides

cated that the panels had relatively large imperfections. It was suggested in [14.163] that
these imperfections stemmed from the fixture and loading apparatus. The investigations there-
fore concluded that panel imperfections could be minimized with proper fabrication tech-
nique, whereas imperfections due to the test set up were difficult to eliminate.

Another experimental and analytical investigation, with emphasis on initial buckling and
postbuckling performance of a graphite-epoxy cylindrical panel in axial compression, was
carried out by Snell and Morely in the mid-eighties [14.166]. Their test program consisted
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

LOCKHEED STAGSC-1 ANALYSIS

~1M END 1~1111

Figure 14.149 Comparison of Becker et al.'s test results on a (90/0) 2 s simply supported panel along
the vertical sides with Lockheed STAGS C-1 imperfection sensitivity analysis (from
Copyright Wiley [14.163])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1182 Composite Structures

of nine panels with radius, length, arc length and thickness of 250 mm, 540 mm, 444 mm
and 2 mm, respectively. All panels were fabricated from 16 plies of XAS-914C prepreg in
varying combinations of 0°, ± 4SO and 90°, in compliance with aircraft loading in bending
and torsion, in order to determine the potential tradeoff between initial buckling stress and
postbuckling strength. Five of the panels were of the following surface lay-up each: ( ± 45 I
+45/04 ),; (±45 4 )s; (+45/0 4 /-45/0 2 ) 1 ; (+45/90/0J-45)c, and (±45/+45/0/+45/0/
-45)s; while four were of (+45/0/-45/0+45/0/-45/0) 1 configuration. Three panels of
the later configuration were manufactured with deliberate geometric imperfections, intro-
duced by interposing a graduated stack of peel-ply layers between the cylindrical mold
surface and the laminate stack. This resulted in a lateral bulge in the form of a pyramidal
hill extending over the whole panel. One of the panels had an imperfection magnitude of
nominally 0.5 mm outwards, the second one of 0.5 mm inwards and the third one of 1.0
mm outwards. The imperfection shapes were measured prior to testing the panels and were
then used in the theoretical calculations.
The loading fixture used in the tests of [ 14.166] was similar to that employed by Wilkins
[14.159], and thus the specimens were nominally clamped along their curved loaded edges
and simply supported along their unloaded straight boundaries.
The data acquired during the tests included overall load displacement curves, corrected
for test machine flexibility (obtained from an X-Y plotter connected to the testing machine
load cell and to displacement transducers); strain histories in all modes of deformation up
to failure (which were recorded by strain gages bonded face to face at nine internal quarter
points of the panel surfaces); geometric imperfection and buckling modes (which were mea-
sured by a displacement transducer mounted on a saddle of a motorized height gage clamped
to the test machine, and which combined with an X-Yplotter provided records of prebuckled
wave forms, imperfections and postbuckled patterns); and lateral displacement at a buckle
peak (measured by a deflection transducer and continuously recorded with load). Further-
more, postbuckled modes were photographed during the test for observation of modal de-
velopment and comparison with theoretical predictions.
The panels were loaded under displacement control by a 400 kN Schenck servo-hydraulic
test machine. Controlled end-displacement was used so that load changes through buckling
could be observed even when buckling was associated in a loss of load capacity.
Snell and Morely's test results [14.166] were compared with the theoretical results yielded
by finite-element (STAGS C-1, [14.88]), finite strip [14.167] and Rayleigh-Ritz [14.168]
analyses. The authors reported that the finite-element buckling predictions overestimated the
test results by about 20 percent. The other methods provided quick results of similar accu-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

racy. It was also reported there that the postbuckling modes variations were laminate lay-up
and imperfection magnitude dependent but the buckling load exhibited very little sensitivity
to mode shape, and that the imperfections of the type introduced in this study had barely
any effect on the buckling capacity of the panels. Furthermore, it was shown that though
the panels demonstrated stable postbuckling behavior, their ultimate failure occurred at loads
lower than their buckling loads (see Figure 14.150). This figure combines the results of two
tests, the first up to initial buckling and unloading and the second loading through buckling
to failure (note that in this figure the full lines represent stable branches, whereas the broken
lines are "jumps" in load from one stable branch to another, under controlled displacement).
The results shown in Figure 14.150 indicate that composite materials permit repeated buck-
ling tests with no loss in performance (an indication of low fatigue sensitivity).
Another experimental and analytical investigation on postbuckling behavior and failure
characteristics of axially loaded selected graphite-epoxy cylindrical panels was carried out
by Knight, Starnes and Waters at NASA Langley in the same period [14.169]. The experi-
mental results were compared with linear bifurcation buckling and nonlinear postbuckling
calculations, including measured initial geometric surface imperfections and large rotations,
as yielded by STAGS C-1 [ 14.88]. The effect of orthotropy on the panel response was
assessed and failure loads and the location of failure were determined by calculating the
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1183

'"' SEP-stobl• •quftlbrltM'Tl path

120

1 Failur~
100 I
Jrd SEP 1
t
z 811
.>!
-o
"
0
--'60

"'
1.0 1.0 30 •. o
Displacement. mm

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Figure 14.150 Equilibrium path of a highly curved panel subjected to axial compression obtained in
the tests of Snell and Morley (from [14.166]. Crown copyright is reproduced with the
permission of the Controller of Her Majesty's Stationery Office)

local stress distributions near regions with severe local bending gradients, in conjunction
with failure criteria and defining first-ply failure.
The test program consisted of 10 panels fabricated from commercially unidirectional Thor-
ne! 300 graphite-fiber tapes, preimpregnated with 450 K cure Narmco 5208 thermosetting
epoxy resin. To determine the influence of material orthotropy on the postbuckling response,
the tapes were layed up to form three panels of 16-ply (±45/0/90/ ±45/0/90) 5 quasi-
isotropic laminates, three panels of 16-ply ( ± 45/90•/ +45)s orthotropic laminates, and two
pairs of panels of 16-ply (±45/+45) 25 and 8-ply (±45/+45)s angle-ply laminates. All
specimens were cured in an autoclave, using the procedure recommended by the resin man-
ufacturer. The panels were ultrasonically inspected, and no defect was detected in any spec-
imen. To achieve uniform axial compressive loading of the panels, their curved edges were
machined flat and parallel. For application of the moire fringe technique to identify the
buckling patterns of the panels and monitor visually changes in the postbuckling radial
displacement of panel surface, the convex side of each panel was painted white. The panels
had a 35.56-cm square plane form with a 38.10 em nominal radius. Since the nominal
thickness of the laminate (0.13 mm nominal ply thickness) differed from the actual one, the
nominal lamina properties used in the analysis were adjusted by the ratio of nominal ply
thickness to measured average ply thickness.
Initial geometric imperfection measurements of the panels, later introduced in the nonlin-
ear STAGS C-1 analysis [14.88], were performed with a contacting direct- current differ-
ential-transformer-type displacement probe in a transversing fixture. Elevation measurements
were dynamically obtained and recorded as the probe moved along arches at 64-mm intervals
along the length of the panel. These measured data presented the rise of the panel surface
above a reference plane and were compared with an ideal rise based on a best-fit radius.
The difference between these two sets was referred to as the initial geometric imperfection.
The method of reduction of the imperfection measurements data and of obtaining the ana-
lytical approximation to the initial geometric imperfections w 0 (x,y) is described in [14.169].
The approximation was obtained from a half-sine Fourier expansion of the measured surface
imperfections and used in the nonlinear analysis as the largest 20 values of the mean im-
perfection amplitudes g,,, selected from the modal content of the measured initial imperfec-
tions. A magnified oblique view of the initial geometric imperfections (ten times the panel
thickness) for two of the quasi-isotropic panels is shown in Figure 14.151 (see also [ 14.169]).
These calculated radial imperfections, which were substantially different from one another,
were added to the perfect geometry of the panel for computation of the buckling load.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1184 Composite Structures

(a) Specimen Al ((w 0 )max = .23t).

(b) Specimen A2 ((w 0 )max = .14t).

Figure 14.151 Analytical approximations to the initial geometric imperfections of the two NASA
Langley quasi-isotropic graphite-epoxy compression cylindrical panels (from [14.169])

The test apparatus employed for testing the panels with measured initial imperfections of
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

[14.169] was similar to that used by Wilkins [14.159]; see also [14.162], [14.163] and
[14.166], thus imposing clamped boundary conditions along the curved edges of the panel
and simply supported conditions along its unloaded edges. Strain gages and direct-current
differential transformers were used, as before, to monitor strains and displacements at se-
lected points. The moire fringe technique was used as well to monitor overall field radial
displacements. Photographs of the moire fringe patterns were taken during the tests.
The panels were slowly loaded to failure under compressive end shortening. The 16-ply
quasi-isotropic specimens were first tested to serve as reference data in evaluating the effect
of orthotropy. The other seven panels (three 16-ply orthotropic, two 16-ply angle-ply and
two 8-ply angle-ply) were loaded to failure to determine the effect of orthotropy on their
postbuckling response and failure characteristics.
The NASA Langley test results of [14.169] were compared with predictions obtained from
linear bifurcation buckling and nonlinear postbuckling analyses (STAGS C-1, [14.88]). The
measured initial geometric imperfections, which were used to accurately model the initial
panel geometry, were incorporated into these analyses. Good correlation between the post-
buckling test results and the STAGSC-1 results up to buckling was reported. The linear
bifurcation buckling analysis yielded buckling loads that were considerably higher than the
experimental ones, as well as the ones obtained by the nonlinear analysis (see Figure 14.152).
Furthermore, it was shown in [ 14.169] that use of the measured imperfection data allowed
accurate prediction of the panel response until local failure occurred in the panels. Failure
initiated near regions with severe local bending gradients and large surface strains that could
lead to local failures. These local failures were associated with the brittle failure character-
istics of the graphite-epoxy material system. Use of failure criteria in conjunction with an-
alytical results from the nonlinear postbuckling solutions qualitatively predicted the load level
and local failure location in the panels that corresponded to the experimental results.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1185

,004

z
I
,003 ®If:®

I

P/EA '®
....---©
.J.u·
.·1 [

.002
,~\_®
(D\\_F!RST·PLY FAILURE

• (P~rllrr.o
0 (?~rloerf
,001 •• •• ·AMALYSIS
=TEST

Figure 14.152 A typical comparison between the NASA Langley


test results on a quasi-isotropic cylindrical com-
.002 .004 .C06 .008 pression panel and STAGS C-1 predictions (from
u/L [14. 169])

It was also shown by Knight et al. [14.169] that, except for the orthotropic specimens,
the panels had significant postbuckling load-carrying capacity, but at loads considerably
smaller than the buckling loads sustained by them. In the case of the angle-ply laminates,
this postbuckling load-carrying capacity was associated with relatively large axial displace-
ments of the panels as compared with those experienced at buckling.
Further experimental and analytical studies on the postbuckling behavior under compres-
sion of graphite-epoxy laminated curved panels were reported by Kobayashi, Shumihara and
Koyama in the mid-eighties (see [14.170]). Their test program consisted of eight panels, one
of each of the following lay-ups: (0/90/0) 5 ; (0/ ± 60)5 ; (60/0/ -60) 5 ; (0/90/ ± 45) 5 ; (0/90h,;
(0/ ±60) 25 , and two with a (±45)25 lay-up. The panels were 170 mm long (nominal), 136
mm arc length and had an internal radius of 500 mm. They were fabricated by a hot-press
curing process from unidirectional prepregged tapes Torayaca T300 in a #2500 epoxy resin.
The panels were tested under controlled end-shortening in an apparatus similar to that
used by Wilkins (see Figure 14.137 and [14.159]). However, their unloaded straight sides
were simply supported by 10-mm diameter high-strength steel rods instead of the commonly
used knife edge supports.
The analytical studies of Kobayashi et al. included the development of an approximate
one-term Galerkin-type solution for predicting the postbuckling response of laminated curved
panels and the application of a finite-element analysis based on a hybrid-type plate element
derived by the first two authors (see [14.170] for details).
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Fair agreement between the finite-element predictions and the test results was reported. It
was shown in this study that the maximum postbuckling load was dependent on the lay-up
of the panel. Also, contrary to the observations in [14.166] and [14.169], the maximum
(failure) postbuckling loads sustained by the panels were found to be considerably higher
than the loads experienced by them at buckling (see [14.170]).

Cylindrical shells- Laminated composite cylinders are fabricated by either filament winding
or prepreg lay-up. Each of these manufacturing processes results in a skin construction that
possesses inherent material characteristics as well as geometrical ones, will influence the
behavior and buckling sensitivity of the shell to a different degree. To emphasize these
differences, the following discussion on buckling tests of composite cylindrical shells distin-
guishes between the filament-wound and prepreg-laid cylinders, and each kind is discussed
separately.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1186 Composite Structures

Filament-wound cylinders- Because of the relative ease of fabricating closed fiber-


reinforced cylindrical surfaces by filament winding, which provides an automatic, cost-
effective manufacturing method, most advanced composite cylinders in present structural
applications are filament wound. Obviously, for these very same reasons, the majority of
experimental studies with cylinders were performed with filament-wound specimens. How-
ever, though an attractive technique, filament winding introduces complex fiber geometries
in regions where material is interwoven during winding. This may affect the correlation
between experiment and theory.
Furthermore, in conducting tests with filament-wound cylinders, experiments have dem-
onstrated that the load-carrying capacity of full-scale wound cylinders is significantly lower
than that of their scaled-down version, used to verify design adequacy in the laboratory. It
was argued by Jensen and Hipp [14.171], that cylinders manufactured by filament winding
contain natural fiber defects: "These defects arise when the fibre tows cross-over one another
during an off-axis winding, often referred to as a helical winding path. Helical winding can
result in two distinct tow patterns: distributed or classical. Using a computer-controlled fil-
ament- winding machine, process parameters can be chosen, such that the undulation loca-
tions are distributed evenly over the surface of the cylinder (distributed pattern), or the tow
undulation can form in distinct circumferential and helical regions (classical pattern). In
either case, weak interfaces exist between the interlaced and/ or non-interlaced layers." The
structural compressive response of a wound cylinder is directly affected by the number of
tow undulations within a helical layer and the percentage of the weak interfaces in the
laminate.
Tests and analytical investigations were carried out by Card and his co-workers at NASA
Langley Research Center in the sixties (see [14.152], [14.172] and [14.173]). NASA also
sponsored similar studies at other research centers, reported by Tsai, Feldman and Stang
[14.174], Jensen and Hipp [14.171] and Hahn eta!. [14.175].
The results of early preliminary investigations with filament-wound glass-epoxy cylinders
were reported in [14.172], and their extensions in [14.173]. Test cylinders 15 in. (381 mm)
in diameter, 15 in. (381 mm) long, the walls of which consisted of alternate layers of cir-
cumferential and ±a helical winding ( ± 25°, ± 45° and 67 .SO), were used in the earlier tests.
The cylinders were relatively thick, with a radius-to-thickness ratio of about 145. To prevent
end failure, the walls of the cylinders were built up near their edges with additional circum-
ferential winding. Two different epoxies were used in fabrication of the test cylinders, Shell
Epon 828 with curing agent D and Shell Epon 826 with curing agent CL.
Two nominally identical shells of each type of lay-up were tested in axial compression
between the platens of a hydraulic testing machine until buckling or failure occurred. The
experimental results were correlated with predictions obtained by the Stein and Mayers
sandwich theory [14.176]. This comparison is depicted in Figure 14.153, where the dashed
line corresponds to failure in shear under axial compression of tubes that have an equal
number of circumferential and helical windings. Card and Peterson [14.172] showed that
cylinders expected to buckle elastically under loads not much higher than this failure in shear
stress would fail under somewhat smaller stresses due to interaction between the two types
of failure. They explained that the epoxy shear stresses became substantial and the cylinders
deformed plastically once they reached loads somewhat smaller than the shear failure load.
Evidence of this interaction at failure was exhibited by the 45o and 67.SO helically wound
cylinders but was absent in the tests on the cylinder with 25o helical angles. The 45° and
67.5° cylinders failed in shear at buckling, whereas barely any damage was observed in 25°
cylinders after the load was removed, and they could be reloaded. Based on the observation
that compressive load induces shear stresses leading to new failure modes, it was concluded
in [14.172] that the correlation between the experimental results and the predicted ones was
satisfactory (see Figure 14.153).
Card presented test results on 51 multilayered glass-epoxy cylinders loaded in compression
at both room temperature and elevated temperatures (see [ 14.173]). Some of these cylinders
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1187

COMPARISCN BETWEa< CALCULATED AND MEASURED BUCK~ING


LOADS FOR GLASS- EPOXY CYLINDERS

16[
14

12 ~ ' ""- <SHEAR FAILURE

'-----------<Y'
0
0

EPOXY CYL. TUBE


828-D C 0'
826-CL a

~ w ~ ~ ~ w m ~ w
HELIX ANGL~. CEG

Figure 14.153 Comparison of Card and Peterson's experimental compression buckling loads of fila-
ment-wound cylinders with predictions based on Stein and Meyers' sandwich theory
(from [14. 172])

were already employed by Card in an earlier test program (reported in [ 14.177]) that included
eight small-diameter tubes as well as some large-diameter cylinders. The experimental results
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

[14.173] were compared with buckling predictions by a Donnell-type laminated orthotropic


cylinder theory and with strength predictions based on existing unisotropic yield criteria (see
references cited there).
The geometry of the test cylinders, the walls of which were composed of six alternating
helically and circumferentially wrapped layers consisting of glass filaments (8 JLm in di-
ameter) embedded in an epoxy material, is depicted in Figure 14.154a and b. The ends of
the cylinders were reinforced to prevent end failures. A layer of glass cloth, about 25 mm
wide and of twice the thickness of the cylinder wall, was added to the small diameter tubes,
whereas a layer 38 mm wide of equal thickness to that of the wall was used to reinforce the
edges of the large-diameter cylinders. Variables in the specimens were the helical warp angle
a (a = 2SO, 45° and 67.5°), the cylinder matrix material (EPON 828 with curing agent D,
EPON 826 with curing agent CL and Hercules Powder Co. formulation no. 25), and the
combination of cylinder length and inside radius (tubes 305 mm and 660 mm, respectively,
and cylinders 381 mm and 381 mm, respectively). The scatter in specimen wall thickness
of the individual specimens was ± 3 percent, and that in the glass volume fraction was also
± 3 percent.
Both the tubes and cylinders were loaded in compression between the platens of a testing
machine. To ensure uniform load introduction into the specimens, their ends were cut fiat
and parallel prior to testing, and they were carefully aligned. All the tubes were tested at
room temperature, while the tests on the cylinders were performed both at room temperature
and elevated temperatures.
The tests on the tubes attempted to correlate the test data with strengths based on the
anisotropic yield criteria proposed in [14.173]:

(14.40)

where (jr. and (jT are stresses parallel and normal to the fiber direction in the unidirectional
layer, Tu is the shear stress in the layer, X, Y are normal and transverse allowable strengths
of a unidirectional layer respectively and S is an allowable in-plane shear stress of a unidi-
rectional layer. Comparison of the test data with the strength predictions showed the strength
predictions for the 4SO wrapped tubes to be overly conservative. Furthermore, as the helical
wrap angle a increased beyond 45°, the strength of the tubes decreased markedly.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1188 Composite Structures

(a)

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
(b)

FIJ(urc 14.154 Fi lament-wound cylindcrlo u~ed in Card'~ comprc"ion tests ( from [ 14.1731): (a) cylinder
geometry. (b) photomicrograph of wall of Ol:uncnt-wound cylinder

In the tests with the unheated cylinders. it was observed Ihat thei r failure wa~ accompanied
by the appearance of two tiers of diamond-shaped buckJcs. uniform ly distributed around their
surface (see Figure 14.155). In the 67.5" wrap angle cyli nder.;, splitting of the cylinder wall
aero~~ the buckles crest wa~ ob~erved. In view of the differences in strength~ experienced

FIJ(urc 14.t55 Failure mode~ obsc1·vcd in Card's compre;sion te'l< of filament-wound cylinders (from
[14. 173])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1189

with the tubes and cylinders of this wrap angle, it was suggested that the appearance of these
cracks was a consequence of the postbuckling deformations rather than of prebuckling ma-
terial failures.
The test results of the unheated cylinders were compared with the predictions obtained
by the analysis derived in [14.173] (see Figure 14.156). Figure 14.156 suggests that the
buckling behavior of a filament-wound cylinder is comparable to that obtained with metallic
cylinders, whereas the observed differences can be largely attributed to the nonlinearity in
the stiffness of the cylinder matrix as the helical wrap angle increases.
Tests of cylinders at elevated temperatures were performed by heating them within a
circular quartz lamp radiator (see Figure 14.157). Two asbestos-cement plates were used to
insulate the radiator from the testing machine. The plates contained circular cutouts to enable
the cylinder edges to bear on the testing machine platens. Uniform heating of the cylinders
was achieved by spraying the exterior of the cylinders with black heat-resistant paint and
inserting circular insulating plates in the ends of the cylinders. Thermocouples were embed-
ded in the wall of each cylinder, at a depth of about half the wall thickness, to monitor and

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
record the induced temperature.
In the tests the cylinders were heated at a constant rate of O.l4°K/s up to the desired
temperature. Then the temperature was held constant and the compressive load was slowly
applied until the cylinder failed. A lamp-voltage controller, which compared a desired tem-
perature input with that measured in a single monitor thermocouple, was employed to reg-
ulate the temperature of the cylinder wall.
As in unheated cylinders, failure of the cylinders at low test temperatures was associated
with loud reports and a sudden drop in the applied load. Buckles could not be directly
observed because of the quartz lamp radiator. Upon unloading and cooling, examination of
the specimens did not reveal any damage. At higher test temperatures, however, failures
appeared to be more silent and less abrupt. Examination of these test specimens revealed
splits in the cylinder wall and, in isolated cases, local blistering. Based on these observations,

10

125

___ ..
15
100

8
Q
50
- - - Nonaxisymmertic buckling In • 0)
-------- Axisymmetric buckling In • 01
Material
I 25
2
3
Material l: v • 0.65: \ • 0.60: C • 0.20;
1
r/t · 1~3: 1/r • 2:tl: t • 0.0525 in. 10.1~ cml
OL-----~----~----~~----~-----=~--~0
0 15 30 45 60 75 90
a,deg

Figure 14.156 Comparison of Card's compression test results of unheated filament-wound cylinders
with buckling calculations (from [14.173])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1190 Composite Structures

(b) Overall Vt/IW,

f igure 14.157 NASA Langley setup lo r clcvatcd-tcrnperature compression testing of filamcm-wound


cyli nders (from [1 4. 173])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

the mode of failure of a heated cylinder, buckling or material failure, was designated. The
test results of these NASA heated cylinders of 114.1731 are shown in Figure 14.158. It is
apparent from thi s figure that the heated fi lament-wound cylinders exhibited large reductions
in compressive strength as the average test temperature was increased (see also the discussion
o n thermal buckl ing in Section 19.3 of Chapter 19). It was suggested that at some temperature
the material in the cylinder wall was so degraded due to heating that buckli ng fai lures
coalesced wi th material failures. Hence, in excess of this material-cri tical temperature. the
compressive strength of the cyli nder became governed by its material strength at high tem-
peratures.
In another study Card [ 14.152) conducted an experimental program on buckling behavior
of filament-wotmd glass-epoxy cylinders (E-glass fi lament in ERL-2256 epoxy resin) aimed
at assessi ng a theory developed for the sensitivity of buckling (to initial geometric imper-
fections) under compressive loading of fiber-reinfo rced cylindrical shells. Following Koitcr
[2.36], [3.15) and [3 .16]. an index of imperfection sensitivity was obtai ned in this theory by
examining the character of the initial postbuckling region for a geomelrically perfect cylin-
drical shell. The theory also considered the effects of nonlinear prebuckling deformations
induced by edge constraints.
In the experimental investigation tests were conducted o n 12 cylinders, 30 in. (762 mm)
long with an internal radius of 30 in. (762 mm) and a nominal wall thickness of 0.072 in.
( 1.83 mm). Six of the cylinders had walls composed of five helically wrapped layers in the
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1191

r ••. °K

300 «X) 500


15
100
¢
cr--....
g. ___,~

8
'-,"'·~¢
'
75

10
'el:,~
~-
; '.~
0~ ~~~ 50 f
I
I
I
~
'I
\
''
\ I

I
--<>---
Material
I
\
~
\ 'ZS

---0--- 2 b ¢
-·<>--- 3

0
oo 100 200 300 «X) 500
r,.. °F
Figure 14.158 Comparison of compressive strengths observed in Card's compression tests of heated
filament-wound cylinders of various composite materials: (1) EPON 828; (2) EPON
826; (3) Hercules Powder Co. Formulation No. 25 (from [14.173])

sequence (±a!±a!± a!± a!± a). The walls of the other six cylinders were composed of
four helical layers and four circumferentially wrapped layers in the sequence ( ± a/901 ±a
1901 ± a/90/ ± a/90), with the last circumferential layer wrap (90) forming the external
surface of the cylinder. Following winding, the cylinders were cured at 250oF (121°C) for
two hours. The variables in the test specimens consisted of the helical wrap angle a(a =
15°, 30°, 4SO) and the cylinder wall configuration.
Sixteen strain gages, 6 in. (152 mm) long, were bonded back to back on each of the
cylinder surfaces. The pairs of gages were oriented about 90° apart in the axial and circum-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

ferential directions at the middle of the cylinder. The axial pairs were employed to measure
the initial strains, used to calculate the axial extensional modulus of the cylinder wall, as
well as to detect bending deformations, which might suggest buckling of the cylinder wall.
The circumferential gages were used to record strains, from which Poisson's ratio, /i,, as-
sociated with the axial strains could be determined. In addition, two linear differential trans-
formers were introduced to measure the relative motion of the platens of the testing machine.
To retain the circular shape of the specimen, a 0.375-in. (9.5-mm) plywood bulkhead was
inserted at each end of the cylinder. Then the cylinder was tested fiat-ended in compression
in the Langley Research Center 1200 kip (5.34 kN) testing machine. Uniform loading of the
cylinder was achieved by carefully aligning the head of the testing machine according to the
axial strain distribution experienced by the axial pairs of gages under small loads.
As in his earlier tests [ 14.173], Card observed also in his later tests [ 14.152] that failure
of the cylinders was associated with a loud report and the appearance of two "classical"
tiers of large, diamond-shaped buckles uniformly distributed around the cylinder surface. No
strain reversal prior to buckling was observed in the load-strain curves recorded during the
tests. Card pointed out that the measured buckling loads were found to depend on the wrap
angle of the test cylinders and varied from 50-90 percent of the predicted ones. This is
apparent from Figure 14.159 where the experimentally observed buckling loads are compared
with calculated ones by the theory derived in [14.152], as well as from Figure 19 of [14.152].
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1192 Composite Structures

1.0

.,,_,
.. -~
- r-
r-
'
r:::
·'
.. 1
• ,. s
'
.2
I"" 1S9 "' a>8
""' 2092lli

1S •s

Figure 14.159 Comparison of Card's compression tests results on filament-wound cylinders with al-
ternating helical and circumferential wrap with predicted buckling loads, (from
[14.152])

Another early experimental and analytical study on the compressive buckling strength of
eleven filament-wound cylinders was reported by Tsai et a!. [14.174]. The cylinders tested
in this study were manufactured from U.S. Polymeric Chemical Co. pre-impregnated fiber-
glass roving (20 and S994 roving with an HTS finish, impregnated with an epoxy resin
system EF 787). The cylinders had diameter to thickness ratios of 167 to 643, and their
walls consisted of three layers: a two-half-layer polar wrap at angles + cp and - cp with respect
to the longitudinal axis of the cylinders (0° -<:: cp -<:: 9°), enclosed by an inner and outer
circumferential wrap.
The buckling loads of the cylinders, assuming clamped boundary conditions, were ana-
lytically determined by a linear anisotropic shell theory developed by Cheng and Ho

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
[14.144]. Tsai and his co-workers reported in [14.174] that the cylinders carried 65-85
percent of the buckling loads predicted by the classical linear analysis. Furthermore, the
buckling mode of the small-diameter specimens was coupled with catastrophic shear failure.
The larger-diameter cylinders exhibited the classical diamond-shaped pattern, maintaining,
however, the basic integrity of the cylinder.
More recently (see [ 14.171 ]), Jensen and Hipp examined the influence of internal material
defects, determined by the winding sequence and undulation density and location and by a
number of weak interfaces in the laminate, on the response of filament-wound cylinders
under axial compression. The undulation location within a layer is a function of the winding
angle, while the percentage and distribution of undulations are dependent on the winding
pattern.
Jensen and Hipp [14.171] also presented details on the manufacturing of the six cylinders
included in the investigation. These cylinders of various winding sequences were manufac-
tured to ascertain the influence of crossover patterns on the buckling behavior: (90,/ ± 30,)s
(where n = 1 or 3), (0/ ± 60) 5 and ( ± 30/90) 5 . The crossover pattern depended on the
different helical angles. Switching the 90° and ± 30° layers allowed the effect of stacking
sequence to be explored. Tripling the cylinder thickness (90) ± 301) 5 increased the number
of the weak interfaces and enabled an investigation of the scaling parameter. In addition to
the classical winding pattern, a distributed pattern was employed to manufacture one of the
(90/ ± 30) 5 cylinders. Four different crossover patterns and undulation densities were pro-
vided by variation in winding sequence and helical pattern, (90/ ± 30) 5 with classical helical
layers, ( ± 30/90)5 with a classical pattern and (90,/ ± 303 )s and (0/ ± 60)s with classical
winding patterns.
The cylinders were 230 mm long and had an inner diameter of 150 mm and an average
thickness of 1.3 mm (6-ply cylinders) or 5.0 mm (18-ply cylinders). To ensure perpendicu-
larity to the loading axis, a special machining procedure was devised (for details see
[14.171]). The first test cylinder was extensively instrumented with strain gages at various
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1193

key locations. Data obtained from this cylinder indicated that one pair of back-to-back gages
would he adequate in the remaining tests.
Depending on the cylinder thickness, the cylinders were tested in a 267 or 1334 kN Tinius-
Olsen test machine. The load was introduced through special end fixtures, which were de-
signed to minimize rotation of the cylinder ends. One of the plates was machined with a
concentric recessed area to position a hemispherical alignment device that helped ensure that
the cylinder was subjected to pure axial loading. Clamped ends were obtained by pouring
molten Cerrobend into grooves in the end plates.
The cylinders failed instantaneously in the gage section, with little prior warning, though
some audible cracking was emitted at loads greater than 75 percent of the ultimate. Inspection
of the cylinders after their failure revealed neither brooming nor crushing of their ends. The
cylinders failed in a mode similar to that reported in the tests of Tsai et al. [14.174]. Failure
was initiated by local buckling near the circumferential crossovers. External hoop layers
delayed the onset of buckling and minimized external damage, but the failure mode was not
significantly altered. It was also demonstrated that the stiffness of the cylinder was reduced
with an increase in the number of weak interfaces. When the test results were compared
with finite element predictions, it was found that the experimental failure stresses were 58-
75 percent of the predicted ones.
Hahn et al. [14.175] summarized the results of a very detailed and systematic experimental
and analytical investigation, the goal of which was to improve the compressive strength of
thin filament-wound cylinders. To achieve this objective and to establish guidelines for the
development of design criteria for filament-wound cylinders, the study attempted a deter-
mination of the relationships between the axial compressive performance of the test cylinders
and their manufacturing procedures and quality. Therefore, considerable attention was given
to the manufacturing procedures of the filament-wound cylinders. Manufacturing problems,
as reflected by imperfections that develop due to improper fabrication procedures and affect
the quality of the shells and subsequently their buckling capacity, were addressed during the
fabrication phase. The imperfections included voids, regions of microbuckled fibers, thick-
ness variations, wrinkles in individual layers, gaps or excessive overlapping of individual
tows during winding, reduction in circumferential tensile strength in the layer, and residual
stresses (which develop via mechanical stresses, thermal stresses, or stresses introduced by
chemical processes). As indicated, however, in [ 14.175] these imperfections are designated
as extrinsic imperfections, meaning that they could be eliminated by changing the manufac-
turing parameters.
The first stage of Hahn et al.'s investigation [14.175] consisted of the manufacturing phase,
which was divided into two parts, and its evaluation. First, various windings and curing
procedures were examined to determine how to vary them in order to yield the combination
that would result in cylinders with the best microstructural quality and most uniform surfaces.
It was found that winding the prepreg tow at an elevated temperature and then using high-
temperature shrink tape apparently provided the best technique.
The actual fabrication of the test cylinders was considered in the second part of the
manufacturing evaluation study. The cylinders were fabricated from Thorne! T650-35 fibers
and Amoco ERL 1908 thermosetting epoxy combined into 1200 filament prepregged tows.
The tows contained approximately 32 percent resin by weight and were approximately 8.23
mm wide. Though the resin was B-staged, the surface tack of this system was substantial.
The cylinders were wound in sequences of: ( ± 30); ( ± 30)" 1 ; ( ± 30)s; ( ± 30/90),s and (90/
± 30)as·
Hahn et a!. identified the following questions as being critical to the design and manu-
facturing of compressively loaded filament-wound cylinders:

How does the crossover pattern-i.e., the number and location of circumferential and helical-
crossover bands-affect the compressive structural response of a single layer, helically-wound
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
cylinder?
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1194 Composite Structures

• How does the anti-symmetry of the laminated-shell region affect the compressive structural
response?
• How does the relative location of crossover bands within a multiple layered cylinder affect the
compressive structural response?
Can the results of a subscale experiment be used to predict the response of larger structures')
Can the quality and load carrying capability of a filament-wound cylinder be improved by
winding at an elevated temperature')
Can other winding parameters such as mandrel material, winding sequence, and thickness affect
the performance of filament-wound cylinders?

In an attempt to answer these questions, Hahn et al. concluded that the research should
be divided into seven experimental programs: in-plane crossover-band spacing, mandrel ma-
terial, cylinder scale, symmetry of the laminated-shell regions, winding sequence, through-
the-thickness crossover-band location and heated winding.
It should be noted that the determination of optimal manufacturing procedures, resulted
in redesign of the sensor-stand pulleys and carriage pulleys employed in filament winding
of composite shells. Also, due to the complexity of the heater apparatus used for elevated-
temperature winding, the cylinders were wound at room temperature first and then cured
with shrink tape.
Following the manufacturing phase, the quality of the cylinders was characterized. Since
Hahn et al.'s investigation [ 14.175] aimed at correlating the compressive performance with
the manufacturing parameters, the quality was characterized at two levels. The first level
included microstructural material characterization of the composite material, as quantitatively
reflected by fiber-volume fraction and void content, obtained from photomicrographs. It was
observed that the quality of the filament-wound cylinders was strongly affected by the man-
ufacturing procedures. The improved quality obtained from winding at an elevated temper-
ature was also evident from the material characterization.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Macroscopic appearance, thickness distribution and initial geometric imperfection, which
represent the deviation of the shell surface from a "perfect" cylinder, constituted the second
level of quality characterization. Thus, the geometry of the cylinders was evaluated locally
by examining the thickness of the shell surface at many points and globally by comparing
the measured shape of the shell with a "perfect" cylinder. The latter imperfections stemmed
from three factors: residual thermal stresses induced during manufacturing, inaccurate ma-
chining of the mandrel or cylinder and stresses introduced by attaching the loading end
fixtures.
Detailed information on the thickness measurements and the pertinent data reduction was
presented in Section 5.2.1 of [ 14.175]. Some roughness was apparent in the thickness dis-
tributions. It was caused by the porous release cloth used during processing, which created
a textured surface in the resin on the outer surface of the cylinders (Figure 14.160). Two

Figure 14.160 Textured surface in the resin on the outer surface observed in the filament-wound
cylinders of Hahn et a!. (from [ 14.175])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1195

special algorithms were used for improving the identification of poor quality regions and for
smoothing the thickness distributions.
The first approach, which was based on averaging the thickness at each point with its
nearest neighbors, is shown in Figure 14.161. Actual and smoothed thickness distributions
obtained by this approach are depicted in Figure 14.162. In the second approach, which used
truncated thickness distribution, the surface roughness became almost invisible, whereas the
dominating factors were easily identified. In applying this method, the thickness at each
point was examined by comparing it to the average thickness with regard to the standard
deviation of the distribution. When the thickness at any point was within two standard
deviations on either side of the mean thickness, the thickness was considered as the mean

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
thickness. For points outside this band, the actual thickness was used (see Figure 14.163 ).
In summarizing the results of the thickness measurements, it was concluded that the cylinder
thickness depended on the amount of twisted and misplaced tows, which caused local ridges
and valleys in the material. The thickness of multiple-layer cylinders was more uniform
when the helically wound plies were placed adjacent to one another since the fibers tended
to nest with one another.
Measurements of the global initial geometric imperfections, the instrumentation employed
to perform the measurements, their data acquisition and reduction were discussed in detail
in Section 5.2.2 of [14.175]. A linear least-squares analysis was performed to obtain the best
fit for the "perfect" hypothetical cylinder. Errors due to misalignment of the longitudinal
axis of the cylinder with the spin axis of the turntable on which the cylinder was located
were considered and removed in the process of reducing the data. The geometric imperfec-
tions observed in the filament-wound cylinders were found to differ dramatically from those
experienced with cylinders fabricated from unidirectional-prepreg sheets (as in [ 10.28] of
Chapter 10). Upon examination of the reduced shapes of the initial geometrical imperfec-
tions, which were measured on the cylinders tested by Hahn et al. [14.175], it was concluded
that on the global level the geometry of filament-wound cylinders depended upon the winding
pattern, scale and ply order.
Testing of the Hahn et al. cylinders engaged three groups of testing equipment:

1. Mechanical, which included a Tinius-Olsen 270 kN screw-driven universal testing ma-


chine, three LVDT's to monitor the end shortening of the cylinder, strain gages and a
data acquisition system for recording the response of the cylinder.
2. Acoustic, which involved a Locan-AT acoustic emission (AE) system manufactured by
the Physical Acoustic Corporation, to detect the occurrence of fracture in the composite
material and buckling of the shell. It should be noted that AE monitoring during buckling
was reported for the first time by Hahn et al. in [14.175]. A typical acoustic energy
observation obtained in these tests is shown in Figure 14.164.

0 Measured thickness locatlon

lsmoot~i-[11+12• ...+18+19] • Smoothed point

EDGE

~ !smooth=+ [11+12+13+14+15+16]

CORNER


~2
(smooth-} [11+12+13+14]

Figure 14.161 Hahn et al.'s algorithm for smoothing of filament-wound cylinders skin thickness dis-
tributions (from [14.175])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1196 Composite Struct ures

(a ) ( b)

Figure 14.162 Thicknes, field of a filamcnl· wound cylinder (from (1-1. 175)): (a) before and (b) aflcr
\lllOOihong "ith Hahn et al: , 3\eraging 1echmquc

3. Visual, wh ich included a 35 mm camera. a standard-~pecd VCR to document the entire


compressive test and a 2000 frames per second high->peed video system. Testing pro-
ced ures were discussed in detai l in r14 . 175].
The tests of Hahn ct al. complied wi th the seven experimental program; mentioned earlier
(i .e. In-Plane Crossover-Band Spacing. Mandrel Material. Cylinder Scale. Symmetry of the
Laminated Shell Regions. Winding Sequence. Through-the-Thickness Cro;,ovcr-Band Lo-
cation and Heated Winding). The te\1 re,uhs of these progmms indicated that the mandrel
material and symmetry of laminated-shell regions barely affected the compressive perform-
ance. (For further discussions o n the ef'f'e cts of these seven manufacturing parameters on the
compressive performance see [ 14. 175.])
Cla.ssical buckling analyses (Donnell and Fliigge type) and a strength analy>is (classical
lamination and Tsai -Wu failure criteria) were performed to supplement the e~perimcntal
investigations and improve the undef\landing of the effects of the above 'e'en manufactur-
ing-related parameters. together with cylinder geometry. constraints and geometric imperfec-
tions, on the compressive buckling behavior of filamen t-wound cylinders. FU11hcnnore, since
the inll uencc of each of the above parameters could not always be determined independently,
the analysi s developed provided a tool for isolating the effects of material and geometric
imperfections. Abo. comparison of the experimental re,ult~ and the analytie:ll predictjons
could be used to a.~scss the validity of the relatively simple analysis. A finite-clement analysi~
with ABAQUS software was perfonncd a~ well. to evaluate the compressh·c perfonnance of
the cylinders.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

In comparing the test results wi th the ana lyt ical predictions. it was found that the cy linders
fai led under loads between 30- 80 percent of the pred icted ones. However. the effects or

Fij:urc 14.163 Thicknc\s Held o r a filament-wound cylinder (of Figure 14. 1621 ;moothcd with Hahn
et al:, trunca1ion 1echniquc (from [14. 17511
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1197

- Lood/Pmcx
- - Oisp/dmox
10000
0.75

7500 ~

0.50 ~
c
5000 w

0.25
2500

0.00 ~-=----....__--' o.Jl..I.,.W~--"--4 0


a 5 10 15 20 25
Time (sec)

Figure 14.164 Typical acoustic energy observation obtained in Hahn et al.'s compression buckling tests
of filament-wound cylinders (from [14.175])

geometric imperfections, prebuckling deformations, and nonhomogeneous composite mate-


rials could not be accounted for in this analysis. Also, when the predicted buckling load was
much smaller than the predicted strength, the cylinder experienced stable postbuckling with
little damage. The amount of damage increased with decrease in the difference between the
predictions.

A good example of a carefully executed experimental investigation on buckling and post-


buckling behavior of large-sized filament-wound cylinders (600 mm diameter, radius-to-
thickness ratio R It = 90-180, length to radius ratio Ll R = 2.5 and a skin consisting of
three to nine layers) was reported by Puck and Ruegg in [ 14.178] and [ 14.179]. The building
of the test setup of their experiments at the Eidgeni:issische Material-Prufungs-Anstalt at
Diibendorf (EMPA, associated with ETH Zurich), the test program and in particular the
detailed treatment of the development of the buckling waves, though dating back over 25
years ago, are of special interest and worth studying even today.
The tests of Puck and Riiegg were carried out in the large (2000 Mp) EMPA displacement-
controlled hydraulic testing machine. Since the capacity of this large machine was more than
20 times the expected buckling loads of the composite cylindrical shells, the test rig was
very rigid indeed, as required for a study extending well into the postbuckling region.
The test setup (see Figure 14.165) was equipped with an internal support cylinder, a
mandrel (a in the figure) that provided a small air gap (c). This gap was 1 mm wide at the
start of loading and grew with increasing load due to the Poisson expansion of the test
cylinder to about 2-2.5 mm. Eight horizontal displacement transducers (e) were attached,
through special holes to the mandrel along one meridian, for measurement of radial deflec-
tions of the inner surface of the test shell. These measurements were recorded on an eight-
channel pen recorder. The mandrel could rotate about the central axis with the aid of a
friction wheel (h) driven by a pneumatic motor (g). Thus, eight 360° records of radial de-
flection were obtained at each load step.
The support mandrel therefore served for a dual purpose: ( l) it afforded a restriction to
the buckle depth and hence ensured a spreading of the buckle pattern over the whole test
cylinder (in the manner explored earlier by Horton at Stanford University for small metal
cylinders, see Subsection 9.2.1 of Chapter 9 and Figures 9.12 and 9.13); and (2) it provided
a rotating prop for the radial displacement transducers.
The measured deflection data included also a fairly complete record of the initial imper-
fections. The imperfection pattern was, however, not yet utilized for an improved prediction
of the buckling load, as employed at Caltech, Delft and Technion (see Chapter 10). The only
conclusion drawn by Ruegg from his initial imperfection data was that they were approxi-
mately of the same shape and magnitude for all his filament-wound shells. A more detailed
comparison with similar specimens of different materials and methods of fabrication, as
--`,`,`````,`,````,,``

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1198 Composite Structures

/
~~ r . <f +
~ c
,[s
8
d //-b
7
a I
6'
6'00!1
5
~
II~
e
J~ ~}=0
rz
r 'J.J'

[ltlllll h t~
=
.~
""1··'
!f
)
f!'.
r:
.... !,., ~.:
~~lfiil c
' ,11
m
n
' ·.:, '/
I

Figure 14.165 Puck and RUegg's setup for compression tests of filament-wound cylinders (from
[14.178])

would probably be canied out today, would have provided useful inputs to the International
Imperfection Data Bank (discussed in Chapter 10).

An early design analysis for buckling under axial compression of filament-wound cylinders
was presented by Ravenhall in [14.180]. The predictions yielded by this analysis were com-
pared with earlier experimental data obtained by Sargent at the Hercules Co. in Utah (see
[14.181]). A knock-down factor k, = 0.66 was determined from this conelation for use with
the theoretical predictions.
Two decades later, Garde [14.182] canied out experiments with filament-wound cylinders
loaded in axial compression. The cylinders were fabricated from E-glass ravings of 300 tex
in CIBA-GEIGY AY 103 epoxy resin with HY 951 hardener manually wound on an alu-
minum cylindrical mandrel with winding angles varying between oo and ± 30° (further details
can be found in [14.182]).
These experimental results were compared with critical buckling calculations employing
the following equation (presented by Baker et al., [14.183]):
N, = 2y/R[B 0 D,(l ~ /L,/L 0 )] 112 (14.41)
where y is a correction factor, R is the shell radius, fL, and fLo are the axial and circumferential
Poisson's ratios, respectively, B 11 is the circumferential extensional stiffness and D, is the
axial flexural stiffness.
The comparison between the theoretical predictions and the test results is depicted in
Figure 14.166. Good agreement between theory, incorporating the conection factor y, and
experiments is apparent from this figure, though the values of y used in the calculations
were based on experimental results for isotropic cylinders. Also, it can be seen from this
figure that cylinders with winding angles between ± 20o and ± 2SO exhibit the best com-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,

pressive performance.

Prep reg-laid cylinders- Extensive analytical and experimental studies on the buckling under
axial compression of cylinders fabricated from prepreg tapes were conducted at the Institute
for Aerospace Studies, University of TorontoLicensee=McDermott
Copyright Wiley
Provided by IHS Markit under license with WILEY
(UTIAS) in the seventies and eighties (see
Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1199

50~
z
~
4C i_
I
~ I
0
<
0
J ol
I
~
!

"'z 2 0"'. >


>

"'
u
::>
aJ
J I

( o ,o I 10 15 20 25 3C 35

Figure 14.166 Comparison of Garde and Lakkad's compression test results onE-glass filament-wound
cylinders with predictions incorporating the correction factor y (from [ 14. 182])

[14.160], [14.161], [14.184], [14.154], [14.6], [10.16] and [14.155]). Special emphasis was
given in these studies to the effect of initial geometrical imperfections of the cylinder.
Buckling of cylinders fabricated from perimpregnated unidirectional glass-epoxy "Scotch-
ply" (XP 250) tapes was experimentally and analytically studied by Tennyson and Mugger-
idge [14.184] while the implications of their results were discussed by Tennyson in [14.6].
The goal of their investigation [14.184] was to determine whether the buckling strength, and
subsequently the load reduction, of well-characterized composite shells containing random
distributions of shape imperfections (assuming them to be free of material imperfections
such as voids, thickness variations and curved fibers) could be adequately estimated based
on some measure of the magnitude of the imperfections present in the shell structure. (Here

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
this measure was chosen as the maximum value of the rms imperfection amplitude.)
The experimental program in [14.184] included tests with 14 three-ply wall shells: 1 ( ~ 70/
70/0), 1 (~20/20/90), 1 (901), 2 (90/~45/45), 1 (90/0/90), 1 (45/0/~45), 1 (~45/45/
90), 2 (0/ ~45/ 45), 2 (30/90/30) and 2 (30/90/ ~ 30). Prior to their buckling tests, the
material properties of each shell were characterized by subjecting it to internal pressure,
torsion and compression loadings. Strain gages were bonded in the axial, circumferential and
± 45° directions, at mid-length on the outer surface, and were employed to monitor the strain
response used to compute the orthotropic in-plane material constants. In order to determine
the filament content of the skin of the shells, coupons were cut from shells and resin burn-
off tests were performed. (A summary of the geometric and material properties of the shells
tested in this study can be found in Tables 1 and 2 of [14.184.])
The random distributions of shape imperfections of each shell were determined by mount-
ing the shell in a transversing apparatus equipped with two diametrically opposed, low-
pressure, linear contacting transducers. The digitized output of these transducers were fed
into a Fourier analysis program, and imperfection amplitudes and an estimated line of power
spectral density as a function of spatial frequency (w) was computed. The data recorded by
the transducers were also used to compute the average shell thickness (for more details see
[14.185]).
The cylinders were tested with their ends clamped. The clamped end constraint was pro-
vided by fitted aluminum end plates that were bonded to the shell wall. In an attempt to
permit repeated tests of the shells, an internal mandrel, having an air clearance of 0.030 in.
(0.76 mm), was installed inside the shell to minimize postbuckling deformation and subse-
quent damaging of the shell. In addition, this procedure provided the means for alignment
of the end platen of the testing machine relative to the cylinder, thus ensuring buckling under
uniform loading.
The test results of [14.185] were compared with the predictions by a nonlinear theory,
including
Copyright Wiley
an initial axisymmetric stress-free radial
Provided by IHS Markit under license with WILEY
displacement and nonmembrane prebuck-
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1200 Composite Structures

ling displacements, derived in thi s study. Since, however. it was assumed theoret ically that
the cylinders were free to rotate duri ng loading. a thrust hearing was inserted between the
top platen and the cylinder (see Figure 14.167). The introduction of th i ~ thrust bearing
increased by 25 percent the buckling capacity of the cyli nder with in-plane anisotropy over
that experienced by a torsionally re•traincd cylinder.
The analytical estimates of the buckling capacity of the imperfect shells were based on
the maximum value of the m1s imperfection amplitude. which was obtained from the statis-
tical representation of the measured initia l geometric imperfection. (Note that the magnitude
of the imperfection ampli tude was up to the order of the 'hell wall thickness.) It was found
from these pred ictions that the init ial shape imperfections signi ficantly reduced the com-
pressive buckli ng strength of the imperfect cylinder' relative to perfect ones. Funhermore,
it was reponed that the agreement between the predicted response of the imperfect cylinders
and the experiments was consistently good. and the discrepancy between theory and tests
did not exceed 20 percent. The difference was attributed in [ 14.184]to an inadequate survey
of the cylinder. so that the maximum nns imperfection amplitude was not necessarily ob-
tained. Pan of this discrepancy cou ld, however, also have stemmed from imperfection asym-
metry, edge constrai nts and material variation.
Hence, the theoretical method proposed in 114.1841. furnished a design method that
yielded accurate estimates of load reductions associated "ith random imperfections of small
mean square amplitude. The method provided conservati'e estimates of the compressive
buckling perforn1ance of circular laminated composite cylinders containing general type im-
perfections up to the order of the shell wall thickness.
A decade later, Tennyson and llansen investigated theoreticall y and experi mentally the
imperfection sensi tivi ty of laminated circular cylinders under axial compression (see [ I 0. 16 1).
The cyl inders in these tests were manuf;~ctured from 3M SP288 TIOO graphite-epoxy prepreg
tapes with the following laminate configurations: eight-ply (0/45/90/ - 45/0/45/90/ - 45),
(0/45/ -45/90),. (0, /-l5/ --l5/90, ). (0/-l5/--t5,145/0,). (021451-45,14510,> and (45/
45/4510.). and four-ply (0190,10) and (9010,190).
Prior to testi ng, the random distributions of the shape imperfections of each shell were
scanned by employing the Toronto transvcrsing apparatus (see [14.185]). These measure-
ments provided the nns imperfection value for several axial generators at various circum-
ferential posit ion~. and the maximum rms used to calculate the equivalent axisynunetric
imperfection amplitude. which wa' u~cd to determine the buckling load of the imperfect
shell.
Ba~ed on the comparison of the te't results with analytical predictions (sec [I 0.16)). it
was concluded that the buckling load of composite >hells could be significantly increased

Figure 14.167 University of Toronto lo3ding fixture with thru;t bearing allowing i'rcc rotation during
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

loading (from ( t4.6])


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1201

through a judicious choice of laminate configurations and their buckling performance could
be improved without paying a severe penalty of increased imperfection sensitivity.
More recently, Sun at UTIAS carried out an optimization study on buckling of laminated-
composite circular cylindrical shells subjected to axial compression, external pressure, torsion
or a combination thereof (see [14.155]). To verify this analysis, the calculated buckling loads,
corresponding to optimal configurations, were compared with experimental data obtained in
tests with eight four-layer cylinders fabricated from preimpregnated Hercules AS/3501-6
graphite-epoxy tape. (For details on manufacturing of the shell specimens see [14.155]). The
calculations of the buckling loads were based on solving the eigenproblem of the von K<ir-
man-Donnell thin-shell equations with clamped boundary conditions, including nonlinear
prebuckling deformations for cylinders containing an axisymmetric shape imperfection, the
amplitudes of which were obtained from imperfection measurements of the actual shells. To
obtain this measured initial geometric shape imperfection, the shell surface scanning method
and data reduction technique used in earlier tests [14.184] and [10.16] were also employed,
while the uniformity of the applied axial load was checked by axial strain gages.
Contrary to the usual assumption in optimum design of composites, it was shown in
[14.155] that the optimized laminates were neither mid-plane symmetric nor balanced. The
test cylinder for axial compression loading had a laminate configuration of (26/ -42/76/
- 3). Good agreement between theory and experiment was found for this optimized laminate
lay-up (see [14.155]). The test cylinder sustained 85.5 percent of the load predicted for the
imperfect shell. Koiter's postbuckling sensitivity parameter "b" computed for this shell was
negative, partially explaining the above discrepancy. Also, it was suggested that factors such
as load eccentricity, manufacturing defects and boundary conditions, other than the nominal
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

clamped ones used in the calculations, might have affected the test data.
Recently, further buckling tests on cylinders manufactured from CFRP tapes were con-
ducted at the Buckling Test Facility, Institute of Structural Mechanics, DLR Braunschweig,
Germany, by Geier et al. and Zimmermann (see [14.186] and [14.188]).
The test program of Geier et al. [14.186] included a set of 10 CFRP shells, the laminate
configurations of which were designed to carry maximum, minimum and intermediate buck-
ling loads. The purpose of the tests was doublefold, experimental confirmation of the com-
puted large range between optimum and pessimum designs (where the optimum may be as
high as 2. 8 times that of the worst one, for laminates consisting of the same number of
layers); and evaluation of the simplifications made in modeling the structure, to determine
the admissible ones for adequate buckling predictions.
The design procedure of the shells (250 mm in radius, 510 mm long, and layers oriented
in + e and - e in respect to the cylinder axis, which yielded optimum laminates ( ± 5119021
± 40) and ( ± 30/90 2 / ± 22/ ± 38/ ±53) and pessimum laminates ( ± 49/ ± 36/0 2 ) and (±51/
±45/ ± 37 I± 19/0 2 ) and a detailed description of their manufacturing are given in [14.186].
The cylinders were tested in the DLR Braunschweig Buckling Test Facility (see Figures
9.38-9.41 of Chapter 9). For more details on the test equipment, instrumentation of the
specimens and test procedures see [14.188].
The test results were compared with analytical predictions obtained by classical buckling
shallow shell theory for orthotropic shells simply supported along their edges, as well as
with analytical solutions for anisotropic shells with clamped boundary conditions. The com-
parison yielded "knock-down" factors (relating experimental results to predictions). The
knock-down factors were found to be relatively high, ranging from 0.80-1.03, their scatter
was small and they were not significantly smaller for the optimal cylinders (about 0.8) than
for the pessimal ones (about 0.9) (see [14.188]).
Since the knock-down factors were considered to depend upon the stiffness properties
used in the computations, the strain gage measurements recorded during the tests were used
to find improved stiffness values. It should be noted that these measurements provide certain
global laminate stiffness properties, whereas in the computations the stiffnesses of the uni-
directional laminae are the basic input properties.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1202 Composite Structures

The DLR test resuiL~ of [14.186]. characterized by high and consistent knock-down factors.
indicated that small strain moderate rotation theory provided a good basis for the prediction
of buckling loads of laminated composite cylinders subjected to axial compre~sion. They
also confi rmed the predicted large range between the opt imum and pess imum designs. A
factor higher than 2.75 was ohserved in the experiments. It was, therefore. concluded that
·'imperfection effects did not overwhelm the innuence wh ich the individual nominal design
had on the buckling load."
Geier and his co-worker; also reported [ 14.187) and )14.188) the result> of analytically
supported experimental studic., on buckling under non -uniform axial load of CFRP cylin-
drical shells. The studies focu~ed on the determination of the loading imperfection \Cnsitivity
(the buckling sensitivity to deviation of the load di~tribution from a uniform o ne). and a
search for imperfection-tolerant fiber composite cylinders. i.e. detemlining skin -laminate lay-
ups for which the predicted buckli ng load of the perfect shell wou ld become im.cnsitivc to
geometrical as well as loadi ng imperfectio ns. Note that the assumption that the cyl inder is
imperfection to lerant is of sign ificant importance in ;tructural modeling for optima l design,
where for simplification it is assumed that the shape a., well as the loading of the ~hell are
perfect.
To achieve the objectives of the test program. four CFRP cylinden.. 250 mm in radius.
510 rum long. fabricated from 0.125-mm prcprcg tape!. with Janlinatc Jay-up<, of (±-11/
± 14). (±75/±75). (±37/±52/±68/0/±60) and CO,/± 19/±37/± 45 /+5 1). were
tested in the DLR Braun>chweig Buckling Test Facility.
Since the imperfection sensitivity of the tested cy linder stems from the en·cct~ of in itial
shape geometrical imperfections and load imperfections, the in fluence of fiber orientations
o n each sensitivity type was studied separately. In order to determine the influence of geo-
metric imperfections. special care was taken to obtain homogeneous uniform loading (for
details sec [ 14.186)). The results of this phase of the experiments confmned that the fiber
orientation markedly influenced the geometric imperfection ~ensitivity. A knock-down factor
of 0.68 was experienced with the ( ± 751 ± 75) laminate.,, whereas a knock-down factor of
0.91 was found for the (01 /± 19/±37/±-15/±51) laminates.
In order to find out more about the influence of load imperfections and their role in
reducing the experimental buck ling load, the shells were loaded in a manner shown in Figure
14.168. Th in metal plates of different thi cknesses (0.05-0.40 mm) were inscned between
the upper layer of epoxy concrete used to obtai n homogenou~ uniform loading anti the top
lo:~ding plate. thus disturbing the uniform loading state. The location of the plates was altered
in turn at 32 equally spaced positions. and a buckling test was performed for c:tch position.
The buckling loads as a function of shape location (for two plate thicknes'e~) arc depicted
in Figure 14.169. This figure demonstrates large bucl.ling load scauer. It wa~ argued (see
[1-1.187] and [I-I. I 88)) that ~incc in the case of pure a'isymmctric loading the buckling loads
should differ only due to the thickness of the epoxy concrete plate, hut should be position

Epoocy concttt~
(07mm)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure t4.16l.l DLR Bmunschweig IC'I procedure ror non·


unirorm axial loading imperfection of compos-
ite C) lindel'l> (rrom [t4.1881)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1203

24 ~tpj=O.CSmm
i I I /\I I
kN \ r- J
v"
I

20 I ~"±Y
lA v \ '/\.. 1\
\ N.~. \l(\1/ :v
-g
16
r" I~~·01 mm y
0 I
l
~12
.s I
~ ~ I
"' 8 I

I i
I !

I I I i
4 12 16 20 24 28 32
Position No.

Figure 14.169 Disturbed buckling loads observed in Zimmermann's tests on nonunifonnly loaded
cylindrical shells as a function of disturbance location of Figure 14.168 (from [14.188])

independent, the observed scatter indicated that there existed at least another non-uniformity
that affected the buckling performance. To confirm this argument, further tests with inter-
changed upper and lower plates of the test shells were performed. The results of these tests
led to the conclusion that the non-uniformity of the curves in Figure 14.169 was due to the
presence of shape imperfections, which disturbed the axisymmetric condition of the cylinder,
in addition to the disturbance introduced by non-axisymmetric loading resulting from the
presence of thin plate inserts.
It was reported in [14.188] that the cylinder with the (0 2 1 ± 191 ± 37 I± 45 I± 51) skin also
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

exhibited the least load imperfection sensitivity. A knock-down factor of 0.89 was observed
for this cylinder when inserting the maximum thickness plate (0.40 mm), whereas with no
insert (uniform loading) a knock-down of 0.91 was experienced. The knock-down factors
experienced with the other cylinder laminate configurations were considerably lower (0.72
and below).
In light of the above results, Zimmermann concluded (see [14.188]), that small loading
imperfections could play an important role in buckling, that their effect strongly depended
on laminate lay-up and that a strong interaction exists between geometric shape imperfections
and load imperfections.
Extensive experimental, analytical and numerical investigations on the buckling behavior
of composite cylindrical shells were carried out by Giavotto, Poggi and their co-workers
[10.38], [14.189]-[14.191], [ 10.18] and [10.28] at the Departments of Aerospace Engineering
and Structural Engineering, Milan Polytechnic, Italy. The tests were performed in cooperation
with Agusta S.Pa Industries in Italy [14.189] and with the Department of Civil Engineering,
Imperial College, London, [10.38], [10.28] and [10.18]. These studies aimed at improving
the knowledge on the behavior of composite materials in shell structures, thus providing the
basis for better exploitation of the material properties and the accompanying controlling
influence of manufacturing processes on shell performance. Furthermore, the results were
intended to provide, through experiments supplemented by numerical analysis, a suitable
database for the development of Eurocodes for composite shells and thin-walled composite
elements subjected to combined loading.
Giavotto et al. identified the following intermediate goals necessary for achieving the
objectives of their investigations:
"a) Assessment of the effect of different manufacturing methods on initial geometric im-
perfections. The statistical properties of geometric imperfections on several series of
specimens made of composite materials with ditTerent lay-up configurations will be
evaluated and characteristic imperfection models for cylinders will be developed. This
will enable realistic tolerances on geometric imperfections to be established and in-
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1204 Composite Structures

corporated in the guidelines. At the same time, this study will produce valuable infor-
mation on the influence of manufacturing on product performance and will provide the
necessary input for analytical models to predict buckling strength of imperfect shells.
b) Experimental assessment of the buckling response of thin wall composite cylinders
under single or combined loading. A series of tests on cylindrical specimens will be
undertaken to estimate the buckling and post-buckling behaviour. All the possible in-
formation regarding displacements, stresses and geometric imperfections will be re-
corded for a further calibration of the numerical tools. The possibility to test a series
of nominally identical specimens will allow an assessment of the reliability of buckling
test predictions.
c) Development and validation of a finite element package and of analytical procedures.
After a proper calibration of the numerical tools it will be possible to generate addi-
tional results to extend the applicability of the guidelines beyond the geometric, ma-
terial and manufacturing range examined experimentally.
d) Development of strength design criteria for composite cylinders under axial compres-
sion, torsion and combined loading using both the experimental and numerical results.
e) Assessment of the suitability of the proposed guidelines in real design conditions and
comparison with existing design methods in terms of weight and strength of typical
applications."
To meet the goals of the investigations, tests with cylindrical shells made of orthogonal
Kevlar prepreg fabric embedded into an epoxy resin and laid upon a cylindrical mandrel
were performed. The cylinders were 700 mm in diameter and 700 mm long. They were
reinforced at their ends to facilitate their fixing into the loading rig (see Figure 14.170). The
stacking sequences of the test cylinders were four ply cross-ply (0/90) 1 and angle-ply
( ± 45) 5 , and eight-ply quasi-isotropic (45/ -45/0/90) 5 , (45/0/90/ -45) 1 and (0/90/ 45/
-45) 5 . The tests of [14.190] were performed with cylinders with stacking sequences of (0/
90) 5 and ( ± 45) 5 .
The cylinders were loaded in the loading rig of Figure 14.171. In the tests of [10.38],
[14.189] and [10.18], axial load in this rig was applied by a hydraulic ram, pushing the
loading platform against four hand-adjustable screw stops, which acted on the four corners
of the loading platform and controlled its displacement by means of four comparators (see
Figure 14.171a). This provided good-accuracy displacement-controlled loading. To measure
the uniformity and accuracy of loading during testing of the cylinder, the compression load
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

and axial displacement were measured at three equally spaced points. The load was recorded
by load cells placed under the lower clamp. The axial displacement was measured by LVDT's
placed on vertical bars fixed to the upper plate and reacted against the lower plate, thus
measuring the distance between the plates (specimen shortening). In the tests of [14.191]

71111PI'I

Figure 14.170 End reinforcements of Giavotto et al.'s test cylinders (from [10.38])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1205

h
I
G
l!.C--
. rt
. HYCfVotJ.. I C
...
....---
SCI!!'-' S lOP LOA.OINCi
i PLAtriJ!M

((llof>AIUI I~

I
Ul'f'Eit CI.AI1P
! j. ~
I I
LVOJ
i -
lRNf$~R
I
l O""CJt 0..11~ -+ -- i
LOAD t CLl

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
...:J uu L
(a )

(b)

Figure 14.171 Milan Polytechnic combined loading testing machine (courtesy of Professor Giavouo):
(a) former testing machine. (b) improved testing machine

the system was improved (see Figure 14.1 71b). T he screw stops were replaced by four ball
screws operated by stepping motors through reduction gears. and the loading displacement
was controlled by a computer, with feedback based on four LVDT measurements, each
coupled to a single screw.
Since uniformity of the applied axial load and min.imization of bending or circumferential
stresses due to clamping action strongly depend on the system used to clamp the specimen
to the loading rig, the Milan Polytechnic group devoted significant e fforts to the development
of the clamping devices used in their tests. In the tests reported in [I 0.38], [ 14.189) and
[I 0.18) the cylinders were at first fixed to the loading rig by pressing, from the inside, the
cylinder rei nforced ends against knurled steel rings by means of two semicircu lar brake
shoes. This system did not provide uni form clamping pressure along the circumference. It
was therefore improved by changing the mechanical c lamp into a hydraulic system shown
in Figure 14 .172a. In this system clamping pressure was provided by oil pushi ng a specially
designed rubber seal against the interior of the specimen. The pressure was now uniform,
but the appl ied ax ial force was eccentric (see Figure 14.172b). The system of Figure 14.172a
was then reversed and the eccentrici ty was eliminated. but under this loadi ng condition small
oil leakage, which could wet and pollu te the specimen , was likely to occur at low oil pressure.
To overcome these problems, the clamping method was again modified into a mechanical
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1206 Composite Structures

rAJCTION
rORCc

~
~
~
w -- -- li
~

':i

-- -
f :i
~
~
a ;;;

(a) (b)

~
E-5

(c )
Figure 14.172 Milan Polytechnic damping devices (courtesy of Professor Giavono): (a) hydrau lic
clamping system, (b) cccemric compression load inlroduced by hydraulic clamping, (c)
modified mechanical clamping device

system (see [ 14.1 9 1) and Figure 14.172c). Here. the ends of the cylinder were inserted into
24 external steel cylindrical sectors, wh ich were pressed against an internal knurled steel
ring (see sections A-A in Fig ure 14.172c). Uni form clamping of the cylinder ends was ob-
tained by tighteni ng the four screws in each section in a crossed way with three subsequent
moment levels, controlled by a dynamometer wrench.
In the tests of [ 10.38]. ) 14.189], [ I 0. I8] and [ 10.28], shape imperfections of the cylinders
were recorded by the scanning apparatus depicted in F ig ure 14.173. In this device two
computer-controlled LYDT's were used to scan the inner and outer surfaces of the cyli nders
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

with extreme flexibi lity on sampl ing pitch in both directions. T he acqu ired data were reduced
by the analyses presented in [10.38 ], ) 14.189 ) and ) 10.28], yielding the imperfection shapes

Figure 14.173 Milan Polytechnic apparatus for imperfection shape survey ( from [ 10. I 8])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1207

of the inne r and outer surfaces of the cylinders (see Figure 14. 174. which represents typical
surfaces of cross-ply a nd angle-ply composite cyl inders). Note that in the outer surfaces
depicted in this fig ure, the thickness variation d ue to the overl apping of the layers. associated
with the stacking sequence, is appare nt. As shown in [ 10.38), Fourier analysis of the inner
surface represents the real geometric shape of the cylinde r.
The initial sha pe imperfectio ns o f recent Mi lan Polytechnic shells (sec [ 14. 19 1]), as we ll
as their growth due to increase in applied load. were recorded by the no ncontact measuring
appara tus depicted in F igure 14.175. This device consisted of a central cylindrical surface
connected to a circular plate tha t was fi xed to the center of the lower clamping plate. Four
prec ision bearings. on which two arms with no rela tive motion were connected. were
mounted o n th is cylinder. The arms supported a slide on which five laser displacement
sensors were placed. T he slide could rota te around the central cylinder and move vertically
through a worm screw. Displacement of the sl ide was provided by two ste pping motors fixed
o n the support. One mo tor was connected through a toothed belt to a pulley, connected wi th
the two a rms, a nd provided rotation o f the sensors around the cyli nder axis. The second
steppi ng motor was connected throug h a toothed belt to a do uble idle pu lley o n the central
support. Th is pulley moved another pulley fixed o n the worm screw, whi ch provided vertica l
displacement o f the slide.
The vertical position of the slide was de tected by an increme ntal encode r connected to
the worm screw. Stru1ing o f every da ta acquisition was regulated by a photoelectric cell ,
which ensu red that da ta acquisition started always from the same point.
The five laser displacement sensors were placed 40 mm away from the internal surface
of the shel l. T hey had a measureme nt range of :!: 10 mm. with a resolution depende nt on
the acqui sition speed ( 15 ttm for 2 ms acquisition). T his allowed recording of both geome tr ic
imperfections on the orde r of tens of micrometers and buc king panerns associated with I0-
20-nu,, displacements.
Data were acqui red alo ng ci rc umfe rences by shifting the sensors from one circumference
to another thro ug h a helicoidal movement. The measurements were take n at equally spaced
poims along the circumferences, on equall y spaced ci rcumfe re nces. Sampli ng pitches in both
directions could be made with extre me flexibility. Usually 220 measures were taken alo ng

c,-,,~_, - 1'1•· Cyl•n•'"'.


t; a t~ l' llll l Surffi(:C
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 14.174 Typical imperfect ion ~hapcs of the intemal and external surfaces of the composite
cylinders tested by Giavouo e l al. (from [10.281)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1208 Composite Structures

LASER
DISPLACEMENT
SENSORS

Figure 14.175 Politechnico di Milano noncontact imperfection measuring device employing laser dis-
placement sensors (from [14.191], courtesy of Giavotto eta!.)

any circumference and 46 measures along any meridian. The internal surface of the cylinder
was recorded about 15-25 times during a buckling test. The time required for complete
scanning of the surface was about four minutes.
The top laser sensor was used to center the two clamps of the cylinder with respect to
one another. The clamps were lined up with the precision of the laser sensors (about 15
p,m).
A typical buckling pattern of the internal surface, measured during a loading test with the
apparatus of Figure 14.175, is shown in Figure 14.176.
The test results of [I 0.38], [ 14.189] and [ 10.18] were compared with theoretical predic-
tions. Knock-down factors ranging from 0.73-0.77 were obtained for the cross- ply cylinders
of [10.38], whereas higher knock-down factors, 0.88 for the cross-ply cylinders and 0.86 for
the angle-ply ones, were observed in the tests of [14.189] and [10.18].

c. Bending Experiments
In numerous applications, thin structural elements buckle by bending. These include large-
sized transportation containers, transport aircraft fuselages, launch vehicles and others. How-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`--

ever, investigations dealing with the buckling behavior of such cylindrical structures, partic-
ularly made of composite materials, are very scarce, and experimental studies are almost
nonexistent. A detailed discussion and literature survey on the behavior in bending, prebuck-
ling, buckling and postbuckling of cylinders, metallic and anisotropic, appeared in the recent
Copyright Wiley
NASA
Provided by IHS Langley-VPI
Markit under license with WILEY sponsored study of lFuchs, Hyer andIncStarnes
Licensee=McDermott [14.192].
- India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1209

•••
...
3 ••

· t 00

....
....
·7 00

·i 00

· 11 00

Fi~ure 1-1.176 A t)pical buckling panem of the internal ; urfacc of a carbon fabric cylinder obscl"\ed
in th<.' comprcs"on tests of Bisagni (counc'y of Dr. Ri'>agni)

To pmvide accurate predictions for the behavior of thin -walled composite cylinders suh-
jectcd to bending. Fuchs. Hyer and Starnes undertook un in-depth numerical and experi-
mental investigation of the bending behavior of graphite-epoxy shells. They assumed that
thi; would provide a better understanding of the influence of laminate stiffness and stnactural
parameters on the bending respon~e characteristics of laminated cylinders. Sub~cquenlly
the;e features would be u~ to gain better insight into the behavior of composite cylinders
under the more complex combined loading states. where bending constitutes one of the load
components. Their test program (~ee [ 14.192)) included ~ix eight-ply graphite-epoxy cylin-
ders with three lay-ups: qua~i-isotropic ( +45;0:90), and two onhotropie ( + 45/0~) 1 and (+ 45/
90!),. two length-to-radius ratios, 2 and 5, and a rad ius-to-thickness ratio of about 160. The
procedures of specimen fabri cation from prepreg tapes of 12-in. (305-mm) wide llercu les
A$4/3502 graphi te-epoxy, the fiberglass tabbing of the cyli nder loaded ends. the C-seanning
of the specimen to detect internal defects, and potting or the specimen ends into the loading
ring~. a~ well as the definition of the cylinder propenie-.. wall thickness. mechanical prop-
enies. and initial surface shape mapping were described in detail in [ 14.192]. The initial
geometric ~hape imperfections were introduced into the analytical predictions for comparison
of experiments with theory.
The bending test program was conducted in the Aircraft Structures Laboratory of NASA
Langley. The cyli nders were tested in a bending test fixtu re specifically constructed to meet
the goals of this study. As stated by Fuchs et al., "The load introduction concept posed a
chall engi ng test fixture design prohlem. The question arose on how to introduce a bendi ng
moment into the speci men wh ile precisely locating and gripping the specimen end~." This
was achieved by the method shown in Figure 14.177. The loaded ends of the cyl inder were
•et between two annular rings. an outer loading ring and an inner loading ring. u~ing low-
melting Cerrobend material. To react tensile loads. the loading end tabs bonded to the cyl-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

inder end were stepped (sec detail A in Figure 14.177b). Al~o. the grooves in the loading
rings prevented the cylinder wall from being pulled out in tension by providing a surface
for the potting compound to react against. ln applying a hcndi ng moment to the cyl inder hy
the approach of Figure 14.177. the potting compound transferred the tensile forces from the
inner and outer loading rings to the fiberglass tabs and the cylinder wall . Compressive loads
were tran~fcrred through direct contact wi th the test fixture. in addition to shear loading of
the cylinder wall. It is apparent that the method of Figure 14. 177 had several adv;tntages
Copyright Wiley direct mechanical fm,tening of the cylinder: gripping wa., relatively unifom1. no preci -
over
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1210 Composite Structures

_____ L__________ _ M

cylinder wall -

JoodJOW=IUI'e

T melnns pom1 alloy


poning compound
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

C1rcumferent1al
R grooves in
loading ring!L

J
~- ----------i~:r_l~~i~-ri~~--- - - - - - - - -
DETAIL A

Figure 14.177 Virginia Tech concept of introduction of bending load into composite cylinders (from
[14.192])

sion machining of the cylinder ends was required and removal of the failed specimen by
reheating the potting compound was facilitated.
Schematics of the bending fixture are depicted in Figure 14.178. The apparatus consisted
of a symmetric design in which the moment was applied at each end of the test specimen
by a hydraulic jack. The loading rings were bolted by 48 bolts on each end of the cylinder
to back plates (see detail A in Figure 14.177), with the cylinder ends aligned with the pivot
axis (see Figure 14.178a). The bolts were incrementally tightened in a star pattern to a final
torque of 80 ft-lb. In the fixture of Figure 14.178, load was transferred from the hydraulic
jacks through the loading pins and loading grips into the moment arm and back plate as-
sembly. A state of pure bending was induced in the cylinder by rotation of the back plate
assembly about the pivot pins (see Figure 14.l78b). It appears from the figure that application
of the moment introduced compression stresses in the top of the cylinder and tension in its
bottom. To avoid relative axial motion between the ends of the cylinder, braces were intro-
duced between the supports (pivot pins). Pressure was applied to the two hydraulic jacks by
a hydraulic pump and was monitored by a digital pressure transducer. A flow divider was
used to equalize the pressure in the jacks. The end rotations of the cylinder, D 1, and D2 (see
Figure 14.178b ), were measured independently by tracking the motion of the cantilevered
tabs with transducers D I through D4 (Figure 14.178b ). Strain gages mounted back to back
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1211

ptvot Jxt:>--......

n,
___.
D, '~,-jf<~=1 S
D2', 1

(b) Side View

Figure 14.178 Schematic of Virginia Tech bending fixture (from [14.192])

near the base of each support (see Figure 14.178b) were employed to monitor bending and
tension of the four supports.
Strain gages were bonded back to back over 60 locations on each cylinder surface, Direct
current differential transformers (DCDT's) were used to record end rotations (see Figure
14.178b) and radial displacements (see Figure 14.179). Load cells, acoustic emission and
the shadow moire method were also employed to monitor the behavior of the cylinder. An
average of 167 channels of data was acquired during each bending test (for details on in-
strumentation setup for each cylinder and DCDT measurement locations, see Tables 3-2 and
3-3 of [14.192], respectively).
In the bending tests the cylinders were loaded in a quasi-static manner until buckling
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

occurred. Then the load was held constant to observe the condition of the specimen and

Figure 14.179 Radial displacement measurements in Virginia Tech bending tests (from [14.192])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1212 Composite Structures

mark the observed deformation pattern. Undamaged specimens were unloaded, rebuckled
and loaded to failure. Cylinders experiencing prebuckling matrix cracking were loaded to
their ultimate failure.
Fuchs et al. compared their experimental results with analytical predictions for both perfect
and imperfect cylinder geometries. For evaluation of the prebuckling responses and gaining
insight into the effects of non-ideal boundary conditions and initial geometric imperfections,
they developed a geometrically nonlinear special-purpose analysis based on Donnell's non-
linear shell equations. STAGS C-1 (see [ 14.88]) geometrically nonlinear finite-element anal-
ysis was used to verify the prebuckling solutions of the special purpose analysis. Based on
the experimental and theoretical results, it was concluded that a geometrically nonlinear
boundary layer behavior characterized the prebuckling responses. This boundary layer was
sensitive to laminate orthotropy, cylinder geometry, initial geometric imperfections, applied
end rotation and non-ideal boundary conditions.
STAGS C-1 was also used to study the buckling and postbuckling behavior of both geo-
metrically perfect and imperfect cylinders. In studying the imperfect cylinders, an analytical
approximation of the recorded shape imperfection was employed to introduce the initial
shape imperfections into the STAGS C-1 analysis (see Appendix F of [14.192]). It was shown
that buckling end rotations, strains, and moments were influenced by laminate orthotropy
and initial shape imperfections. Observed buckling moments correlated well with predictions
that accounted for geometrical imperfections.
In postbuckling it was shown that experimental buckling patterns and measured strain
profiles were very similar in some cases. In the failure tests it was found that ultimate failure
could be attributed to an interlaminar shear failure mode along the nodal lines of the post-
buckling patterns.

d. Shear Buckling Experiments


Use of composite shear panels instead of metallic ones in thin-walled primary structures,
particularly those that exploit postbuckling strength, can oiler considerably more load-bear-
ing capacity as well as enhance structural efficiency. Therefore, composite shear webs are
increasingly being introduced in the design of new generations of advanced thin-walled
structures, particularly in aerospace applications. Nevertheless, the behavior of composite
shear webs has hardly been touched by research, and barely any studies appeared in the
open literature; in particular, practically no postbuckling and experimental investigations have
been reported.
In view of this deficiency, a research program was initiated by Wolf and Kossira at the
Institut fiir Flugzeugbau und Leichtbau (IFL) of the Technical University Braunschweig,
Germany, to extend the frontiers of knowledge on the buckling behavior of composite cy-
lindrical shear panels, covering both analytical and experimental aspects [14.193] and
[ 14.194]. The experimental part of the program yielded a fundamental understanding of the
nonlinear behavior and the collapse mechanisms in the postbuckled regime and provided
data for assessing the applicability of nonlinear analysis tools, such as the FiPPS finite-
element code developed at the IFL TU Brunswick [ 14.193].
Wolf and Kossira [ 14.194] reviewed the few experimental investigations on composite
curved shear panels that were reported in the open literature. These investigations included
the 1947 studies of Kuenzi [14.195], the 1962 investigations of Bert, Crisman and Nordby
[14.196], the 1974 experiments of Wilkins and Olson [14.197] and the 1979 and 1986
Deutsche Airbus (MBB) tests (for details of the MBB experiments, which employed a test
rig similar in concept to that depicted in Figure 12.32a and described in Subsection 12.3.8,
see [14.194]).
Bert et al. (see [14.196]) carried out full-scale tests on thin-walled sandwich shells, cylin-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

drical panels and full-scale cylinders, which were fabricated from two-ply epoxy fiber glass
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1213

(E-glass, I 8 I style fabric, 828 Z epoxy) lam inate facings separately honded to an alumi num
honeycomb core. The materials used for manufacturing the specimens, the special tooling
and fabrication processes and procedures for both the pre prcg material and shell specimens
are described in detail in their paper. To simulate the typical load ing in an aircraft panel
bordered by spars and stringers. Be11 ct al. tested their shear panels in a special gripping
dev ice consisting of ring end grips and grooved edge grips (Figure 14. I 80). T he edge and
end g rips were linked by ti lting pins and the device was therefore analogous to the picture
frame (d iscussed in Subsection 8.4.2 of Chapter 8). The gripping device wa~ mounted in a
torsion-bendi ng test ing rig via concentric steel rings on both ends. A torsion arm, connected
at o ne end to the torque plate A (F ig ure 14. 180) and loaded at its other end by a hydraulic
jack, was used to introd uce the tOrque into the panel. T he correlati on of buckl ing test results
of the composite panels with ca lculated va lues was simi lar to that experienced wi th homo-
geneous isotropic ones.

Wilki ns and Olson [ 14.197] employed in thei r experimental program a test technique
simi lar to that of Bert et al. [ 14.196]. Their tests foc used o n answering the following three
questions:
I. Can the experimental buckli ng load of a curved panel be determined no ndestruct ively?
2. Can the experimental buckli ng load be predicted by analysis?
3. How much load beyond buckling can such a panel sustai n?
To ach ieve their goals Wi lkins and Olson performed tests on o ne eight-ply :!: 45 stacking
graphite-epoxy and one eight-ply :!: 45 stacking boron-epoxy specimen of the type shown in
F ig ure 14.1 8 1. These complete composite cyl inders, 9 in. (228.6 mm) long and 12 in. (304.8
mm) in radius, were assembled of four quarter-panel sections, which were joi ned by T-shaped
fasteners and auached to L-shaped end-rings, thro ug h wh ich torque was imroduced into the
specimen to induce the shear stresses. Note that because the panels consisted of o nly a few
plies, stacking resulted in imba lance in bending. T herefore, the shear capability of the cyl-
inders strongly depended on the direction of shear application, and hence each cylinder was
twisted in both directio ns with respect to the cylinder axis.
To mon itor and record the behavior of the specimens. strain gage roselles were installed
back to back on their inner and o uter surfaces. at the center of each of the four panels. Also,
electrical de nection gages were insta lled inside the spec imens to record their internal de-
flections. The strain gage data were used for nondestructive determi nation of the buckli ng
load of each panel by a "modified" Southwell method, which was an extension of the studies
of [9.94] and [9.214]. Determination of the buckli ng loads by the Sout hwell method required,
however. loading the specimens up to 75-90 percent of their buckli ng load.
The test results were compared wit h classical shell theory buckli ng predictions for both
clamped and simply supported loaded edges. Good correlation between the buckli ng loads
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`-

Figure 14.180 Ben et al."s gripping device for shear Jest of composite cylindrical panels (from [ 14. 196])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1214 Composite Structures

Figure 14. 181 Wilkins and Olson's curved-panel assembly for shear tests of composite cylindrical
panels (from [14. 197 ))

determined by the Southwell method and the predictions correspondi ng to the clamped
boundary conditions was obtained by Wilkins and Olson. They also reported that the shear
panels had a postbuckling strength.

Wolf and Kossira pointed o ut in 1992 (14.194] that the test techniques of 114.1951-
[14. 197] and the Deutsche Ai rbus (MBB) experiments did not comply with the necessary
requi rements fo r experimental data to be suitable for verification of analytical methods, i.e.
there was no clear defi ni tio n of loading and boundmy conditions, particularly for postbuck-
ling analysis, and no simple analytical or numerical modeling of test configurations. so that
all relevant physical effects could be included in the analysis. was provided. Conseq uently,
Wolf and Kossira searched for an alternative technique. bearing in mind the requirements
for cost effectiveness and the applicability of the test technique also to panels of various
curvatures. Taking these requirements into account, Wolf and Kossira chose the picture frame
method used in testing of flat shear panels (and discussed in Subsection 8.4.2 of Chapter 8)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

for their experiments.


Since the picture frame is a planar testing device, four pairs of alu minum filler modules,
milled tO fi t the panel curvature and bolted to the frame edge members, were used to ac-
commodate the curved panels int.o the plane picture frame (see F igure 14.182). Notches were
cut at the corners o utside the test area (400 x 400 mm' ) to allow unconstrained rotation of
the frame edge members. It is apparent from Figure 14.182 that usc of the plane picture
frame for loading the curved panels introduces load eccentricity. Wolf and Koss ira fo und,
however, that the picture frame had enough bending stiffness to eliminate this problem when
testing shallow panels.

fitter modules rotational oxls

Figure 14.182 Wolf and Kossira's picwre frame for shear loading of curved composite panels (from
[14. 194])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1215

It should be noted that real curved panels are loaded in a direction tangential to their
surface, whereas in the planar picture frame the specimen is exposed to deformation in a
projection plane parallel to the middle plane of the frame (Figure 14.182). To validate the
picture frame technique for testing of shallow shells like those used in the Braunschweig
tests, the responses of both a fiat panel and a curved one of the same plan form were
numerically simulated by a finite-element analysis. It was shown in [14.193] and [l4.194]
that both panels experienced identical responses. It was therefore deduced that the applied
load introduction concept had no effect on the responses of the curved tested panels, even
in the postbuckling deep regime.
Hence, Wolf and Kossira concluded that the picture frame method had a number of ad-
vantages compared to the other test methods of [14.195]-[14.197]: "It was applicable in
combination with any picture frame device, the load could be applied using standard testing
machines, both curved and fiat panels could be investigated using the same equipment and
the panel curvature could be varied with minor effort." Moreover, it was shown that since
for a constant shape parameter Z = a 2 I rt (where a and r are the height and the radius of
the panel, respectively, defined in Figure 14.182, and t is the panel thickness) and identical
material properties, curved panels exhibit a similar buckling behavior in the linear elastic
regime, the behavior of a deep panel (with large a/ r) may be simulated by testing a shallow
panel (large r) with reduced thickness.
To determine the shear deformation of the panel, the elongation of one diagonal of the
panel and shortening of the other, due to a tensile load applied by a 200 kN capacity test
machine, were measured by inductive displacement transducers attached to the frame. Ge-
ometric shape imperfections and buckling modes were also determined by a deflection trans-
ducer, attached to a carriage running on a slideway along the horizontal diagonal of the

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
panel. Strain gage rosettes positioned back to back were used to measure normal and shear
strains at predetermined points of interest. Prior to testing, the initial shape imperfections
were determined along the horizontal diagonal.
The tests were performed with two types of shallow panels, fabricated from unidirectional
graphite-epoxy prepreg sheets. The panels had a radius of 4000 mm and either four-ply
( ± 45), or eight-ply ( ± 45) 2s lay-ups. The test results of the panels were compared with the
predictions obtained by the nonlinear finite-element code FiPPS. Good correlation was ob-
served between the experimental and theoretical results. A typical comparison is presented
in Figure 14.183. Wolf and Kossira concluded that this type of good agreement proved the
efficiency and applicability of the adopted picture frame method.

e. Lateral Loading Buckling Experiments

Panels under point and pressure loadings- Buckling of curved panels subjected to lateral
point or pressure loadings is associated with the snap-buckling phenomenon. For isotropic
materials, the problem was the subject of many analytical studies as well as some experi-
mental investigations. However, for laminated composite materials snap-buckling received
very little attention, and apparently the only fully reported experimental investigation in the
open literature is that of Marshall, Rhodes and Banks, published in 1977 (see [14.198]. As
indicated by Marshall et a!., their tests aimed at complementing their analytical investigation
on nonlinear behavior of thin orthotropic curved panels under lateral loading (presented in
[ 14.199]). Then, a decade later, some vague hints about another study on sandwich panels
appeared in [14.200].
The panels tested by Marshall et a!. were of a spherical shape and orthotropic wall con-
struction. They were fabricated from unidirectional E-type Tyglas impregnated in a matrix
of Crystic 272 polyester resin. The manufacturing process of the panels, the extensive survey
performed to determine their thickness statistically and ascertain their initial curvature, as
well as the experimental characterization of the composite material mechanical properties
and fiber volume fraction, were described in detail in [14.198].
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1216 Composite Structures

! 35
PANEL S2 ·1

ooe Exl
--- FE~-!, gl!om nonl
PANEL
- FEH,geolft.nonl • COLLAPS
& :not tmlur!

20 '"""'"'"!"--, _
FIRST :IBR( fAILURE -----tS a--
15
0

0 •
0
10 0
FIRST .~LY FAILURE o_______ 8 7 _-
{MATRIX: FAILURfl o .
0
0

0
0 -1 -2 -3 -4 -5 -6 -7 -8 -9 -10
~It-

Figure 14.183 Typical comparison of Wolf and Kossira's shear test results with their FEM predictions
(from [14.193])

Each fabricated panel was machined to a rectangular planform of 406.4 X 203.2 mm 2


(aspect ratio of A = 2.0). After completion of the test with this aspect ratio, the panel was
subsequently machined in steps of 25.4 mm each so that it could be further tested at different
aspect ratios A = 1.75, 1.50, 1.25 and 1.00. Thus, five experiments were performed per
panel, which provided consistency to the experimental values.
The experimental rig used to test the panels under central point loading is depicted in
Figure 14.184 and the rig employed to subject them under pressure loading is shown in
Figure 14.185. In the tests, the panels were either simply supported by edge support boxes
(Figure 14.186a) or clamped (Figure 14.186b). Referring to the central loading (Figure
14.184), increments of deflection were introduced by means of the upper screw rod C. In
response to this deflection, a moment was developed about the fulcrum B, thus transmitting
a force to the proving ring P. The distance from the fulcrum to the ring P was equal to that
from the fulcrum to the rod C. Thus, the reaction force from the test panel was directly
measured by the proving ring.
The datum position for measurements was achieved, by bringing the point of the rod C
into slight contact with the panel and then applying a very slight initial load in the proving
rings by the jacking screw A. By successive adjustments of C and A, a datum position could
be obtained. Hence, to establish the response of the panel, a series of deflections could be
applied to the panel, and the corresponding forces that caused them could be measured. To
check the repeatability of the test results, two sets of results were taken from each test, and
only a negligible difference was observed between them.
In the lateral pressure tests, the support boxes were inverted such that the test panel formed
one side of the pressure chamber (Figure 14.185). Pressure was slowly supplied to the
chamber via a control valve, thus slowly increasing the pressure on the panel surface. A
time interval was allowed for steady-state conditions to be reached, to ensure reliable deter-
mination of the panel pressure versus deflection behavior. As in the point load tests, two
sets of records were taken at each test to examine the repeatability of the experiment.
The test results were correlated with theoretical predictions obtained by the analysis de-
veloped by Marshall et al. [14.199]. A typical comparison study for lateral pressure loading
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(A = 1.00) is shown in Figure 14.187. Very good agreement between theory and experiment
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1217

/ Botoi'ICt Lt"t'
a /~t'Kr~we<J
WftQh l fu!crumB
/ leve-r
-a:.. !. .: p
&!',....
ffi
r.:~
~IOC) f:

II "'I
0•"'1./"'i p c ;>
'"9
J OclotfloCJ - - ~
Sl(f f W A
LO•~r «"ft•e~
•Od(
ao..

Figure 14.18-1 Marshall et al.'> test rig for testing compos11e cur>ed panels under lateral ccmral point
loading (from [ 14. 198])

is observed in this figure. The figure shows that whereas snap-buckling was predicted for
clamped boundary condit ions, no snap was expected for si mple suppo rts; indeed. very close
to the ex perimemal behavior. Similar resuhs 10 those depicted in Fig ure I 4. I 87 were reported
in 114. 198] for the central point loading.
h was also shown by Marshall et al. that in comrast 10 the theoretical predictions. which
a~~umed symmetric deflections throughout the loading range and the occurrence of snap-
buckling. unsymmctric dcHcctions and bifurcation buckling rather than snap buckling were
experienced in the tests for A i!: 1.50. This unsymmetric behavior became increasingly evi-
dcm with increase in aspect ratio. As poimed out by Marshall et al., ·•11 is imponant from
a design point of view since analytical predictions ba•ed on symmetrical panel de fl ections
wou ld be grossly in error." Marshall et al. indicated that probably they were the only ones

Ill Ill

Fi~ure 14. 185 Marshall et nl.'s test rig for testing composite cutvcd panels under ln1cral uniform
pressure loading (from [ 14.198])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1218 Composite Structures

Ca l

Figure 14.186 Edge <uppo11S used in Mar~hall cl al.'s buckling


lc>ls of composi1e curved panel; under laiCral
loading (from fl4. 198]): (a) supporl hoxcs used
for ; imply supported panels. (b) suppon boxes
(b) u>ed for clamped panels

to observe the experiment:ll evidence of unsymmetrical buckling in laterally loaded cuned


panels.

Cyli11ders subjecT To e.rtenral pressure- Apparently no ad hoc experimenL~ on buckling of


composite shells under ex ternal pressure have been reported in the open literawre. However,
experimental and theoretical studies o n the buckli ng of glass-reinforced plastic cylindrical
shells subject to external pressure were carried out by Abu-Farsakh and Lusher 114.20 I ) as
pan of a broader-scope program on buckling under combi ned axial comprc•sion and external
pressure. Five specimen' out of 20 included in thei r experiments. made of a glass-reinforced
plastic-type woven roving (WR) cloth (warp:weft = I: I) embedded in heat-rel.i,tant polyester
resin. were tested under external pressure only. All of the cylinders had a nominal diameter
of d - 305.3 nun and a shell wall of I mm in thickne\s, constructed of three Iayen. of equal
nominal fiber thickness (0.30 mm). each having the same fiber orientation. Cylinders with
length to radius ratios I. 2, 3 and 4 were invest igated in the program.
The test fixture of 114.2011 is depicted in Figure 14.1R8. In this rig the top and hottom
ends of the cylinder ( I) were inserted in grooves (6) :md (7) in the upper loadi ng ring (3)

.600f
-
w
p,_.,\1,., ~OOtd 1•1 00

•' t~eu~ll
.0 CfP I )72~ ·
.;
.:•oo • Loodof'q
'" •
I(
Ul\loo~tng
M t4:WI ot ~OWI9
On.d uniOOdtnq

..
~-J • •
Fi~:urc 14.187 Comp:uison of cxperimcnlal responses of coonp<>silc curved panel subjccl~d to laleral
prcs;urc ob>crvcd in Mar,.hall el al.'s t~st~ with their theorelical prcclicl ion' (from
[14. 1981)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1219

(b)

(o)
figure 14.188 Urmel'\11) of London P"'"""' chamber (from [I~ 101J): (a) details of end fixities and
connecuon of upper loodrng nng to P"'"""' chamber. (b) general 're" of pressure
ch•rnber in axial comprc"ion machine

and lower baseplate (2), which were fil led with a low-melt ing-point al loy (MCP-70) to
provide end fixit y. To apply cxtemal pressure on the cylinder. a pressure chamber was formed
by enclosing the ~pccimen by a series of steel cyl indrical segmenLs bolted to one another in
\tages according to tlte length of the 'pccimen (see Figure 14.188b). The lower cylindrical
segment was bolted to the baseplate (2). and the upper 'egment (5) was connected to the
upper loading ring (3) through a nexible gasket element (8). The gasket was clamped by a
ring (9) to the upper loading ring (3) and by ring ( I 0) to the upper cylindrical segment (5).
thus enclosing the pressure chamber. A diaphragm pump (Edwards High Vacuum , Model
Dl) provided external pressure. A Mano;tat control va lve (John Watson and Smith Ltd ..
type MR3/1/SR/240V) controlled the air pre\sure during the tesls. The pressure inside the
te't rig was recorded by a pressure trJn\ducer <RDP Electronics Ltd .. type P5/50) backed
up by a mechanical pressure gage.
Axial compression loading was applied through the loading plate (4) hy a Mohr and
Federhoff testing machine. A 76-mm diameter steel ball placed at the center of the plate (4)
was employed to en~ure, as far as possible. symmetry and uniformity of the applied axial
load. In add itio n three load cells (I I). on top of the three screw jacks placed under the
bottom plate (2). were used for monitoring and adjusting the unifonnity of the applied axial
load distribution. a~ well as the magnitude of the applied axial load.
E,aJuation of the effect of initial geometric imperfections on the buckling capacity of
glas!>-reinforced pla,tic cylinders CORPs) was one of the major goals of the progran1. To
meet this objective. two displacement tran•ducers were placed on a central 25-mm outer
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

diameter accurately mach ined stainless steel tube to measure the init ial geometrical shape of
the cylinders. a~ well as to record the top and mid-height radial displacement;, during the
experiment. A stepper motor ( 1.8"/step) was used in double \leps to ro tate the central shaft.
The experimental re!>ults were compared with predictions obtained by a nonlinear finite-
clement analy'i' program. NONL5. de,elopcd within the framework of the im estigation of
114.201]. It wa> >hown that buckli ng loads of composite cylindrical shells subjected to hy-
dro•tatic pressure were less sensitive to presence of initial geometric imperfections than
axially loaded shells. In some of the tests it was observed that the action of imperfections
might be reversed and result in a slight increase in buckling pressure.

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1220 Composite Structures

f. Combined Loading Buckling Experiments


Most curved panel and shell clements have to be ucsigned to resist combined states of
loadi ng. Nevertheless. the majority of past analyses. and in part icular experimental studies.
were limited to buckling subject to a single simple type of loadi ng. Only a few investigations
of the buckling of unisotropic panels and shells subjected 10 combined loading were reported.
Among these arc the studies of Bert, Crisman and Nordby [ 14.202]. Wilkins and Love
( 14.203]. Booton and Tennyson ( 14.154]. Herakovich and Johnson ( 14.2~1 and Abu-Farsakh
and Lu~her [14.2011.11 should be noted that designs that neglect interaction effect~ result in
le~~ weight-efficient panels and shells. Moreo' cr. important issues such as geometrical im-
perfection sensitivity and the possibility of selecting appropriate laminate orientation~ yield-
ing imperfection-tolerant laminate configurations, load imperfection sensitivity and the innu-
ence of the sequence of loading on correlation between theory and experiments ure then not
pmpcrly assessed.
Bert ct al. [ 14.202] reported the resu lts of an experimental and analytical investi gation on
buckling under combined bending and torsion of large sandwich-type circular cy lindrical
and tnmcated conical shells with glass-epoxy facings and aluminum honeycomb core~. This
study complemented an earlier one performed by Ben and his co-workers [ 14.1961. in which
the information and data about manufacturing and te~ting of the specimens. including the
prcprcg used to fabricate the cylinder facings. the considemtions associated with the buckling
design of the sandwich-type cyli nders and the description of the University of Toronto 3
million in.-lb (339 MN mm) torsion-bending machine employed for testing (sec Figure
14.189). were given.
Torque in the loading machine was applied by mean~ of a hydrau lic jack and '' hand-
operated pump (A in Figure 14. 189). Pure bending moments were introduced by cables
connected with a common jack (D in the figure). Roller~ prevented any ~ign ifi cant com-
pressive load from being applied to the cylinders. Tite te!.l cyl inders were mounted in the
torsion-bending machine 'ia +in. ( 101.6-mm) deep concentric steel ring~ (8 in Figure
14.189). The specimens were fitted and aligned in the machine and then tack bonded. To
maintain alignment, the mounting rings were subsequently connected by steel columns and
the assembly was removed for casting the ends of the cylinders in place.
The test results were Cll111pared with analytical predictions yielded by the following critical
interaction equation. proposed a decade earl ier by Wang e1 al. (see [14.205]):
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

( 14.4 1)
where (T.,)r&s and (u.,),..., arc the critical buckling torsional and bending stres~es. respec-
tively. under combined torque and bending loading. and (T, ), and (u" )8 arc the correspond-
ing critical stresses under torque or bending only. The limited experimental data was found
to fit this interaction curve [Eq. ( 14.41 )] reasonably well. Seveml other interaction equations
were suggested in [14.2081. Si nce. however, only three specimens were tested, the te~l results
could not conclusively valid<lle any one of the interaction eq uations for use in design.

Copyright Wiley Figure 14.189 Unovcl'\ily of Oklahoma torsion-bendong machine (from (14. 1961>
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1221

Wilkins and Love [14.203] and Herakovich and Johnson [14.204] studied experimentally
the buckling of graphite-epoxy and boron-epoxy cylinders subjected to combined compres-
sion and torsion. Wilkins and Love investigated the effects of the material type, boron-epoxy
or graphite-epoxy and of laminate configurations, four-ply ( ± 45) 5 and six-ply (0/ ± 45}1, on
the buckling interactions. Laminates of such stacking sequences exhibit significant bending-
twisting coupling, and their torsional stiffnesses for positive torques are different from those
for negative torques.
Wilkins and Love [ 14.203] presented detailed information on the fabrication of the test
cylinders. The overall geometry of the cylinders was maintained constant, an outside diameter
of 15 in. (381 mm) and a test section length of 15 in. (381 mm). The lateral stiffness survey
method of Craig and Duggan [9.219] was used to define low stiffness areas, where buckling
was assumed to initiate, and subsequently to determine the critical load by means of strain
gage rosettes installed in these areas (for a detailed description of the application of the
method see Subsection 15.3.1 of Chapter 15). A typical interaction curve obtained by Wilkins
and Love is depicted in Figure 14.190. Comparison of the test results with analytical pre-
dictions revealed an apparent knock-down factor of 65 percent. The most interesting aspect
of the tests results of Figure 14.190 is the near-linearity of the interaction curve over a wide
range of stresses. It was found by calculations that within the linear range of the interaction
curve the torsional buckling stress was reduced by 10-20 percent of the increase in the
applied compressive stress.
In the studies of Herakovich and Johnson [14.204], the test cylinders of each material,
AVCO 5505/4 boron-epoxy and Modmor I/Narmc 5208 graphite-epoxy, were fabricated
with the following lay-ups: four-ply (+AS),, eight-ply ( -45 2 / 45 2 ) 5 and eight-ply (0/ ± 45 I
90) 5 symmetric laminates and a four-ply ( -82.5/30/20/ -82.5) unsymmetric laminate. The
cylinders were approximately 152.4 mm in diameter, most of them were 508 mm long and
their wall thickness ranged from a minimum of 0.53 mm to a maximum of 1.37 mm, in
accordance with the number of plies. Respectively, the radius-to-wall thickness (RI t) of the
cylinders varied from 55-142 and their length to radius ratio (LIR) from 4.4-6.7.
Aluminum fixtures bonded to the end of the specimens were used to introduce load into
the cylinders. The end fixture consisted of a stepped cylindrical base, which extended into
the composite cylinder, and an aluminum collar around the outside of the cylinder. Fixed-
end conditions were achieved by adhesive bonding of all surfaces.
The tests were performed at Virginia Tech on an MTS axial-torsional combined loading
machine with a rated capacity of 222.4 kN axial load and ± 2.26 kN-m torque. Cylinders
with torsional capacity in excess of 2.226 kN-m were tested on a pure torsion machine at
NASA Langley Research Center. In the early tests load control was used, but this proved to
be an unsatisfactory loading mode for conducting more than one test with the same specimen.

1
~ 2000
l±45ls
SOLID POINTS ARE
"' 1500
...: ANALYTICAL PREDICTIONS
~
1000

500
<r • 7800
0~--L---~--~--~--~-cr---
0 I000 2000 3000 4000 5000 6000
COMPRESSION STRESS (psi)

Figure 14.190 Typical buckling interaction observed in tbe tests of Wilkins and Love of a composite
Copyright Wiley cylinder subjected to combined axial compression and shear loading (from [14.203])
--`,`,`````,`,````

Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1222 Composite Structures

Therefore, subsequent tests were run under displacement control or combinations of dis-
placement and load control, which proved to be more satisfactory for determining buckling
without failure. The combined mode was used for the pure compression or torsion tests of
the cylinders made of unsymmetric laminates, to eliminate introduction of undesired loads
due to coupling between in-plane forces and bending moments associated with unsymmetric
laminates. A disadvantage of the combined mode is that proportional loading cannot be
applied.
Herakovich and Johnson compared their experimental results with predictions obtained by
a computer program developed by Wu (for details see [14.204]). They presented both the
test results and theoretical predictions on a dimensionless load plane, whose ordinate is the
normalized axial load ratio ·Rc and whose abscissa is the normalized torsional load ratio R,
to generate the critical interaction curves. Such a typical critical combination curve is shown
in Figure 14.191 for the unsymmetric graphite-epoxy cylinder of lay-up (-82.5/30/20/
-82.5). Herakovich and Johnson observed good correlation between the experimental and
predicted interaction curves. They also found that the shape of the interaction curves was
strongly affected by the cylinder laminate lay-up configuration. It is worth emphasizing that
good practice in comparing theory with experiments requires that real laminate data will be
the input in the computations. In predicting the buckling behavior this requires that material
moduli be experimentally determined as exact as possible. The tests of Herakovich and
Johnson, in which the moduli of each specimen were individually measured prior to per-
forming the buckling test and later used in the correlation study, are a good example of
practicing the above approach.

Booton and Tennyson at UTIAS, Toronto, conducted a theoretical and experimental in-
vestigation of the buckling of imperfect glass-epoxy (Scotchply XP250) circular cylindrical
shells under the simultaneous action of torsion, external pressure and axial compression (see
[14.154] and [9.269]). Their study aimed at demonstrating the influence of laminate lay-up,
geometrical imperfection amplitude and cylinder length on the interaction behavior.
The University of Toronto experimental program of [14.154] and [9.269] consisted of
buckling tests on two three-ply laminated Scotchply XP250 glass-epoxy preimpregnated tape
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

wound cylinders, ( -70/70/0) and (45/0/ -45). The cylinders were supplied by the Struc-
tures Division of the Flight Dynamics Laboratory, Wright-Patterson Air Force Base, and

+ THEORY
+ 0 SPEC. •1~
++ 0 SPEC. •8
1.0 1- +
+
-
+ +
+
+ 0

+0
0.5 1- -
+
+

+
+-

0
I I
·2 .0 ·1.0 1.0 2.0

Figure 14.191 Typical correlation between critical combinations of axial compression and torsion load-
ings of composite cylinders observed in the tests of Herakovich and Johnson and pre-
Copyright Wiley
Provided by IHS Markit under license with WILEY
dicted interaction curves (from [14.204])
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1223

were of the same type employed in the earlier Toronto tests of [14.184]. Each specimen was
clamped at the ends with aluminum end plates bonded to the shell wall with an epoxy plastic.
It was then placed over an internal mandrel in order to prevent excessive postbuckling de-
formations and subsequent fracture, thus allowing repeatable testing of each cylinder. Note
that the lower end plate was actually a ring.
Due to the method of manufacture, the shells contained unintentional random shape im-
perfections (for details on measuring these imperfections to obtain both thickness data and
median surface profiles and the procedure of evaluating the largest root-mean-square value
of the imperfection, used to calculate an equivalent axisymmetric imperfection amplitude,
see Subsection 14.4.1b and [14.185]).
The experimental results of [14.154] and [9.269] were correlated with the approximate
interaction formulae proposed there. Typical nondimensional interaction plots are shown in
Figure 14.192 (in these figures k = R/R,, R,, R, and 7{, are the ratios of TIT,,, PIP" and
pip,,, respectively, T,,, Pu and p,, are the critical torque, critical axial load and critical
pressure, respectively, when each of the loads acts alone, and T, P and p are the applied
torque, axial load and pressure corresponding to the simultaneous critical combinations of
these loads. It is apparent from Figure 14.192 that experimental results agreed reasonably
well with predictions based on the largest measured rms imperfection amplitude and a critical
axisymmetric buckling mode. Correlations of similar quality between theory and experiments
were observed in [14.154] and [9.269] for both the test cylinders and for all of the combi-
nations of loadings investigated there. On the basis of these observations, it can be suggested
that the reduction in buckling strength in the presence of small amplitude random imperfec-
tion can be obtained by calculating the buckling loads on the basis of an "equivalent"
deterministic imperfection (axisymmetric mode).

14.4.2 Stiffened Panels and Shells


Very few experimental and theoretical investigations on buckling and postbuckling behavior
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

of composite-stiffened-curved panels and shells (all of which happen to be related to aero-


space applications) are documented in the open literature. Among these are the analytical
and experimental studies carried out at the Aircraft Division of the Northrop Corporation,
Hawthorne, California, by Agarwal [14.206] and by Deo, Agarwal and Madenci [14.8]; at
the NASA Langley Research Center by Knight and Starnes [14.207]; at the Lockheed Mis-
siles and Space Co. Inc., Palo Alto, California, by Bushnell et al. [13.134]; and at the Israel
Aircraft Industry ( IAI) by Shalev, Segal and Frostig [13.117]. Joint programs of NASA
Langley Research Center, Lockheed Engineering and Sciences Company and Boeing Com-

10 10

~ • -2 I

'o-----!:--.u..._-~
00

Figure 14.192 Typical correlation between critical combinations of torsion, (R, = k,R,) external pres-
sure and axial compression observed by Booton and Tennyson in tests of imperfect
Copyright Wiley glass-epoxy cylinders and predicted interaction curves (from [14.154))
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1224 Composite Structures

mercia! Airplane group and the Douglas Aircraft Company in California and the ALENIA
Company in Italy were reported by McGowan et al. [14.208] and by Bucci and Mercuria
[14.128]. These studies aimed at demonstrating that, as in the design practice of metal aircraft
structural components, which in order to enhance structural efficiency allows local buckling
between the stiffeners to occur below limit load (sometimes significantly below, as in aircraft
fuselage), the design of postbuckled composite stiffened-curved panels and shells can be as
viable and even more efficient. To meet this goal, the above tests were conducted to gain
better understanding of the strength limits and failure characteristics of these postbuckled
composite structural components. The following discussion will focus on the experimental
investigations by the Northrop Corporation, NASA Langley and Lockheed.

The Northrop Corporation tests- Agarwal [14.206] presented the results of a combined
experimental and analytical study on the postbuckling behavior of four curved-hat-stiffened
panels (see Figure 14.193) subjected to axial compression. In the analytical part a design
methodology for composite panels was defined and evaluated by correlation with the exper-
imental results, which were obtained from the tests in which the response of the four curved
panels was determined.
The panel configuration selected was typical of a moderately loaded section of a fighter
aircraft fuselage, where buckling was expected at 25 percent of the design limit load. As
indicated by Agarwal, tests of a complete cylinder would have been desirable in order to
model accurately the exact behavior of the fuselage under the prescribed loading. However,
such tests were expensive and very time consuming, so tests with "equivalent" cylindrical
panels (i.e. sustaining a buckling loading equal to that of the complete cylinder) of appro-
priate dimensions and boundary conditions were conducted to simulate the behavior of a
complete cylinder (see discussion in Subsection 14.4.1 b). The width and boundary conditions
of the equivalent panel were determined according to the guidelines provided in [14.209].
To allow maximum interaction between the stiffeners and skin of the panel, the dimensions
of the specimens were optimized for weight (for further details on the determination of the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

panel dimensions, as well as on their fabrication, see [14.206]).


To prepare the panels for testing, their ends were potted in an epoxy compound and then
machined fiat and parallel within close tolerances, a procedure usually adopted for such tests.
The panels were tested fiat between the loading heads of a 50,000 lb (222.4 kN) Baldwin
static test machine. Care was taken to ensure proper head alignment. Strain gages mounted
back to back were employed to ensure uniformity of load distribution and determine the
onset of buckling. The moire technique was used to observe the buckle patterns of the panels
and their development during loading.
Agarwal found that the panels designed with the methodology he proposed failed at ap-
proximately the design failure load and that the observed failure mode was the same as the
predicted one, i.e. crippling of the stiffeners, leading to local separation from the skin, along
with severe damage to the stiffener. Thus, the simulation of complete shells by testing
"equivalent" panels appeared quite satisfactory.

II
lj

lrl
II
Ull
I

Copyright Wiley
Figure 14.193
FIBEA

itf'AT\l
Northrop composite-hat-stiffened curved panels (from [14.2061)
'

li
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1225

Deo et al. [14.8] presented an extensive static and fatigue test program on curved stiffened
graphite-epoxy and metal panels, loaded well into the postbuckling region under either com-
pression or shear loadings. The experimental program aimed at validating the semi-empirical
static analysis methodology proposed in this investigation for predicting the behavior of the
deeply postbuckled panels, as well as of the various design considerations associated with
the development of the methodology (see detailed discussions in Sections 3.1 and 3.2 of
[14.8]). The tests also assessed the applicability of the rigorous Rayleigh-Ritz analysis
method developed in the study, filled data gaps related to design of postbuckled aircraft
structural components (see details in Table 3.1 of [14.8]) and established a reliable database
for the responses of these structures.
Sixteen hat-stiffened composite panels designed in compliance with that semi- empirical
methodology were included in the program. Hat-section stringers were selected because of
their superior efficiency relative to other cross-section shapes. The panels were fabricated
from Hercules A370-5H/3501-6 woven graphite-epoxy and Hercules AS/3501-6 graphite-
epoxy tapes. Fabrication of the panels was discussed in Section 3.4 of [14.8]. Ten of the
panels were tested in shear (see details of panels in Figure 14.194a). Following the design
methodology, a panel web lay-up of (45o/90/45 2 ), a stringer spacing of 10 in. (254 mm)
and frame (J-type) spacing of 24 in. (610 mm) representative of actual fuselage structures
were chosen for the shear panels. The ]-sections were used because of fabrication and at-
tachment considerations. Six of the panels were tested under axial compression (see details
of panels in Figure 14.194b). A web lay-up (45/0/90/0/45), and stiffener spacing of 12.2
in. (31 0 mm) were selected for the compression panels.
Two panels of each loading group were tested under static loading, and the other specimens
were tested in fatigue. The static tests were conducted to obtain data on skin buckling strains,
stiffener crippling strains, skin and stiffener strains at stiffener-web separation and the strain
distributions in the panels as a function of the applied load. Moreover, these data were
required for assessing the applicability of the semi-empirical analysis methodology and for
subsequent development of a design guide for composite stiffened curved shear and com-
pression panels.
The test panels were instrumented with strain gages and LVDT's for monitoring the out-
of-plane displacements during the tests. The moire grid technique was employed for visual
observation of the out-of-plane deflections and the buckled patterns of the panels.
The compression tests were conducted in a I 00,000 lb (445 k:N) capacity Tinius Olsen
test machine. The curved panel ends were potted in an epoxy compound for load introduc-
tion.
The shear panels were loaded in a test fixture designed and developed at Northrop (see
Figure 14.195) that was capable of introducing shear stresses in the panels with no accom-
panying adverse stresses. In the test fixture the curved panels were enclosed by two flat
panels, which were considered part of the fixture. These two panels were connected at a
point midway between the test panel center of curvature and the test panel. One end of the
tube formed by the test panel and flat panels was clamped against all the applicable degrees
of freedom. The other end of the tube was connected to the loading frame plate. It can be
seen in Figure 14.195 that in order to allow free rotation of the loading frame shaft about
the tube centroid, the loading frame support plates were slotted. Shear load was introduced
by a torque applied to the loading plate via two 50,000 lb (222 k:N) load cylinders and a
torque arm of 74 in. (1880 mm). The fixture was designed so as to allow testing of curved
panels in a wide range of sizes and curvatures with minimal alterations.
Based on comparison of the test results of the compression panels with those of the semi-
empirical analysis, Deo ct al. found that prediction of the ultimate load of the compression
panels by this analysis was very accurate and could hence be readily used for design pur-
poses.
In comparing the test results of the shear panels with predictions yielded by the modified
tension field theory, adapted to use with curved shear panels, Deo et al. found the theory to
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
7 PLIES GA IE TA.PE
(0")

TA.PE (to')

0,0572

VIEW F VIEW D

<0.00 ---------!

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
·-·
VIEW D

SECTION 11.-A

ALL DIMENSIONS IN inch.


(a)

FIBER ORIENTATION

1.25
1.25

90'
-45.*>45'
o· IT 51.6~ EPOXY
FIBEJ;l ORIENTATION POTTNG lYP
EACH ENO All
STRINGERS

PLAN 'v1EW
SEC A· A.
(b)

Figure 14.194
Copyright Wiley Northrop composite-curved-stiffened panels (from [14.8]): (a) hat-stiffened shear pan-
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
els-details of geometry and lay-ups
No reproduction or networking permitted without license from IHS
of panel skin and stiffeners, (b) configuration of
Not for Resale, 02/13/2019 01:32:58 MST
hat-stiffened compression panels
Curved Panels and Shells 1227

Figure 14.195 Northrop's test fixture specially designed for combined shear and compression loading
of curved stiffened-panels (from [14.8])

be applicable to composite curved stiffened shear panels. However, the ultimate load pre-
dictions were found to be conservative by about 35 percent.
Deo et a!. also observed that:
• The postbuckling stiffness of the compression panels was reduced by about 40 percent
compared with their initial stiffness.
• The loss in stiffness after buckling of the shear panels was about 55 percent.
• The failure mode of the compression and shear panels was stiffener-web separation (note
that since the measured loads in the compression tests, corresponding to this mode of
failure, closely agreed with the semi-empirical predictions based on a crippling failure
mode, it could be concluded that the stiffener-web separation mode of failure in the
compression panels was induced by stiffener crippling that preceded it).
Based on the shear panel tests, Deo et a!. developed an empirical equation to predict the
stiffener-web separation. They demonstrated that the stringer or ring-forced crippling strains
corresponded to the stiffener-web separation strains.
Concerning the fatigue test results of the buckled panels, Deo et a!. observed that repeated
buckling had no effect on the initial buckling of compression or shear panels. Based on the
fatigue failure modes experienced in the tests, they also developed an approach to life pre-
diction for buckled composite curved stiffened compression and shear panels.

The NASA Langley tests- Knight and Starnes reported in [14.207] the results of an exper-
imental and analytical study of the postbuckling behavior and failure characteristics of eight
curved !-stiffened graphite-epoxy (Thorne! 300 tapes preimpregnated with 450 K cure
Narmco 5208 thermosetting epoxy resin) panels loaded in axial compression. Note that the
stiffeners of all specimens were secondarily bonded to the skin. The panels were designed
and manufactured by the Lockheed-Georgia Company (for details see [14.207]). Panels of
three different lengths (508, 660 and 813 mm), three different stiffener spacings (102, 140
and 178 mm), three different panel widths (arc lengths) (381, 495 and 610 mm) and a
nominal radius of 2160 mm were included in the test program. The stacking sequences of
the panel skin, stiffener web and stiffener cap were (±45/02 /+45/90 2 )s, (0/ ±45/+45/02 /
±45/0,) and (0/ ±45/+45/02 / ±45/02 /90/0 2 /90 2 /0 2 )., respectively. Note that except for
the curvature, the dimensions of the tested panels were identical with those of the stiffened
fiat panels of [14.102] shown in Figure 14.95. Also, the stacking sequences were identical
with those of design type 2 in Table 14.2.
Following cure, the panels were ultrasonically inspected. No detectable defects were re-
Copyright corded in any specimen. The panel curved ends were again potted in an epoxy-resin material
--`,`,`````,`,````

Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1228 Composite Structures

to prevent brooming of the fibers during loading of the panel. The potting material was
enclosed in a steel frame for protection. The curved ends of the panel were machined fiat
and parallel to ensure uniform loading. The unstiffened side of the panel was painted white
for application of the moire techniques. On several of the panels the initial geometrical
imperfections were surveyed. These records were later used as input in the STAGS C-1
[14.88] analysis to evaluate the influence of measured surface imperfections on the predicted
postbuckling behavior of the panels.
The panels were loaded in axial compression by a 4.45 MN capacity hydraulic testing
machine. The specimens were fiat-end tested without supporting their straight lateral edges.
Strain gages were used to monitor strains, and DCDT's were employed to monitor displace-
ments at selected points. Moire fringe patterns were recorded photographically.
Both a linear buckling analysis using the PASCO computer code [14.210] and nonlinear
postbuckling calculations using STAGS C-1 were performed to predict the behavior of the
panel. The buckling calculations were executed to predict the initial buckling of the panel,
as well as to determine the level of modeling detail required to accurately predict the post-
buckling response of the panel. The POSTOP computer code [14.211] and [14.212] was
used to determine the stress distributions in the interface region between the panel skin and
the stiffener attachment flange.
The STAGS C-1 predictions were compared with the test results. Good correlation between
analysis and experiments was found up to the load where local failure initiated, provided
that two effects were accounted for: (1) since the analytical buckling mode had considerable
deformation in the stiffener webs and attachment flanges, these stiffener components had to
be modeled with shell elements of appropriate stiffness to obtain satisfactory correlation with
the postbuckling tests results; and (2) use of measured initial shape imperfections in modeling
the panel skin geometry.
Knight and Starnes also found that all the specimens experienced a stiffened-panel buck-
ling mode, where both the skin and the stiffeners deformed. Failure of all panels initiated in
a skin-stiffener interface region. The existence of high membrane strains at the stiffeners
after skin buckling were apparently suf1icient to trigger a local skin stiffener separation,

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
which led to comprehensive panel failure. Furthermore, they observed that local damage,
induced prior to failure, could change the panel stiffness enough to affect the postbuckling
response in repeated buckling tests.

The Lockheed tests- The tests carried out by Bushnell et al. at Lockheed Missiles and Space
Company in the mid-eighties (see [13.134]) were discussed in Subsection 13.6.3 of Chapter
13. There the discussion focused on the unique features of the test frame developed to
perform tests on large-scale, curved graphite-epoxy hat-stiffened panels.
The specimens tested in [13.134] represented part of the fuselage of a large air transport:
194 in. (4298 mm) in diameter, 30 in. (762 mm) long, 27 in. (686 mm) wide (arc length)
and 8 in. (203 mm) stringer spacing, and were designed to buckle locally at loads well below
the ultimate load of 3000 lb!in. (525 N/mm). The optimum design of these large-scale
panels was obtained by the PANDA 2 code (see [13.135]-[13.137]).
Early optimum design blade-stiffened panels employed in the experiments of [13.134]
failed at loads significantly below the desired ultimate load. Therefore, based on the studies
of Agarwal (see [14.206]) and others, a hat-type cross-section was chosen for the stringers.
Optimum design studies with PANDA 2 yielded the panel configuration depicted in Figure
14.196. Since these large, composite-curved, stiffened panels were expensive to fabricate and
prepare for testing, small fiat specimens with a cross-section identical to that of the optimum
design of the large panels obtained with PANDA 2 were first fabricated and tested to master
the techniques for fabrication of high-quality curved panels and their testing. Following the
tests with such seven small fiat panels, three large-scale curved panels were fabricated and
tested. Details of the fabrication of both the small flat and large scale curved panels are
given in Section 5 of [13.134].
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
CLOTH TAPE ClOTH
2 PliES ~0.006/PLY
,..............
2 PLIESI0.006 2' PLIES fQ.0052 IN!PLY
,............__
If= 0.17111 in. (ll PLIES (IISc,.•45c f On /90T IOIIT f90T /OAIT /90T I OIJT /9~ /On /'JOT /OH l~.li\ /~Sc J)

1._ = O.On in. (II PliES OF CLOTH 00.006/PlY J - - - - lon.---------1 t( = 0. 17,.8


l115cf-115c!-\\"5cl /r 'i_ TO 'i_ ----1•-~•i+--
-_j_Af
Ill PLACES) TYPICAL --1-
UNIFORM TAPER OVER
, [ o. 375 in. (TYP)

t
1,_ : 0.0}"
(7 PLIES)
DETAIL A DETAIL B l'JII tn R
~0. 006IN 10. OOS2 IN to. 006 ~~~ ~
tn./PLY in./PLY m /PLY ! \

tb"'2lPLIES ~
[ IJ5cf4.q,\ /OlfT/!OT /0 4 T/'JOT/OqT /'90T /OH/-45c !II\] ::::::::...----
~ ~ llltn
CLOTH TAPE CLOTH 8 tn
+45° CLOTH 2 PLIES 1' PLIES 2 PLIES T'r'p

0.
0.006/PLY 0.00'52 m./PLY 0.006fPLY
:;·~
on.

29 PLIES OF TAPE. AS ABOVE ---.. .

All LAYERS OF ADHESIVE

CLOTH = HMF HO/!JII " · 006 tn./PlY


TAPE -:: TlOOI'JH iiO.OOS2 tn./PlY

l LAYERS OF ADHESIVE FILL WITH ROLLED TAPE (FIBERS HAVE G-d~ ORIENTATION I

~ I
" "'"" ' j 0''" ~ 1
'1% I f:::~
4 LAYERS OF ADHESIVE AT 1 "ui!LAYER

~'""'" "'"~" uw """'" "'"" m• ""

) jJ 2 LAYERS OF ADHESIVE
DETAil A

Figure 14.196 Bushnell et al.'s optimum design of compression hat-stiffened panels-details of geometry and lay-ups of panel skin and of hat-stiffeners cross-section
(from [13.134])
...
~
co --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1230 Composite Structures

Though the investigation of [13.134] represented a good example of a modem design and
test methodology of composite curved stiffened panels, local stress concentrations, not buck-
ling, caused failure of two panels. For these panels, agreement between test and theory was
found to be reasonably good with respect to failure loads of these panels and their failure
modes. The third panel failed in a wide-column buckling mode at a load well below that
predicted by PANDA 2 for a panel clamped at its loaded ends. It was suggested in [13.134]
that this premature failure was caused by excessively soft end plates on the test specimen,
whereas the calculations assumed clamped loaded ends.
Nevertheless, the test results on the small fiat panels provided important information. They
showed that early delamination, initiating from the unloaded edges of the panel, could be
prevented by use of many small C-clamps applied to short aluminum tabs distributed along
the sides of the panel in such a way that they neither affected the distribution of axial load
across the panel (by maintaining small gaps between each tab to its neighbors) nor signifi-
cantly stiffened the unloaded boundaries against wide-column general instability. Also, early
delamination and pop-off of the stringer was prevented by introducing thin-film adhesive in
areas where these modes of failure were observed in specimens with no adhesive (see Figure
14.196).

14.4.3 Corrugated Cylinders


In the early eighties a study on failure by buckling in bending of a large-scale corrugated
cylinder fabricated from T00/5208 graphite-epoxy unidirectional tapes was conducted by
Davis at NASA Langley Research Center (see [14.213]). The wall of the cylinder consisted
of three flat corrugated sheets wrapped to the proper cylindrical shape (for fabrication details
see [14.214]-[14.216]). Overall configurations of the 3-m diameter shell, as well as details
of the open corrugations and laminate lay-ups, are given in Figure 14.197. Note, however,
that the shell was 3.05 m long or, including the steel loading rings, 3.15 m long.

Grd~h i le-epoxy
I ·4s),o.os6
L
f-J--~:··-- - ~~~
·~ln ,)#-~ ?=~-
1.55

VIIWC
SECT ION D-0 Graphi te-rpm:y (, 45) Gr-lphl tP-epox~
143/5208 E·GLA.SS (3 plies) (•45/05/•4~5-)
typJCal
Gr<:~ph1te-epoxy (·45) · 11.400

-~~---
T
i.19 _L_=-1· --------
0.056 \ , .

t_[l_J7nd n:-el~
.
~.1.650--1- ~-epoxy
1
---·-u.116
-
1 w-.t,l11Pd wet. Romr-:emperat~rc ( ·45) 5 typical
arlhe<; lVP
!S2.4 radills VlfW B
I

'>hPll <,pl rep,


three cqua 1

Figure 14.197 Configuration and details of Davis's 3-m diameter graphite-epoxy corrugated cylindrical
shells tested in bending (from [14.213])
--`,`,`````,`,````,

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Curved Panels and Shells 1231

Figure .14.198 Details of loading ring and scalloped auachment rings used to introduce bending load
into Davis's large-scale corrugated cylinder (from ( 14.2 13])

Figure 14.199 Test llxture (showing conical loading fixtures) used for loading in bending of Davis's
corrugated cylinder (from [1 4.2 13 ])

o.20J
J.._ /~---~....._.......

Figure 14.200 Spacer blocks placed between scalloped auachment ring and adapter ring to eliminate
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
rolling moment in the loading rings Licensee=McDermott
Provided by IHS Markit under license with WILEY
in Davis's bending test of the large-scale corrugated
Inc - India/8215328006, User=G, Boopathi
cylinder (from [14.2 131)
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1232 Composite Structures

2.0

I Analys1s w1thout
spacer blocks

;---4-__
5
_ ---i __
Analy51S with

] 1.0 j_ spdcerblods

--- ----------e
§
Scalloped Position of
attachment stiffener
nng and r1ng 1
doubler zone
0 Test without spacer blo~ks

j 0 Test with spacer blocks

.1 .2 ,3
Ax i a 1 loc at i 011/ ax i a 1 1ength , X/~

Figure 14.201 Effect of placing of spacer blocks between the adapter ring and the scalloped attachment
rings on eliminating the rolling moment induced displacements near the scalloped at-
tachment rings-comparison of test results with BOSOR 4 predictions (from [14.213])

Bending load was introduced into the shell wall by a set of concentric scalloped attachment
rings, bolted to each end of the cylinder and to aluminum angle-section adapter rings, that
transferred the load from the loading rings to the scalloped attachment rings (see Figure
14.198). The loading rings were bolted to heavy conical loading fixtures, one at one end of
the cylinder mounted on a rigid wall and the other one at the other end attached to the
bending loading frame (see Figure 14.199). This bending test fixture, excluding the conical
fixtures, was described in Subsection 9.7.2 of Chapter 9 (see Figure 9.75b).
A preliminary test of the cylinder at 30 percent of the design load revealed a severe load-
introduction problem at the end of the cylinder. An in-plane ring rolling moment in the
loading rings, which produced substantial shell wall bending at the ends of the cylinder, was
generated due to a 5.56-cm offset between the centroidal radius of the loading ring and the
mid-surface radius of the shell wall.
The BOSOR 4 [2.92] analysis, using a model that included the loading ring, the adapter
ring, scalloped attachment rings and the corrugated shell wall, showed that the scalloped
attachment ring rotated with the rolling of the loading ring and caused a large inward dis-
placement of the shell wall near the scalloped attachment rings. Because of the low shell
wall hoop stiffness, this displacement induced high shell wall bending strains in addition to
the axial strains. Further analysis indicated that the problem could be eliminated by using
properly sized spacer blocks with beveled ends (see Figure 14.200) placed between the
adapter rings and the scalloped attachment rings.
In the tests following the introduction of the spacer blocks, it was observed that the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

bending strains in the shell wall were significantly reduced (see Figure 14.201), thus allowing
successful buckling tests of the cylinder. In the ultimate load test the shell sustained the
desired design ultimate load test before buckling and failed at 101 percent of the design
load.

14.5 Concluding Remarks


Before closing this chapter, it should again be emphasized that there is a scarcity in the open
literature of tests on buckling and collapse of composite structures. In particular, only very
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1233

few experiments exist on composite stiffened shells and panels. Hence, these topics should
be given priority in future investigations.

References
14.1 Curry, J.M., Effect of Ply Drop-Offs on the Strength of Graphite-Epoxy Laminates, M.Sc. thesis,
Virginia Polytechnic Institute and State University, May 1986.
14.2 DiNardo, M.T. and Lagace, P.A., Buckling and Postbuckling of Laminated Composite Plates
with Ply Dropoffs, AIAA Journal, 27, (10), Oct. 1989, 1392-1398.
14.3 Lager, J.R. and June, R.R., Compressive Strength of Boron-Epoxy Composites, Journal of
Composite Materials, 3, Jan. 1969, 48-56.
14.4 Greszczuk, L. B., Micro buckling Failure of Circular Fiber-Reinforced Composites, AIAA Jour-
nal, 13, (10), Oct. 1975, 1311-1318.
14.5 Leissa, A.W., Buckling of Laminated Composite Plates and Shell Panels, Report AFWAL-TR-
85-3069, AF Wright Aeronautical Laboratories, June 1985.
14.6 Tennyson, R.C., Buckling of Laminated Composite Cylinders: A Review, Composites, Jan.
1975, 17-24.
14.7 Finch, D.C .. The Buckling of Symmetric and Unsymmetric Composite Plates with Various
Boundary Conditions, TELAC Report-84-3, Technology Laboratory for Advanced Composites,
Department of Aeronautics and Astronautics, Massachusetts Institute of Technology, Cam-
bridge, Mass., Feb. 1974.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

14.8 Deo, R.B., Agarwal, B.L. and Madenci, E., Design Methodology and Life Analysis of Post-
buckled Metal and Composite Panels, Vol. I, AFWAL-TR-85-3096, AF Wright Aeronautical
Laboratories, Dec. 1985.
14.9 Jensen, D.W., Buckling and Post buckling Behavior of Unbalanced and Unsymmetric Laminated
Graphite/Epoxy Plates, Ph.D. thesis, Massachusetts Institute of Technology, Feb. 1986.
14.10 Whitney, J.M., Structural Analysis of Laminated Anisotropic Plates, Technomic, Lancaster,
Penn., 1987.
14.11 Kapania, R.K. and Raciti, S., Recent Advances in Analysis of Laminated Beams and Plates,
Part I: Shear Effects and Buckling, AIAA Journal, 27, (7), July 1989, 923-934.
14.12 Jones, R.M., Mechanics of Composite Materials, Scripta, Washington, D.C., 1975.
14.13 Hausner, J.M. and Stein, M., Numerical Analysis and Parametric Studies of the Buckling of
Composite Orthotropic Compression and Shear Panels, NASA TN D-7996, Oct. 1975.
14.14 Whitney, J.M., The Effect of Transverse Shear Deformation on the Bending of Laminated Plates,
Journal of Composite Materials, 3, 1969,534-547.
14.15 Chia, C.Y., Nonlinear Analysis of Plates, McGraw Hill, New York, 1980.
14.16 Prabhakara, M.K. and Chia, C.Y., Postbuckling Behavior of Rectangular Orthotropic Plates,
Journal of Mechanical Engineering Sciences, 15, (1), 1973, 25-33.
14.17 Davis, J.G., Jr. and Zender, G.W., Compressive Behavior of Plates Fabricated from Glass Fil-
aments and Epoxy Resin, NASA TN D-3918, April 1967.
14.18 Ashton, J.E. and Love, T.S., Experimental Study of the Stability of Composite Plates, Journal
of Composite Materials, 3, April 1969, 230-242.
14.19 Kicher, T. P. and Mandell, J. F., A Study of the Buckling of Laminated Composite Plates, AJAA
Journal, 9, (4), April 1971, 605-613.
14.20 Willey, B.T., Instability of Glass Fiber Reinforced Plastic Panels Under Axial Compression,
USA AVLABS Technical Report 69-48, Department of Aeronautics and Astronautics, Stanford
University, Stanford, Calif., Sept. 1971.
14.21 Verchery, G., Etude du flambement de structures en materiaux composites, Raport 010, ENSTA
(Paris), 1973.
14.22 Ashton, J.E. and Waddoups, M.E., Analysis of Anisotropic Plates, Journal of Composite Ma-
terials, 3, 1969, 148-165.
14.23 Reissner, E. and Stavsky, Y., Bending and Stretching of Certain Types of Heterogeneous Aeo-
lotropic Elastic Plates, Transactions of the ASME, Journal of Applied Mechanics, 28, (3), Sept.
1961, 402-408.
14.24 Banks, W.M. and Harvey, J.M., Experimental Study of Stability Problems in Composite Ma-
terials, in: Stability Problems in Engineering Structures and Components, T.H. Richards and P.
Stanley, eds., Applied Science Publishers, London, 1978.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1234 Composite Structures

14.25 Marshall, I.H. and Banks, W.M., The Instability of Composite Structures with Initial Imperfec-
tions, in: Stability Problems in Engineering Structures and Components, T.H. Richards and P.
Stanley, eds., Applied Science Publishers, London, 1978.
14.26 Banks, W.M., The Postbuckling Behavior of Composite Panels, in: Proceedings, International
Conference on Composite Structures, Geneva, 1975, 272-293.
14.27 Banks, W.M., Harvey, J.M. and Rhodes, J., The Non-Linear Behavior of Composite Panels
with Alternative Membrane Boundary Conditions on the Unloaded Edges, in: Proceedings,
Second International Conference on Composite Materials, Toronto, Ont., Canada, 1978, 316-
336.
14.28 Fok, W.C., Postbuckling Behavior of Plates with Discontinuous Change of Thickness, Ph.D.
thesis, University of Strathclyde, 1975.
14.29 Jensen, D.W. and Lagace, P.A., Influence of Mechanical Couplings on the Buckling and Post-
buckling of Anisotropic Plates, AIM Journal, 26, (10), Oct. 1988, 1269-1277.
14.30 Lagace, P.A., Jensen, D.W. and Finch, D.C., Buckling of Unsymmetric Composite Laminates,
Composite Structures, 3, 1986, 101-123.
14.31 Minguet, P.J., Buckling of Graphite/Epoxy Sandwich Plates, TELAC Report 86-16, Technology
Laboratory for Advanced Composites, Department of Aeronautics and Astronautics, Massachu-
setts Institute of Technology, Cambridge, Mass., May 1986.
14.32 Minguet, P.J., Dugundji, J. and Lagace, P.A., Buckling and Failure of Sandwich Plates with
Graphite-Epoxy Faces and Various Cores, in: Proceedings, 28th AIM! ASMEIASCE!AHS
Structures, Structural Dynamics and Materials Conference, Monterey, Calif., April 1987, 394-

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
401.
14.33 Lagace, P.A., The Behavior of Highly Loaded Structural Elements, TELAC Report 83-19,
Technology Laboratory for Advanced Composites, Department of Aeronautics and Astronautics,
Massachusetts Institute of Technology, Cambridge, Mass., Nov. 1983.
14.34 Mandel, J.F., An Experimental Study of the Buckling of Anisotropic Plates M.Sc. thesis, Case
Western Reserve University, School of Engineering, June 1968.
14.35 Starnes, J.H. Jr., and Rouse, M., Postbuckling and Failure of Selected Flat Rectangular Graphite-
Epoxy Plates Loaded in Compression, in: Proceedings, AIM! ASME/ASCEIAHS 22nd Struc-
tures, Structural Dynamics and Materials Conference, Atlanta, Ga., April 6-8, 1981, AIAA
Paper No. 81-0543.
14.36 Noor, A.K., Starnes, J.H., Jr. and Waters, W.A. Jr., Numerical and Experimental Simulations
of the Postbuckling Response of Laminated Anisotropic Panels, in: Proceedings, A/MIASM£
IASCEIAHSIASC 31st Structures, Structural Dynamics and Materials Conference, Long
Beach, Calif., April 2-4, 1990, 848-861.
14.37 Engelstad, S.P., Reddy, J.N. and Knight, N.F., Jr., Postbuckling Response and Failure Prediction
of Flat Rectangular Graphite-Epoxy Plates Loaded in Axial Compression, in: Proceedings,
AIMIASMEIASCEIAHSIASC 32nd Structures, Structural Dynamics and Materials Confer-
ence, Baltimore, Md, April 8-10, 1991, 888-895.
14.38 Almroth, B.O. and Brogan, F.A., The STAGS Computer Code, NASA CR-2950, 1978.
14.39 Buskell, N., Davis, G.A.O. and Stevens, K.A., Postbuckling Failure of Composite Panels, in
Composite Structures 3, I. H. Marshall, eel., Elsevier Applied Science, London, New York,
1985, 290-314.
14.40 Chang, K., Jensen, J., Matthews, R. and Parekh, A., Buckling Strength of Carbon/Carbon
Panels, in: Proceedings, AIMIASMEIASCE!AHSIASC 31st Structures, Structural Dynamics
and Materials Conference, Long Beach, Calif., April 2-4, 1990, 914-919.
14.41 Bowen, D.V. and Ayers, K.B., Buckling and Failure of Metal Faced CFRP Sheets in Compres-
sion, Composites, March 1975, 69-74.
14.42 Starnes, J.H., Dickson, J.N. and Rouse, M., Postbuckling Behavior of Graphite-Epoxy Panels,
in: Proceedings, ACEE Composite Structures Technology Conference, Seattle, Wash., August
13-16, 1984, 137-159 (NASA CP 2321).
14.43 Rouse, M., Postbuckling of Flat Unstiffenecl Graphite-Epoxy Plates Loaded in Shear, in: Pro-
ceedings, AIMIASME!ASCEIAHS 26th Structures, Structural Dynamics and Materials Con-
ference, Orlando, Fla., April 15-17, 1985, 605-616.
14.44 Romeo, G. and Gaetani, G., Effect of Low Velocity Impact Damage on the Postbuckling Be-
havior of Composite Panels, in: Proceedings, 17th !CAS Congress, Stockholm, Sweden, Sept.
1990, 994-1004.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1235

14.45 Agarwal, B.L., Postbuckling Behavior of Composite Shear Webs, A/AA Journal, 19, (7), July
1981, 933-939.
14.46 Rich, M.J. and Foye, R.L., Low Cost Composite Airframe Structures, in: Proceedings, 3rd
Conference on Fibrous Composites in Flight Vehicle Design, NASA TMX-3377, Nov. 1975,
225-242.
14.47 Kaminski, B.E. and Ashton, J.E., Diagonal Tension Behavior of Boron-Epoxy Shear Panels,
Journal of Composite Materials, 5, Oct. 1971, 553-558.
14.48 Pimm, J.H., Advanced Composite Tension Field Tests and Evaluation, National Sampe Sym-
posium and Exhibition, San Francisco, Calif., 1979, 1405-1416.
14.49 Foreman, C.R., Design Concepts for Composite Fuselage Structure, in: Fibrous Composites in
Structural Design, E.D. Leone, D. W. Oplinger and J.J. Burke, eds., Plenum Press, New York,
1980, 103-123.
14.50 Ashton, J.E., Analysis of Anisotropic Plates II, Journal of Composite Materials. 3, 1969, 470-
479.
14.51 Ashton, J.E. and Love, T.S., Shear Stability of Laminated Anisotropic Plates, Composite Ma-
terials: Testing and Design, ASTM STP 460, Feb. 1969, 352-361.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
14.52 Tsongas, A. G. and Ratay, R.T., Investigation of Diagonal Tension Beams with Very Thin Stiff-
ened Webs, NASA CR 101854, July 1969.
14.53 Darevics, V.M. and Hoy, J.D., SST Technology Follow-on Program, Phase I, Intermediate Shear
Beam Analysis, Federal Aviation Agency. Washington, D.C., Report FAA-SS-72-11, May 1972.
14.54 Weller, T., Singer, J., Kossira, H. and Arnst, G., Durability of Composite and Metal Deeply
Postbuckled Panels, Joint Research Report, Faculty of Aerospace Engineering, Technion, Israel
and Institut fiir Flugzeugbau und Leichtbau, TU Braunschweig, Germany, July 1995.
14.55 Weller, T., Singer, J., Kossira. H. and Arnst, G., Durability of Composite and Metal Deeply
Postbuckled Panels, Joint Research Report, Faculty of Aerospace Engineering, Technion, Israel,
and Institut fUr Flugzeugbau und Leichtbau, TU Braunschweig, Germany, December 1995.
14.56 Weller, T.. Messer, G. and Libai, A., Repeated Buckling of Graphite-Epoxy Shear Panels with
Bonded Metal Stiffeners, TAE Report No. 546, Department of Aeronautical Engineering, Tech-
nion-Israel Institute of Technology, Haifa, Israel, 1984.
14.57 Bush, H.G., Experimental Evaluation of Two 36" X 47" Graphite/Epoxy Sandwich Shear Webs,
NASA TM X-72767, August 1975.
14.58 Chamis, C.C., Buckling of Anisotropic Composite Plates, in: Proceedings ASCE, Journal of
the Structural Division, 95, (STIO), 1969, 2119-2139.
14.59 Chamis, C.C., Buckling of Anisotropic Plates, Closure and Errata, in: Proceedings ASCE, Jour-
nal of the Structural Division, 97, 1971, 960-962.
14.60 Harris, G.Z., The Buckling and Postbuckling Behavior of Composite Plates under Biaxial Load-
ing, International Journal of Mechanical Sciences, 17, 1975, 187-202.
14.61 Bert, C.W. and Chen, T.L.C., Optimal Design of Composite Material Plates to Resist Buckling
under Biaxial Compression, Transactions JSCM, 2. (1), 1976, 7-10.
14.62 Lukoshevichyus, R.S., Minimizing the Mass of Reinforced Rectangular Plates Compressed in
Two Directions in a Manner Conductive toward Stability, Polymer Mechanics, 12, (6), 1977,
929-933.
14.63 Schmit, L.A. Jr., and Farshi, B., Optimum Design of Laminated Fibre Composite Plates, Inter-
national Journal of Numerical Methods in Engineering, 11, (4), 1977, 623-640.
14.64 Libove, C., Buckling Pattern of Biaxially Compressed Simple Supported Orthotropic Rectan-
gular Plates, Journal of Composite Materials, 17, 1983, 45-48.
14.65 Tung, T.K. and Surdenas, J., Buckling of Rectangular Orthotropic Plates under Biaxial Loading,
Journal of Composite Materials, 21, 1987, 124-128.
14.66 Arnold, R.R. and Parekh, J.C, Buckling, Postbuckling, and Failure of Flat and Shallow-Curved,
Edge-Stiffened Composite Plates Subject to Combined Axial Compression and Shear Loads,
in: Proceedings, AIAAI ASMEI ASCEI AHS, 27th Structures, Structural Dynamics and Materials
Conference, Part 1, San Antonio, Tex., May 19-21, 1986, 769-782.
14.67 Romeo, G. and Frulla, G., Buckling of Simply Supported and Clamped Anisotropic Plates under
Combined Loads, in: Proceedings, International Conference Spacecraft Structures and Me-
chanical Testing, Noordwijk, The Netherlands, April 24-26, 1991, ESA SP-321, 1, 161-166.
14.68 Romeo, G. and Frulla, G., Analytical and Experimental Results on Composite Panels under
Combined Biaxial Compression and Shear Loads, Aerotechnica Missili e spazio, July-Dec.
1991, 107-116.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1236 Composite Structures

14.69 Romeo, G. and Frulla, G., Postbuckling Behavior of Anisotropic Plates under Biaxial Com-
pression and Shear Loads, in: Proceedings of the 18th Congress of the International Council
of the Aeronautical Sciences, Beijing, P.R. China, September 20-25, 1992, (2), 1936-1944.
14.70 Romeo, G. and Frulla, G., Nonlinear Analysis of Anisotropic Plates with Initial Imperfections
and Various Boundary Conditions Subjected to Combined Biaxial Compression and Shear
Loads, Intemational Journal of Solids and Structures, 31, (6), 1994, 763-783.
14.71 Bonanni, D.L., Johnson, E.R. and Starnes, J.H., Jr., Local Buckling and Crippling of Stiffener
Composite Sections, NASA TM-101133, June 1988.
14.72 Bonanni, D.L., Johnson, E.R. and Starnes, J.H., Jr., Local Crippling of Thin-Walled Graphite-
Epoxy Stiffeners, in: Proceedings, AIAAIASME!ASCEIAHS, 29th Structures, Structural Dy-
namics and Materials Conference, Part I, Williamsburg, Va., April 18-20, 1988, 313-323.
14.73 Spier, E. E., Crippling/Column Buckling Analysis and Test of Graphite/Epoxy-Stiffened Panels,
in: Proceedings, A1AA/ ASMEI SAE 16th Structures. Structural Dynamics and Materials Con-
ference, Denver, Colo., May 27-29, 1975, 1-16.
14.74 Spier, E.E., On Experimental versus Theoretical Incipient Buckling of Narrow Graphite/Epoxy
Plates in Compression, in: Proceedings, AIAA/ASMEIASCE!AHS 21st Structures, Structural
Dynamics and Materials Conference, Part I, Seattle, Wash., May 12-14, 1980, 187-193.
14.75 Spier, E.E., Postbuckling Fatigue Behavior of Graphite/Epoxy Stiffeners, in: Proceedings,
AIAAIASME!ASCEIAHS 23rd Structures, Structural Dynamics and Materials Conference, Part
I, New Orleans, La .. May 10-12, 1982. 511-527.
14.76 Spier, E.E. and Klouman, F.L., Empirical Crippling Analysis of Graphite/Epoxy Laminated
Plates, in: Composite Materials: Testing and Design (4th Conference), ASTM STP 617, 1977,
255-271.
14.77 Peery, D.J .. Aircraft Structures, McGraw-Hill, New York, 1950, 382-384.
14.78 Arnold, R.R. and Mayers, J., Buckling, Postbuckling and Crippling of Materially Nonlinear
Laminated Composite Plates, international Journal of Solids and Structures, 20, (9/ 10), 1984,
863-880.
14.79 Pagano, N.J. and Pipes, R.B., The Influence of Stacking Sequence on Laminate Strength, Jour-
nal of Composite Materials, 5, 197!.
14.80 Pagano, N.J. and Pipes, R.B., Some Observations on the Interlaminar Strength of Composite
Laminates, International Journal of Mechanical Sciences, 15, 1973.
14.81 Whitney, J.M. and Kim, R.Y., Effect of Stacking Sequence on the Notched Strength of Lami-
nated Composites, Composite Matenals: Testing and Design (4th Conference), ASTM STP
617, 1977.
14.82 Tyala, S.T. and Johnson, E.R., Failure and Crippling of Graphite-Epoxy Stiffeners Loaded in
Compression, Report No. CCMS-84-07 and VPI-E-84-19, Virginia Polytechnic Institute and
State University, Blacksburg, Va., June 1984.
14.83 Reddy, A.D., Rehfield, W.L., Bruttomesso, R.I. and Krebs, N.E., Local Buckling and Crippling
of Thin-Walled Composite Structures under Axial Compression, in: Proceedings, AIAA I ASMEl
ASCEI AHS 26th Structures, Structural Dynamics and Materials Conference, Part I, Orlando,
Fla., April15-17, 1985, 804-810.
14.84 Reddy, A.D. and Rehfield, L.W., Postbuckling and Crippling Behavior of Graphite-Epoxy Thin-
Walled Airframe Members, in: Proceedings, 41st American Helicopter Society Forum, Fort
Worth, Tex., May 1985, 609-6!5.
14.85 Rehfield, L.W. and Reddy, A.D., Observations on Compressive Local Buckling, Postbuckling
and Crippling of Graphite/Epoxy Airframe Structure, in: Proceedings, AIAA!ASMEIASCE/
AHS 27th Structures, Structural Dynamics and Materials Conference, San Antonio, Tex., May
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

19-21, 1986, 301-305.


14.86 Causbie, S.M. and Lagace, P.A., Buckling and Final Failure of Graphite/PEEK stiffener Sec-
tions, in: Proceedings, AJAA/ ASMEI ASCE! AHS 27th Structures, Structural Dynamics and Ma-
terials Conference, Part I, San Antonio, Tex., May 19-21, 1986, 280-287.
14.87 Wang, C., Pian, T.H.-H., Dugundji, J. and Lagace, P.A., Analytical and Experimental Studies
on the Buckling of Laminated Thin-Walled Structures, in: Proceedings, AIAAIASMEIASCEI
AHS 28th Structures. Structural Dynamics and Materials Conference, Part I, Monterey, Calif.,
April 6-8, 1987. 135-140.
14.88 Alrnroth, B.O., Brogan, F.A. and Stanley, G.M., Structural Analysis of General Shells, 1, 2,
Report No. LMSC-D633873, Applied Mechanics Laboratory, Lockheed Palo Alto Research
Laboratory, Palo Alto, Calif., Dec. 1982.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1237

14.89 Hashin, Z., Failure Criteria for Unidirectional Fiber Composites, Journal of Applied Mechanics,
47, (2), June 1980, 329-334.
14.90 Frostig, Y., Siton, G., Segal, A., Sheinman, I. and Weller, T., Postbuckling Behavior of Lami-
nated Composite Stiffeners and Stiffened Panels under Cyclic Loading", AIAA Journal of Air-
craft, 28, (7), July 1991, 471-480.
14.91 Segal, A., Siton, G. and Weller, T., Durability of Graphite-Epoxy Stiffened Panels under Cyclic
Postbuckling Compression Loading, in: Proceedings, ICCM & ECCM, 6th International Con-
ference on Composite Materials and 2nd European Conference on Composite Materials, F.L.
Matthews, N.C.R. Bushnell, J.H. Hodgkinson and J. Morton, eds., Imperial College, London,
June 20-24, 1987, 5.69-5.78.
14.92 Sheinman, I. and Frostig, Y., Postbuckling Analysis of Stiffened Laminated Panels, Journal of
Applied Mechanics, 55, (3), 1988, 635-640.
14.93 Sheinman, 1., Frostig, Y. and Segal, A., Bifurcation Buckling Analysis of Stiffened Composite
Laminated Panels, in: Buckling of Structures: Theory and Experiments: The Josef Singer An-
niversary Volume, Studies in Applied Mechanics, vol. 19, Elishakoff et a!., eds., Amsterdam,
1988, 355-380.
14.94 Williams, J.G. and Stein. M., Buckling Behavior and Structural Efficiency of Open-Section
Stiffened Composite Compression Panels, AIAA Journal, 14, (11), 1976, 1618-1626.
14.95 Brown, R.T., Computer Programs for Structural Analysis, in: Engineered Materials Handbook,
vol. I, Composites, ASM International, Metals Park, Ohio, 1988, 268-274.
14.96 Spier. E.E. and Anderson, J.A., Test Semi-Empirical Analysis of a Carbon/Epoxy Fabric Stiff-
ened Panel, in: Research in Structures, Structural Dynamics and Materials 1990, NASA Con-
ference Publication 3064, work-in-progress papers from a conference held in Long Beach,
Calif., April 2-4, 1990, 3-35.
14.97 Williams. J.G. and Mikulas, M.M., Jr., Analytical and Experimental Study of Structurally Ef-
ficient Composite Hat-Stiffened Panels Loaded in Axial Compression, in: Proceedings, AIAAI
ASME/ SAE 16th Structures. Structural Dynamics and Materials Conference, Denver, Colo.,
(1), May 27-29, 1975, 1-21.
14.98 Mikulas, M.M., Jr., Bush, H.G. and Rhodes, M.D., Current Langley Research Center Studies
on Buckling and Low-Velocity Impact of Composite Panels, NASA TM X-3377, in: Proceed-
ings of the Third Conference on Fibrous Composites in Flight Vehicle Design, Williamsburg,
Va., Nov. 4-6, 1975, 633-663.
14.99 Stein, M. and Williams, J.G., Buckling and Structural Efficiency of Sandwich-Blade Stiffened
Composite Compression Panels, NASA TP 1269, Sept. 1978.
14.100 Davis, R.C. and Starnes, J.H., Jr., Design Detail Verification Tests for a Lightly Loaded Open-
Corrugation Graphite-Epoxy Cylinder, NASA TP 1981, March 1982.
14.101 Williams J.G., Anderson, M.S., Rhodes, M.D., Starnes, J.H., Jr. and Stroud, J.S., Recent De-
velopments in the Design, Testing, and Impact Damage Tolerance of Stiffened Composite Pan-
els, in: Proceedings of the Fourth Conference on Fibrous Composites in Structural Design, San
Diego, Calif., Nov. 14-17, 1978, E.M. Leone, D.W. Oplinger and J.J. Burke, eds., Plenum
Press, New York, London, 1980, 259-291.
14.102 Starnes, J.H., Jr., Knight, N.F. and Rouse, M., Postbuckling Behavior of Selected Flat Stiffened
Graphite-Epoxy Panels Loaded in Compression, in: Proceedings, AIAAIASMEIASCEIAHS
23rd Structures, Structural Dynamics, and Materials Conference, New Orleans, La., May 10-
12, 1982, 464-477.
14.103 Lopez, O.F., Residual-Strength Tests of L-1 011 Vertical Fin Components after 10 and 20 Years
of Simulated Flight Service, Proceedings, ACEE Composite Structures Technology Conference,
NASA CP 2321, Seattle, Wash., August 13-16, 1984, 1-16.
14.104 Stein, M., Postbuckling of Eccentric Open-Section Stiffened Composite Panels, in: Proceedings,
AIAAI ASMEI ASCEI AHS 29th Structures, Structural Dynamics and Materials Conference, Wil-
liamsburg, Va .. Aprill8-20, 1988, 57-61.
14.105 Sydow, P.O. and Stuart, M.J., Experimental Behavior of Graphite-Epoxy Y-Stiffened Specimens
Loaded in Compression, in: Proceedings, First NASA Advanced Technology Conference, NASA
CP 3104, Part 2, Jan. 1991,953-968.
14.106 Schuette, E., H., Barab, S. and McCracken, H.L., Compression Strength of 24S-T Aluminum-
Alloy Flat Panels with Longitudinally Formed Hat-Section Stiffeners, NACA TN 1157, 1946.
14.107 Hickman, W.A. and Down, N.F., Compressive Strength of 24S-T Aluminum-Alloy Flat Panel
with Longitudinal Formed Hat-Section Stiffeners Having Four Ratios of Stiffener Thickness to
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Skin Thickness, NACA TN 1553, 1948.


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1238 Composite Structures

14.108 Hickman, W.A., and Down, N.F., Data on the Compressive Strength of 75S-T6 Aluminum-
Alloy Flat Panels, Having Small Thin, Widely Spaced, Longitudinal Extruded Z-Section Stiff-
eners, NACA TN 1978, 1949.
14.109 Hickman, W.A. and Down, N.F., Data of 75S-T6 Aluminum-Alloy Flat Panels with Longitu-
dinal Extruded Z-Section Stiffeners, )IACA TN 1824, 1949.
14.110 Down, N.F. and Hickman, W.A., Design Charts for Flat Compression Panels Having Longitu-
dinal Extruded Y-Section Stiffeners and Comparison with Panels Having Formed Z-Section
Stiffeners, NACA TN 1389, 1947.
14.111 Agarwal, B.L., and Davis, R.C., Minimum-Weight Designs for Hat-Stiffened Composite Panels
under Uniaxial Compression, NASA TN D-7779,1974.
14.112 Viswanathan, A.V. and Tamekuni, M.. Elastic Buckling Analysis for Composite Stiffened Panels
and Other Structures Subjected to Biaxial Inplane Loads, NASA CR 2216, 1973.
14.113 Tripp, L.L., Tamekuni, M. and Viswanathan, A.V., User's Manual BUCLASP-2. A Computer
Program for Instability Analysis of Biaxially Loaded Composite Stiffened Panels and Other
Structures, NASA CR 112226, 1973.
14.114 Halstead, D.W., Tripp, L.L., Tamekuni, M. and Baker, L.L., User's Manual BUCLAP-2: A
Computer Program for Instability Analysis of Laminated Long Plates Subjected to Combined
Inplane Loads, NASA CR 132298, 1973.
14.115 Vestergen, P. and Knutsson, L., Theoretical and Experimental Investigation of the Buckling and
Post Buckling Characteristics of Flat Carbon Fibre Reinforced Plastic (CFRP) Panels Subjected
to Compression or Shear Loading, in: Proceedings of the lith Congress of the International
Council of the Aeronautical Sciences (/CAS), J. Singer and R. Staufenbiel, eds., Lisbon, Por-
tugal, Sept. 10-17, 1978, 217-223.
14.116 Romeo, G., Experimental Investigation on Advanced Composite-Stiffened Structures under Uni-
axial Compression and Bending, A/A4. Journal, 24. (11), Nov. 1986, 1823-1830.
14.117 Romeo, G., Analytical and Experimental Results of Advanced Composite Stiffened Panels under
Combined Loads, in: Proceedings of Workshop Composite Design for Space Applications, ES-
TEC, Noordwijk, Oct. 15-18, 1985, ESA SP 243, Feb. 1986, 79-86.
14.118 Falzon, B.G. and Steven, G.P., Buckling and Postbuckling of Stiffened Thin Skinned Carbon
Fibre Composite Panels, Department of Aeronautical Engineering, University of Sydney, Jan.
1993.
14.119 Thuis, H.G.S.J., The Design, Fabrication and Testing of Discrete Stiffened CFRP Compression
Panels with Different Stringer Concepts, NLR-TP-93256-U, June 1993.
14.120 Arendsen, P. and Wiggenraad, J.F.M .. PANOPT User's Manual, NLR CR 91255 L.
14.121 Trentin, C., Alesi. H. and Scott, M.L., Postbuckling Performance of Stiffened Fibre Composite
Panels, Report CRC-ASCP 93001, Cooperative Research Center for Aerospace Structures, Ltd.,
Victoria, Australia, Sept. 1993.
14.122 Scott, M.L. and Rees, D.A., Design and Performance of Postbuckling Composite Structures,
in: Proceedings, Second Pacific International Conference on Aerospace Science and Technology,
Melbourne, Australia, March 20-23, 1995, 647-654.
14.123 Romeo, G. and Frulla, G., Postbuckling Behaviour of Graphite/Epoxy Stiffened Panels with
Initial Imperfections Subjected to Eccentric Biaxial Compression Loading, Journal of Nonlinear
Mechanics, 32, (6), 1997.
14.124 Wiggenraad, J.F.M., The Postbuckling Behavior of Blade-Stiffened Carbon-Epoxy Panels
Loaded in Compression, NLR MP 85019 U Report, National Aerospace Laboratories, NLR,
The Netherlands, Feb. 1985.
14.125 Blass, C. and Wiggenraad, J.F.M., Development and Test Verification of the Ariane 4 Interstage
2/3 in CFRP, in: Proceedings, AIA/ASMEIASCEIAHS 27th Structures, Structural Dynamics
and Materials Conference, Part 1, San Antonio, Tex., May 19-21, 1986, 307-313.
14.126 Frostig, Y., Segal, A., Sheinman, I, and Weller, T., Postbuckling of Flat Stiffened Graphite/
Epoxy Panels Under Cyclic Compression, in: Proceedings, ECCM3, Third European Confer-
ence on Composite Materials, Bordeaux, France, March 20-23, 1989, A.R. Bunsell, P. Lamicq
and A. Massiah, eds., 333-340.
14.127 Bushnell, D. and Bushnell, W.D., Minimum-Weight Design of a Stiffened Panel via PANDA 2
and Evaluation of the Optimised Panel via STAGS, Computers and Structures, 50, (4), 1994,
569-602.
14.128 Bucci, A. and Mercuria, U., CFRP Stiffened-Panels under Compression, in: The Utilization of
Advanced Composites in Military Aircraft, AGARD-785, San Diego, Calif.. Oct. 7-11, 1996,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

12-1-12-14.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1239

14.129 Laananen, D.H. and Renze, S.P., Buckling of Open-Section Bead-Stiffened Composite Panels,
Composite Structures, 25, 1993, 469-476.
14.130 Smith, C.S., Compressive Strength of Transversely Stiffened FRP Panels, in: Aspects of the
Analysis of Plate Structures, D.J. Dawe, R.W. Horsington, A.G. Kamtekar and G. H. Little,
eds .. Clarendon Press, Oxford, 1985, 149-174.
14.131 Davis, R.C., Stress Analysis and Buckling of }-Stiffened Graphite-Epoxy Panel, NASA TP
1607, Feb. 1980.
14.132 Rouse, M., Postbuckling and Failure Characteristics of Stiffened Graphite-Epoxy Shear Webs,
in: Proceedings, AIAAI ASMEI ASCEIAHS 28th Structures, Structural Dynamics and Materials
Conference, Monterey, Calif., April 6-8, 1987, 181-183.
14.133 Whetstone, W.O., SPAR Structural Analysis System Reference Manual-System Level 13A, Vol.
I: Program Execution, NASA CR-158970-1, 1978.
14.134 Thomson, R.S. and Scott, M.L., Testing and Analysis of Thin Stiffened Composite Shear Panels,
in: Proceedings, Second Pacific International Conference on Aerospace Science and Technology,
Melbourne, Australia, March 20-23, 1995, 655-662.
14.135 Rhodes, J.E., Postbuckling and Membrane Structural Capability of Composite Shell Structures,
in: Advances in Composite Materials, Proceedings 3rd International Conference on Composite
Materials, Paris, August 26-29, 1980, A.R. Bunsell, C. Bathias, A. Martrenchar, D. Menkes
and G. Verchery, eds., 1707-1720.
14.136 Sobel, L. and Sharp, D., Comparison of Analytical and Experimental Results for the Postbuck-
ling Behavior of a Stiffened, Flat, Composite Shear Panel, in: Proceedings, 35th AIAA/ ASME
/ASCE/AHSIASC Structures, Structural Dynamics and Materials Conference, Part I, Hilton
Head, SC, April 18-20, 1984, 449-457.
14.137 Antona, E. and Romeo, G., Analytical and Experimental Investigation on Advanced Composite
Wing Box Structures in Bending Including Effects of Initial Imperfections and Crushing Pres-
sure, in: Proceedings of the 15th Congress of the International Council of the Aeronautical
Sciences. vol. 1, London, Sept. 7-12, 1986, 255-261.
14.138 Romeo, G. and Frulla, G., Nonlinear Analysis of Graphite-Epoxy Wing Boxes under Pure
Bending Including Lateral Pressure, AIAA Journal of Aircraft, 32, (6), Nov.-Dec. 1995, 1375-
1381.
14.139 Romeo, G., Frulla, G. and Busto, M., Nonlinear Angle of Twist of Advanced Composite Wing
Boxes under Pure Torsion, AIAA Journal of Aircraft, 31, (6), Nov.-Dec., 1994, 1297-1302.
14.140 Romeo, G. and Frulla, G., Postbuckling Behavior of Graphite-Epoxy Wing Boxes Panels under
Pure Torsion, in: Proceedings of the 20th Congress of the International Council of the Aero-
nautical Sciences, vol. 1, Sorrento, Naples, Sept. 8-13, 1996, 1156-1166.
14.141 Brooks, W.G., The Design and Construction of a Postbuckled Carbon Fibre Wing Box Structure,
in: Proceedings of the 15th Congress of the International Council of the Aeronautical Sciences,
vol. I, London, Sept. 7-12, 1986, 238-243.
14.142 Frey, M., Graf, B. and Messmer, S., Numerical and Experimental Investigation ofPostbuckling
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

of a Composite Box Structure, in: Proceedings of the 19th Congress of the International Council
of the Aeronautical Sciences, vol. 3, Anaheim, Calif., Sept. 18-23, 1994, 2955-2963.
14.143 Dong. S.B., Pister, K.S. and Taylor, R.L., On the Theory of Laminated Anisotropic Shells and
Plates, Journal of the Aerospace Sciences, 29, (8), August 1962, 969-975.
14.144 Cheng, S. and Ho, B.P.C., Stability of Heterogeneous Aeolotropic Cylindrical Shells under
Combined Loading, AIAA Journal, 1, (4), April 1963, 892-898.
14.145 Ho, B.P.C. and Cheng, S., Some Problems in Stability of Heterogeneous Aelotropic Cylindrical
Shells under Combined Loading, AIAA Journal, 1, (7), July 1963, 1603-1607.
14.146 Ugural, A.C. and Cheng, S., Buckling of Composite Cylindrical Shells under Pure Bending,
AIAA Journal, 6, (2), Feb. 1968, 349-354.
14.147 Chehil, D.S. and Cheng, S., Elastic Buckling of Composite Cylindrical Shells under Torsion,
AIAA Journal of Spacecraft, 5, (8), August 1968, 973-978.
14.148 Lei, M.M. and Cheng, S., Buckling of Composite and Homogeneous Isotropic Cylindrical
Shells under Axial and Radial Loading, Journal of Applied Mechanics, Transactions of the
ASME, Dec. 1969, 791-798.
14.149 Geier, B., Energy-Based Task Formulations for Buckling Problems of Laminated Composite
Shells, Zeitschriftfiir Flugwissenschaften und Welfraumforschung, 10. (4), 1986, 215-227.
14.150 Chaudhuri, R.A. Balaraman, K. and Kunukkasseril, V.X., Arbitrarily Laminated, Anisotropic
Cylindrical Shell under Internal Pressure, AIAA Journal. 24, (11), Nov. 1986, 1851-1858.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1240 Composite Structures

14.151 Vinson, R.L. and Sierakowski, R.L., The Behavior of Structures Composed of Composite Ma-
terials, Martinus Nijhoff, Dordrecht, 1986.
14.152 Card, M.F., The Sensitivity of Buckling of Axially Compressed Fiber-Reinforced Cylindrical
Shells to Small Geometric Imperfections, NASA TMX-61914, June 1969.
14.153 Khot, N.S. and Venkaya, V.B., Effect of Fiber Orientation on Initial Postbuckling Behavior and
Imperfection Sensitivity of Composite Cylindrical Shells, Technical Report AFFDL-TR-70-125,
Air Force Flight Dynamics Laboratory, Wright-Patterson Air Force Base, Ohio, Dec. 1970.
14.154 Booton, M. and Tennyson, R.C.. Buckling of Imperfect Anisotropic Circular Cylinders under
Combined Loading. in: Proceedin;?s, AIAA/ ASME 19th Structures, Structural Dynamics and
Materials Conference, Bethesda, Md., April 3-5, 1978, 351-358.
14.155 Sun, G., Optimization of Laminated Cylinders for Buckling, UTIAS Report No. 317, Institute
for Aerospace Studies, University of Toronto, June 1987.
14.156 Simitses, G.J., Buckling of Moderately Thick Laminated Cylindrical Shells: A Review, Com-
posites, Part B 27B, 1996, pp. 581-587.
14.157 Geier, B., Buckling of Fibre Composite Shells: Specific Problems, in: Proceedin;?s of the In-
ternational Conference on Stability ol Structures, ICSS-92, vol. 1, Coimbatore, India, June 7-
9, 1995, S. Rajasekaran and S. Sridharan, eds., Allied Publishers, New Delhi, 57-70.
14.158 Hui, D., Asymmetric Postbuckling of Symmetrically Laminated Cross-Ply Short Cylindrical
Panels Under Compression, Composite Structures, 3, (1), 1985, 81-95.
14.159 Wilkins, D.J., Compression Buckling Tests of Laminated Graphite-Epoxy Curved Panels, AIAA
Journal, 13, (4), April 1975, 465-470.
14.160 Tennyson, R.C., Chan, K.H. and Muggeridge, D.B., The Effect of Axisymmetric Shape Imper-
fections on the Buckling of Laminated Anisotropic Circular Cylinders, Transactions of the
Canadian Aeronautics and Space Institute, 4, (2), Sept. 1971.
14.161 Tennyson, R.C., Muggeridge, D.B., Chan, K.H. and Khot, N.S., Buckling of Fiber-Reinforced
Circular Cylinders under Axial Compression, TR-72-102, Air Force Flight Dynamics Labora-
tory, Wright-Patterson Air Force Base, August 1972.
14.162 Bauld, N.R. and Khot, N.S., A Numerical and Experimental Investigation of the Buckling
Behavior of Composite Panels, Computers and Structures, 15, (4), 1982, 393-403.
14.163 Becker, M.L., Palazotto, A.N. and Khot, N.S., Experimental Investigation of the Instability of
Composite Cylindrical Panels, Experimental Mechanics, 22, Oct. 1982, 372-376.
14.164 Khot, N.S. and Bauld, N.R., Jr., Further Comparison of the Numerical and Experimental Buck-
ling Behaviors of Composite Panels, Computers and Structures, 17, (1), 1983, 61-68.
14.165 Satyamurthy, K., Khot, N.S. and Bauld, N.R., An Automated, Energy-Based, Finite-Difference
of the Elastic Collapse of Rectangular Plates and Panels, Computers and Structures, 11, 1980,
239-249.
14.166 Snell, M.B. and Morely, N.T., The Compression Buckling Behavior of Carbon-Fibre Reinforced
Plastic, in: Proceedings of the Fifth International Conference on Composite Materials, ICCM,
vol. 5, August 1985, 1327-1355.
14.167 Stroud, W.F. and Anderson, M.S., PASCO: Structural Panel Analyses and Sizing Code, Capa-
bility and Analytical Foundations, NASA TM 80181, 1980.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

14.168 Zhang, Y. and Matthews, F.L., Initial Buckling of Curved Panels of Generally Layered Com-
posite Materials, Composite Structures, 1, (!),Jan. 1980, 3-30.
14.169 Knight, N.F., Starnes, J.H., Jr. and Waters, W.A., Jr., Postbuckling Behavior of Selected Graph-
ite-Epoxy Cylindrical Panels Loaded in Axial Compression, Paper No. 86-0881-CP, in: Pro-
ceedin;?s, AJAA/ ASMEI ASCEI AHS 27th Structures, Structural Dynamics, and Materials Con-
ference, San Antonio, Tex., May 19-21, 1986.
14.170 Kobayashi, S., Sumihara, K. and Koyama, K., Compressive Buckling Strength of Graphite-
Epoxy Laminated Curved Panels, in: Proceedin[;s, 15th Congress of the International Council
of the Aeronautical Sciences, JCAS, London, Sept. 7-12, 1986, 244-254.
14.171 Jensen, D.W. and Hipp, P.A., Compressive Testing of Filament-Wound Cylinders, Composites,
in: Proceedin;?S of the 8th International Conference on Composite Materials (TCCM/8), Hon-
olulu, Hawaii, July 15-19, 1991, 35-F-1-35-F-9.
14.172 Card, M.F. and Peterson, J.P., On the Instability of Orthotropic Cylinders, NASA TN D-1510,
Dec. 1962, 297-308.
14.173 Card, M.F., Experiments to Determine the Strength of Filament Wound Cylinders Loaded in
Axial Compression, NASA TN D-3522, August 1966.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1241

14.174 Tsai, J., Feldman, A. and Stang, D.A., The Buckling Strength of Filament Wound Cylinders
under Axial Compression, NASA CR-266, July 1965.
14.175 Hahn, H.T., Jensen, D.W., Claus, S.J. and Hipp, P.A., Structural Design Criteria for Filament
Wound Composite Shells, NASA CR-195125, Jan. 1994.
14.176 Stein, M. and Mayers, J., Compressive Buckling of Simply Supported Curved Plates and Cyl-
inders of Sandwich Construction, NACA TN 2601, 1952.
14.177 Card, M.F., Experiments to Determine Elastic Moduli for Filament-Wound Cylinders, NASA
TN D-3110, 1965.
14.178 Puck, A. and Ri.iegg, C., Experimentelle Untersuchungen zur Beulstabilitat von gewickelten
GFK-Zylindern unter axialer Drukbeslastung, Kunststoffe, 64, (12), 1974,3-11.
14.179 Ruegg, C., Beulstabilitat von gewickelten GFK-Zylindern unter axialer Last, SIA-Tagung Kun-
stoffe im Bauingenieurwesen, Zurich, Oct. 17-18, 1975, 342-372.
14.180 Ravenhall, R., Stiffness and Buckling in Filament-Wound Motors, AIM Journal of Spacecraft,
1, (3), May 1964, 260-263.
14.181 Sargent, T.V., Compressive Testing of Eighteen-Inch I.D. Spirally Cylinders, Hercules Powder
Co., Report MTI-123, Feb. 1960.
14.182 Garde, A.N. and Lakkad, S.C., Experimental Studies on the Instability of Axially Loaded Cy-
lindrical FRP Shells, in: Proceedin[is, 5th International Congress on Experimental Mechanics,
Montreal, June 10-15, 1984, 92-96.
14.183 Baker, E.H., Kovalevsky, L. and Rish, FL., in Structural Analysis of Shells, McGraw-Hill, New
York, 1972, chap. 11, 293-296.
14.184 Tennyson, R.C. and Muggeridge, D.B., Buckling of Laminated Anisotropic Imperfect Circular
Cylinders under Axial Compression, Journal of Spacecraft, 10, (2), Feb. 1973, 143-148.
14.185 Tennyson, R.C., Muggeridge, D.B. and Caswell, R.D., Buckling of Circular Cylindrical Shells
Having Axisymmetric Imperfection Distributions, AIM Journal, 9, (5), May 1971. 924-930.
14.186 Geier, B., Klein, H. and Zimmermann, R., Buckling Tests with Axially Compressed Unstiffened
Cylindrical Shells Made from CFRP, in:BucklinR of Shell Structures, on Land, in the Sea and
in the Air, Proceedings International Conference, J.F. Jullien. ed., INSA Lyon, September
17-19, Elsevier Applied Science, London, New York, 1991, 498-507.
14.187 Geier, B., Klein, H. and Zimmermann, R., Experiments on Buckling of CFRP Cylindrical Shells
under Non-Uniform Axial Load, in: Proceedings of the International Conference on Composite
Engineering (ICCEI 1), D. Hui, ed., August 28-31, 1991, 1043-1044.
14.188 Zimmermann, R., Buckling Research for Imperfection Tolerant Fiber Composite Structures, in:
Proceedings, Conference on Spacecraft Structures, Materials and Mechanical Testing, Noord-
wijk, The Netherlands, March 27-29, 1996 (ESA SP-3856, June 1996), 411-416.
14.189 Giavotto, V., Poggi, C., Castano, D., Guzzetti, D. and Fezzani, M., Buckling Behavior of Com-
posite Shells under Combined Loading, in: Proceedings, 17th European Rotorcraft Forum,
Berlin, Sept. 1991, 84-1-84-13.
14.190 Poggi, C., Taliercio, A. and Capsoni, A., Fiber Orientation Effects on the Buckling Behavior
of Imperfect Composite Cylinders, in: Proceedings, International Colloquium on Buckling of
Shell Structures on Land, in the Sea and in the Air, Lyon, France, Sept. 17-19, 1991, J.F.
Jullien, ed., 114-123.
14.191 Bisagni, C., Buckling and Postbuckling Behavior of Composite Cylindrical Shells, in: Pro-
ceedings, !CAS 1996, 20th Congress of the International Council of the Aeronautical Sciences,
Sorrento, Napels, Italy. Sept. 8-13, 1996,1, 1148-1155.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

14.192 Fuchs, J.P., Hyer, M.W. and Starnes, J.H., Jr., Numerical and Experimental Investigation of the
Bending Response of Thin-Walled Composite Cylinders, Report CCMS-93-19, VPI-E-93-11,
Center for Composite Materials and Structures, College of Engineering, Virginia Polytechnic
Institute and State University, Sept. 1993.
14.193 Wolf, K. and Kossira, H., The Buckling and Postbuckling Behavior of Curved CFRP Laminated
Shear Panels, in: Proceedings, !CAS 88, 16th Congress of the International Council of the
Aeronautical Sciences, Jerusalem, Israel, August 28-Sept. 2, 1988, 1, 920-930.
14.194 Wolf, K. and Kossira, H., An Efficient Test Method for the Experimental Investigation of the
Postbnckling Behavior of Curved Composite Shear Panels, in: Proceedings, ECCM-CTS, Eur-
opean Conference on Composite Testing and Standardisation, Amsterdam, September 8-10,
1992.
14.195 Kuenzi, E.W., Stability of a Few Curved Panels Subjected to Shear, Forest Products Laboratory,
Report No. 1571, 1947.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1242 Composite Structures

14.196 Bert, C.W., Crisman, W.C. and Nordby, G.M., Fabrication and Full-Scale Structural Evaluation
of Glass-Fiber Reinforced Plastic Shells, Journal of Aircraft, 5, 1968, 27-33.
14.197 Wilkins, D.J. and Olson, F., Shear Buckling of Advanced Composite Curved Panels, Experi-
mental Mechanics, 14. 1974, 326-330.
14.198 Marshall, l.H., Rhodes, J. and Banks, W.M., Experimental Snap-Buckling Behavior of Thin
GRP Curved Panels under Lateral Loading, Composites. 8, (4), April 1977, 81-86.
14.199 Marshall, I.H., Rhodes, J. and Banks, W.M., The Non-Linear Behaviour of Thin Orthotropic
Curved Panels under Lateral Loading, Journal of Mechanical Engineering Science, 19, Feb.
1977, 30-37.
14.200 Ward, H.S. and McCleskey, S.F., Buckling of Laminated Shells Including Transverse Shear
Flexibility, in: Proceedings, AIAAIASME/ASCEIAHS, 28th Structures, Structural Dynamics
and Materials Conference, Part L Monterey, Calif., April 6-8, 1987, 141-147.
14.201 Abu-Farsakh, G.A.F.R. and Lusher, J.K., Buckling of Glass-Reinforced Plastic Cylindrical
Shells under Combined Axial Compression and External Pressure, AIAA Journal, 23, (12), Dec.
1985, 1946-1951.
14.202 Bert, C.W., Crisman, W.C. and Nordby, G.M., Buckling of Cylindrical and Conical Sandwich
Shells with Orthotropic Facings, AIAA Journal, 7, (2), Feb. 1969, 250-257.
14.203 Wilkins, D.J. and Love, T.S., Combined Compression-Torsion Buckling Tests of Laminated
Composite Cylindrical Shells, Journal of Aircraft, 12, (11), Nov. 1975, 885-889.
14.204 Herakovich, C.T. and Johnson, E.R., Buckling of Composite Cylinders under Combined Com-
pression and Torsion-Theoretical/Experimental Correlation", in: Test Methods and Design Al-
lowables for Fibrous Composites, ASTM STP 743, C.C. Chamis, ed., American Society for
Testing and Materials, West Conshohocken, Pa., 1981, 341-360.
14.205 Wang, C.T., Vaccaro, J.R. and Desanto, D.E., Buckling of Sandwich Cylinders under Combined
Compression, Torsion. and Bending Loads, Journal ofApplied Mechanics, 22, 1955, 324-328.
14.206 Agarwal, B.L., Postbuckling Behavior of Composite-Stiffened Curved Panels Loaded in Com-
pression, Experimental Mechanics, 2::, 1982, 231-236.
14.207 Knight, N.F., Jr. And Starnes, J.H., Jr., Postbuckling Behavior of Selected Curved Stitiened
Graphite-Epoxy Panels Loaded in Axial Compression, AIAA Journal, 26, (3), March 1988,
344-352. Also in: Proceedings AIAAI ASMEIASCEIAHS 26th Structures, Structural Dynamics
and Materials Conference, Orlando, Fla., April 15-17, 1985, Paper No. 85-0768-CP.
14.208 McGowan, D.M., Young, R.D., Swanson, G.D. and Waters, W.A., Compression Tests and Non-
linear Analysis of a Stringer- and Frame-Stiffened Graphite-Epoxy Fuselage Crown Panel, Fifth
NASA/Dod Advanced Composites Technology Conference, Seattle, Wash., August 22-25,
1994, Paper No. A94-33140.
14.209 Sobel, L.H. and Agarwal, B.L., Buckling of Eccentrically Stringer-Stiffened Cylindrical Panels
under Axial Compression, Journal of Computers and Structures, 6, (3), June 1976, 193-198.
14.210 Anderson, M.S. and Stroud, W.J., A General Panel Sizing Computer Code and Its Application
to Composite Structural Panels, AIAA Journal, 17, (8), August 1979, 892-897.
14.211 Dickson, J.N. and Biggers, S.B., POSTOP: Postbuckled Open Stiffener Optimum Panels-
Theory and Capability, NASA CR-172259, 1984.
14.212 Biggers, S.B. and Dickson, J.N., POSTOP: Postbuckled Open Stiffener Optimum Panels-
User's Manual, NASA CR-172260, !984.
14.213 Davis, R.C., Buckling Test of a 3-Meter-Diameter Corrugated Graphite-Epoxy Ring-Stiffened
Cylinder, NASA TP-2032, July 1982.
14.214 Johnson, R., Jr., Reck, R.J. and Davis, R.C., Design and Fabrication of a Large Graphite-Epoxy
Cylindrical Shell, in: Proceedings, AIAA I ASME 19th Structures, Structural Dynamics and Ma-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

terials Conference, Bethesda, Md., April 3-5, 1978, 300-310.


14.215 Penton, A.P., Johnson, R., Jr. And Freeman, V.L., Fabrication of Composite Shell Structure for
Advanced Space Transportation, in: Selective Applications of Materials for Products and En-
ergy, vol. 23 of National SAMPE Symposium and Exhibition, Society for the Advancement of
Material and Process Engineering. Covina, Calif., 1978, 137-149.
14.216 Johnson, R., Jr., Design and Fabrication of a Ring-Stiffened Graphite-Epoxy Corrugated Cylin-
drical Shell, NASA CR-3026, 1978.

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
15
Nondestructive Buckling
Tests

15.1 Nondestructive Methods for Buckling Tests


One of the disadvantages of the common buckling test is its destructive or terminal nature.
After the experimental team has spent many weeks and days preparing the test, it is over in
a very short time. The loaded structure buckles and fails, usually with large plastic defor-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

mations. Though with modern automated data acquisition systems most of the measured
information can be retrieved, the same test cannot be repeated! Therefore the use of a buck-
ling experiment as a proof test is limited to significantly imperfection-insensitive structures,
a very small group in practice, or to special experiments designed specifically for repeated
buckling, like certain combined loading tests. Thus the desire and need for nondestructive
buckling test method evolved.
Nondestructive methods for structural testing have been actively employed and widely
used for decades (see for example [15.1]-[ 15.5]). They employ indirect measurements that
do not damage the test objects to detect flaws and measure performance properties of struc-
tures. Each NDT method requires some form of probing media and some detection system
and can accordingly be grouped into categories (such as radiography, ultrasonics, electro-
magnetic induction, magnetic-field tests, liquid-penetration, thermal, acoustic emission and
others). The applications of NDT methods have been primarily aimed at detection of point
responses, such as crack initiation and propagation, or in the case of composite structures,
the detection of hidden dclaminations. Overall, or whole-field responses have only been dealt
with more recently. Buckling has therefore scarcely been considered.
Nondestructive methods for buckling appeared first for the simplest case, that of the simple
column. The earliest NDT methods employed for columns were the Southwell method, dis-
cussed in Chapter 4, Subsection 4.6.4 (where more recent examples, like [4.44], were also
mentioned); and the vibration correlation methods, which originated from the similarities
between the bucking and vibration behavior of columns.
The similarities between buckling and vibration behavior of columns and other structures,
as well as the resulting correlations methods, will be discussed in the next section. Before
we embark on this discussion, we should, however, point out that all nondestructive methods
for buckling can be classified according to their character into two classes: static and dynamic
methods; and according to their approach into two groups: (1) those for determination of
the actual boundary conditions leading to better calculations of the buckling loads, and (2)
those for direct determination of the buckling loads.

Buckling Experiments: Experimental Methods in Buckling of Thin-Walled Structures: Shells, Built-Up Structures, Composites
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
and Additional Topics – Volume 2. J. Singer, J. Arbocz and T. Weller
No reproduction or networking permitted without license from IHS
Copyright © 2002 John Wiley & Sons, Inc.
Not for Resale, 02/13/2019 01:32:58 MST
1244 Nondestructive Buckling Tests

15.2 Vibration Correlation Techniques {VCT)


15.2.1 Correlation between Vibration and Buckling
As pointed out in Chapter 2, Section 2.2.1, the free harmonic vibrations of an axially com-
pressed column are governed by the differential equation
Elw,,,u + Pw," + pAW, 11 = 0 (15.1)
where w is the lateral deflection of the column, A its cross-section and p the mass per unit
volume, and the appropriate boundary conditions. Equation (15.1) is identical to Eq. (2.266)
of Chapter 2, designated here as Eq. (15.1A),
(15.1A)
where F =(PIE!) and p =(pAlE!). One may also reiterate that this differential Eq. (15.1)
could be obtained either by application of d' Alembert's principle to the static beam column
equation (by replacing the distributed lateral load by the inertia load) or by application of
Hamilton's principle (variation of the total potential and kinetic energy), which yields also
the appropriate combinations of boundary conditions.
For harmonic vibrations,
w =' W(x) sin wt (15.2)
and substitution of Eq. (15.2) in Eq. (15.1), or using separation of variables as in Chapter
2, yields
2
EIW,"'' + PW," -w pAW = 0 (15.3)
or
(15.3A)
where again Eq. (15.3A) is identical to Eq. (2.269) of Chapter 2.
For simple supports and constant El, the boundary conditions are, as in Eq. (2.2) of
Chapter 2 or Eq. (4.3) of Chapter 4,
W(O) = W,,.JO) = W(L) = W,"(L) = 0
which are satisfied by
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

W = C sin( 1TXIL) (15.4)


Substitution of Eq. (15.4) into Eq. (15.3) yields the frequency squared
w 2 = (1T 2 /pAU)[(p 2 EIIU)- P] (15.5)
or
(15.5A)
2
where P = (1T Ell U) is the Euler load of the simply supported column.
Equation (15.5) is the well-known linear relationship between the frequency squared w 2
and the axial compressive load P for a simply supported column, shown in Figures 15.1 and
15.2. For P = 0, the frequency becomes the natural frequency for free vibrations of a w
simply supported beam. AsP increases, the frequency (here the fundamental one) decreases.
When P = PD the buckling load, the frequency falls to zero, which characterizes buckling.
The same relation applies to the nth mode of vibration and buckling of the simply sup-
ported column,
W, = C, sin(nm:/ L) (15.4A)
for which the frequency squared is
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Vibration Correlation Techniques (VCT) 1245

I 2
i
' I'

'I"J
10
i
!
0.8
I

!--------

0.4
~I
0.2 ~
~j
0
"-'i. i
0 0.2 0.4 0.6 0.6 L2
'"
(:J
Figure 15.1 Variation of frequency squared with axial compression for a uniform column with clamped
ends (from [15.14]). Upper and lower bounds of frequency squared ratios (w/ w0 ) 2 , ob-
tained by energy methods, are plotted versus the end load ratio (PIP,) in the figure,
as three straight lines forming a triangle, which bounds the true curve of the (w/ w0 f-(P
I PtJ relation. The extremely small deviation from linearity of this curve is apparent

(15.5B)
where P" = (n 2 1r 2 El I U) is the buckling load corresponding to the nth mode. Equations
(15.5) or (l5.5B) can be rewritten as
(w,,w"of = 1 -(PIP") (15.6)
to emphasize the linear relationship (w110 here is the nth frequency of the unloaded column).

Figure 15.2 Vibration con-elation techniques (VCT) for a column: the indirect method for definition
of the end fixity and the direct method for determination of the buckling load. x = ex-
perimental points --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1246 Nondestructive Buckling Tests

The relationship between frequency squared and compressive load is exactly linear only
in the case of simple supports, where the vibration mode is identical to the buckling mode.
For other boundary conditions the straight line is replaced by shallow curves that hardly
deviate from the straight line, as was shown for example by Lurie for clamped columns
using energy methods (see Figure 15.1 and [15.14]). Note that in the figure w0 designates
the fundamental frequency of the unloaded column.
The concept of relating vibration characteristics to buckling loads was considered at the
beginning of the 20th century, when Sommerfeld [15.6] tested a clamped-free column, ver-
tically mounted, with a variable mass attached to the free end. He observed that the natural
frequency of the column decreased as the mass increased, approaching zero as the tip mass
reached the weight needed to buckle the column. In the decades that followed, the buckling-
vibration relation was studied theoretically by a number of researchers (see for example
[15.7]-[15.12]), but experimental investigations appeared only at the middle of the century
(for example [15.13]-[15.17]).
After the early theoretical studies [ 15.7] and [15.8] had established the linear relationship
between the frequency squared and the axial compressive load for a simply supported col-
umn, the investigators extended their analyses to columns with other boundary conditions
and to other structural elements (like [15.10], [15.11] or [15.14]. The comprehensive theo-
retical studies of Massonnet, who analyzed the behavior of uniform beams, plates and cy-
lindrical as well as spherical shells with various boundary conditions [15.10] and [15.11],
laid the foundation for future work. He showed that in all cases studied, the buckling load
and the square of the natural frequency were very nearly linearly related, and that this linear
relationship became exact if the mode of free vibrations was identical to the buckling mode.
Lurie [15.14] amplified this for columns, as mentioned earlier in this subsection (see Figure
15.1), and studied rigid-joint trusses and plates. He and others also carried out experiments
to evaluate the theories and establish practical nondestructive vibration correlation techniques
(see for example [15.14]-[15.17]). These techniques and their more recent developments are
discussed in the following subsections.

15.2.2 Vibration Correlation Techniques (VCT) for Determination of


Boundary Conditions and Buckling Loads in Columns
For columns, the primary goal of the vibration correlation techniques has usually been the
direct nondestructive determination of the buckling loads; but the definition of the actual
boundary conditions, or end fixity, has also been considered a significant task. Indeed, the
early studies (such as [15.9], [15.12], [15.18] and [15.19]) focused on this task of definition
of boundary conditions, which has retained its importance today, especially for in situ mea-
surements.
For columns, the deviations from linearity in the frequency squared-compressive load
relations are small (as shown in Figure 15.1), since their fundamental vibration modes are
similar to the corresponding buckling modes for all boundary conditions. This was recon-
firmed again by Burgreen [ 15.19], who showed that for equal end fixities the maximum error
in assuming linearity was about 2 percent. Therefore, one can neglect these small deviations
and consider the relations straight lines. Then the procedure for determination of the end
fixity of columns is rather simple.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

The end fixity can be represented by a rotational elastic restraint K (see Figure 15.2),
whose value defines the relative position of its frequency squared-load line in relation to
those of the simply supported column (K = 0) or the clamped one (K __, oo). Since the
influence of the elastic restraint on the frequency is practically identical to its effect on
buckling of the column, the relative position of wl to w~, and w~1 is also identical to that of
the buckling load in relation to P,, " and P, '·"·
Thus, in the experiment one measures the fundamental frequency of the actual column at
increasing loads, up to about half the roughly guessed end fixity, and draws the best-fit line
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Vibration Correlation Techniques (VCT) 1247

between the w2 points (see Figure 15.2). Measurement in the presence of some axial load
ensures that the column has settled and that the measured end fixity will represent that
effective in the real service conditions of the loaded column.
The position of the w 2 - P line yields an appropriate elastic restraint K. The corresponding
buckling load can then be calculated. In this manner VCT is employed for determination of
the boundary conditions-the indirect method. One can, however, extrapolate the best-fit line
drawn through the experimental points (as shown by the broken line in Figure 15.2) to the
abscissa to obtain P,,,k directly-the direct VCT method.
Probably the earliest vibration correlation experiments on elastically restrained columns
were carried out at the Guggenheim Aeronautical Laboratories of the California Institute of
Technology by Chu [15.13] and a couple of years later by Lurie [15.14]. Both test programs
employed rigid-joint frames subjected to end loads (see Figure 15.3) to provide two axially
compressed columns in each test. The frames were made of 5 I 16 X 1 in. cold-rolled steel
strip, brazed together at the four corners, and measured approximately 12 X 20 in. (305 X
508 mm). The axial load was applied through knife edges resting in hardened-steel seats.
Whereas in the earlier tests the free vibrations were measured, in the later ones an oscil-
lator was built, which imposed an exciting force on the frame. By varying the frequency of
the oscillator so as to develop resonance, the natural frequency could be found at each load
increment. This is essentially the technique used today in most VCT experiments. Strain
gages were attached to both vertical members of the frame to determine whether the vibration
modes were symmetric or unsymmetric. Measurements were recorded automatically on a
two-channel recording oscillograph.
Lurie first tried to repeat Chu's results and indeed corroborated that the first frame vibrated
in the unsymmetric mode. The frame remained stable beyond the theoretical symmetrical
buckling load of 7330 lb, and frequency measurements continued with increasing load. The
load was increased further until the frame buckled in the unsymmetrical mode. The mode
of buckling was therefore of the same form as the mode of vibrations, validating the VCT
approach. The variation of frequency squared with load is shown in Figure 15.4 as curve 1.
Extrapolation of this curve to the abscissa (where w = 0) gave a buckling load of 11,150
lb, as compared to the theoretical unsymmetric critical load of 11,360 lb for this frame, i.e.
the VCT prediction was within 2 percent of the theoretical buckling load. The experimental
buckling load was not recorded in this test.
A test on a second frame of the same dimensions yielded similar results, which are plotted
as curve 2 in Figure 15.4. The actual experimental buckling load was measured here as
10,250 lb, compared with the VCT extrapolated value of 11,450 lb and the theoretical value
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

of 11,360 lb. Again the VCT prediction was within I percent of the theoretical buckling
load, but nearly II percent above the experimental buckling load.
A third frame of similar dimensions was tested with eccentric knife-edge seats (which had
been moved 0.2 in. towards the centerline of the frame). The purpose of this eccentric loading

Figure 15.3 Rectangular frames used in GALCIT VCT experiments on elastically restrained columns:
symmetrical and unsymmetrical modes (from [15.14])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1248 Nondestructive Buckling Tests

wz
eoor--~~----+-~~~

2,000 6,000 8,000 10.000 12,000


p

Figure 15.4 GALCIT VCT experiments on rectangular steel frames: some test results (curves I, 2, and
5 from [15.14])

was to ensure symmetric buckling. The vibrations were, however, still unsymmetrical,
whereas buckling was apparently symmetric. The measured buckling load of 5500 lb was
rather low, possibly because of the load eccentricity.
To study the discrepancies with theory, the experimental boundary conditions were re-
examined. The frame was supported on knife edges of square cross-section that were firmly
held in 90° V-blocks, which prevented any lateral movement between the two supports. The
load was applied by similarly held knife edges. To study the influence of free lateral move-
ment between supports, such freedom was obtained by replacing the right-hand bottom and
top 90° V-block supports by hardened-steel knife-edge seats, which permitted lateral move-
ment, and leaving the left-hand supports as before.
A fourth frame, of the same dimension as the previous three, was now tested with these
modified boundary conditions. The symmetric mode now became the stable form for all
loads. The frequency squared versus load for this frame is plotted as curve 5 in Figure 15.4,

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
which is again practically linear. Extrapolation to the abscissa yielded the VCT predicted
buckling load of 7330 lb, which corresponded here exactly to the theoretical critical load
but exceeded the experimental buckling load of 6950 lb by about 5 percent. Lurie's figure
included additional partial results, which were inconsequential and are therefore omitted in
Figure 15.4.
Lurie's experiments thus demonstrated the practical feasibility of VCT for prediction of
buckling loads. He also pointed out two necessary preconditions for the use of VCT in
columns and frames: (1) the column or members of the frame should be long enough to
ensure elastic instability, (2) the boundary conditions should not vary with loading to ensure
identity of those measured by the vibrations and those acting at buckling.
A year later, Johnson and Goldhammer at the U.S. Navy David Taylor Model Basin
employed the vibration correlation technique to determine the buckling load of a built-up
column, actually a stiffened panel, supported only at the top and bottom, with five T-stiffeners
welded to the plate [15.16]. Their main purpose was to determine the degree of restraint
exerted on the ends of their panel-column by the testing machine and to assess the influence
of various modifications in the panel on the effective length of the column. For the same
panel to be reused, a nondestructive method was essential, and VCT proved to be very
suitable.
The 75-in. long stiffened panel (to whose loaded ends heavy bearing plates were welded)
was loaded axially in a hydraulic universal testing machine, and vibrations at the fundamental
frequency were excited and measured by an electromagnetic vibration generator in conjunc-
tion with an electronic power supply. The plot of applied load versus the square of the
natural frequency, assumed to be a straight line, is shown in Figure 15.5. The least-squares
fit straight line through the 10 measured points was extended to zero frequency to obtain
the critical load. A simple formula was also derived to facilitate prediction of the buckling
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Vibration Correlation Techniques (VCT) 1249

100
I
'
! I

I
I
I
j
I

J •
~~- - ~

-
-t±
80 ~·
I
l
~ i I
r----- l I I
60

40
I ""-..
I"-- -t- I
I

1"-- l
20 I [b--. I
I t I f1t.... '
~T
J I 1'-- I
2 00 400 600 800 1000
2
(FREQUENCY fIN (CPS )

Figure 15.5 U.S. Navy David Taylor Model Basin VCT tests on a stiffened panel-column: compressive
load versus frequency squared (from [15.16]). Moment of inertia of cross-section 0.4667
in.•

load by this process from the measured points, without plotting them. The sample calculation
yielded P" = 89,400 lb, and hence an effective column length of 39.3 in. Since the buckling
load was obtained nondestructively, the panel could be modified (actually stiffened) and
retested to yield in the same manner the effective length of the altered configurations. Thus,
the practicality of VCT for columns was verified, and the authors also recommended its in
situ application for assessment of the actual effective lengths.
The use of vibration correlation techniques for prediction of buckling loads and determi-
nation of the end fixity of columns (including built-up ones) has since become an accepted
practice. An example is [15.20], discussing the experimental procedure employed at the
Northrop Corporation in California in the late sixties. A Rene 41 (nickel alloy) brazed
honeycomb column was tested, and it was vibrated by a shaker attached to the center of the
column by means of vacuum pad. The weight of the moving part of the shaker was less
than 5 percent of the weight of the column and hence hardly affected the mass of the
vibrating column. The necessary attention was paid to the important issue of mass addition
and its influence on the natural frequency.
Another example were the careful experiments on the buckling of axially compressed
aluminum and composite columns and plates, carried out by Chailleux et al. at the Ecole
des Mines (ENSTA) in Paris in the mid-seventies [8.82]. The static part of their study, and
in particular the application of the Southwell plot to plates, was already discussed in Chapter
8, Section 8.3.4. They also conducted, however, tests by a dynamical method, the vibration
correlation technique, on the same columns and plates. Figures 15.6(a) and (b) show sche-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

matically their VCT test setup and measurement systems for columns and plates, respectively.
For columns, the test setup (Figure 15.6a) was relatively simple. The bending curvature
K was measured near the middle of the column by the customary pair of strain gages (2).
The flexural vibrations of the column were controlled by the electrodynamic exciter (5),
supplied by a low frequency sinusoidal current generator (4). The exciter was suspended by
four springs, and its moving part was very light to minimize additions of mass to the vibrating
column. The amplitude of the vibrations was maintained at a low level to avoid nonlinear
effects in the vibrations. At each load step the first resonance frequency was found with the
aid of Lissajous figures on the oscilloscope (9).
To reiterate, the phase difference between the exciting force (i.e. the current which supplies
the exciter) and the deflection (i.e. the alternating part of the signal measured by the strain
gages during the vibration) yields at resonance an ellipse with horizontal and vertical axes
on the oscilloscope-called a Lissajous figure. The Ecole des Mines researchers asserted
that the Lissajous figure procedure was more precise (yielding a precision better than 0.5
percent) than the often-used search for maximum amplitude of vibrations, though opinions
of later researchers differ on that point.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1250 Nondestructive Buckling Tests

!~------~p
Qq
(7) 1----- ----P

~-c:::-.;:-=;--~l
@ ~K
CD lol - ,2) 0
~-~--~
I

CD
.,z,
.
I
~~- (~)u
icDl I
(!_'

---~f
(II) I L" .
~I
(a) (b) \@~1

Figure 15.6 Ecole des Mines (ENSTA), Paris, VCT experiments on columns and plates: schematic
diagrams of test setup and measurement systems (from [8.82]): (a) for columns, (b) for
plates. In both diagrams: (1) specimen, (2) strain gages, (3) amplifier, (4) low-frequency
sinusoidal current generator, (5) electrodynamic exciter, (6) displacement pickup, (7) com-
pensation, (8) galvanometric light recorder, (9) oscilloscope, (10) voltmeter, (II) frequency
meter. Note that numbers (6)-(8) apply only to (b)

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
The dynamical diagram, or load versus frequency squared plot, for a typical column (Col-
umn no. 4: length 200 mm, width 10.35 mm and depth 2.23 mm, made of resin epoxy
unidirectionally reinforced by three layers of boron fibers, six layers of glass fibers and again
three layers of boron) is shown in Figure 15.7. The diagram presents three zones. In zone I
the load on the column is still small and the end clearances have not yet completely dis-
appeared. This zone is always very narrow. In zone II the measured points are positioned
along a straight line, which can be extrapolated to the ordinate to yield the buckling load
P". In zone III, when the loads approach the buckling load the measured points diverge
rapidly from the straight line, due to unsymmetrical vibrations induced by the accumulated
bending of the specimen.
The precision in the extrapolated P,, obtained by VCT was found to depend on the size
of the straight line zone II in the plot. When zone II extended to about three-quarters of the
VCT plot, the P" obtained from it was very close to that obtained by the static Southwell
method and to the theoretical prediction, within a few percent. When the straight line zone
was narrower, the dynamically obtained P,., tended sometimes to be significantly larger than
the theoretical prediction and that obtained by the Southwell plot, (up to about 20 percent
in a few cases), indicating that VCT might then be unreliable, though the dynamical diagram
exhibited a narrow linear portion.

Pdyn' 45 7 kg

10

2 2
f I Hz l
Figure 15.7 Ecole des Mines (ENSTA), Paris, VCT experiments on columns: dynamical diagram, load
versus frequency squared, for a typical laminated composite column (epoxy resin with
unidirectional boron and glass fibers), with two edges hinged (from [8.82])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Vibration Correlation Techniques (VCT) 1251

Eight columns of rectangular cross-section and a span-to-depth ratio of about 100 were
tested. Five of these were made of epoxy resin, reinforced with unidirectional boron and
glass fibers, and three were made of aluminum reinforced by unidirectional or angle-plied
boron fibers. All columns were tested with three types of boundary conditions:
1. H.H.: two simply supported (hinged) ends on knife edges
2. H.C.: one end simply supported and the other clamped
3. C.C.: two clamped edges
The VCT predicted buckling loads were very close to (within a few percent of) the theoretical
and Southwell plot ones for the H.H. and H.C. edges, but the differences for some of the
C.C. boundary conditions were larger.
In general, it was shown that with sufficient care the VCT yielded reliable results for
columns. The corresponding Ecole des Mines experiments on plates are discussed in the
next subsection.
Before we leave columns, we should point out that for their case VCT is indeed rather
simple, as is evident in the student experiments carried out routinely at the Technion in
Haifa. The usual column buckling experiment in the undergraduate student laboratory was
extended to include vibrations of the loaded column by attachment of an exciter to the center
of the columns as well as an accelerometer on the opposite side for measurement of the
frequencies (see Figure 15.8a). The specimens were steel columns of rectangular cross-

(a)

left= 380 mm

.6
_[J_)2
( wo .4 A

.2

(b)
Figure 15.8 Student experiments employing VCT on columns at Technion-Israel Institute of Tech-
nology, Haifa: (a) schematic diagram of test setup: (I) column, (2) end fitting, (3) load
cell, (4) upper platen of testing machine, (5) bridge circuit, (6) electromagnetic shaker,
(7) accelerometer, (8) amplifier, (9) oscilloscope, (10) oscillator, (11) frequency counter.
(12) lower platen of testing machine; (b) typical frequency ratio squared (w/w0 f) versus
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

axial load P plot for a simply supported column of slenderness ratio (L/ p) = 263 mm
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1252 Nondestructive Buckling Tests

section, held in roller-bearing end fittings (see Figure 6.16) for simple supports (SS) or
clamped (C) in the same end fittings with the rotating axis fixed by set screws. In a typical
laboratory session about 20 columns of four lengths, for which (LIp) varied from 159.4 to
263.3 and therefore elastic buckling was ensured, were tested with three combinations of
boundary conditions: SS-SS, C-C and C-SS.
In each test the buckling load was predicted by VCT and by a Southwell plot, the points
being obtained by increasing the load in steps to about 60-80 percent of the calculated
(Euler) buckling load for the nominal boundary conditions.
In the dynamical part of the test, application of VCT, the column was excited by the
shaker (no. 6 in Figure 15.8a) with increasing frequencies, the response being measured by
the accelerometer (7), and the first resonant frequency was found with the aid of Lissajous
figures on the oscilloscope (9). The squares of these resonant frequencies, or rather the
squares of the frequency ratios, were plotted versus the axial load (see for example Figure
15.8b).
Though there was some scatter in these plots, fairly consistent predictions of the buckling
load could be obtained by extrapolating the best-fit lines of plots like Figure 15.8b. In
parallel, static central deflections were measured on the same columns and their buckling
loads were predicted from the corresponding Southwell plots. The VCT predictions P,, com-
pared quite well with Psourh and Pt/,· For example, for the SS-SS supported column of Figure
15.8b, the VCT predicted P,, = 267 kg was 92 percent of Psouth = 289 kg and 89 percent
of the calculated P," = 299 kg. Similar results were obtained in other columns of these
undergraduate student experiments, while only very few problems were encountered, veri-
fying the usefulness and relative simplicity of the method.

15.2.3 VCT for Determination of Boundary Conditions and Buckling


Loads in Plates
While the employment of vibration correlation techniques for columns has been found to be
feasible and practical from the early attempts in the fifties, their application to plates en-
countered considerable difficulties that were overcome only in the seventies.
Within the bounds of linear theory, the analysis for vibrations of an axially loaded column
can easily be extended to plates (see [15.14]), yielding the general result: "For any thin plate
of polygonal shape and uniform thickness which is simply supported along all the edges and
subjected to a uniform thrust N per unit length, the frequency w follows the relationship"
(w,, w, 0? = 1 - (N IN,,_,) (15.7)
(where w, 0 is now the nth frequency of the unloaded plate). This is the linear law for plates,
which is exact for simply supported polygonal plates. For perfect plates of arbitrary planform,
or with other boundary conditions, the deviation of the frequency ratio squared versus load
curve would, as for columns, be expected to be small. For circular plates this was shown by
Massonnet [15.11] indeed to be so.
In practice, however, the initial curvatures in thin plate have a significant effect that may
cause a considerable deviation from linearity. Lurie, for example, found in his experiments
on simply supported rectangular 24S-T duralumin plates of (bit) = 300 [15.14], both un-
stiffened and stiffened by a central channel stiffener, that the natural frequency decreased
only slightly with increase in load, yielding a frequency squared curve, as shown in Figure
15.9, curve A. The experimental w 2 points soon diverged from the theoretical linear law
predictions. Lurie first checked the reliability of his experimental setup. By freeing the two
unloaded edges, he effectively obtained a wide column, whose experimental w 2 values were
positioned on a linear theoretical curve, as expected, which "acquitted" the setup of respon-
sibility. Hence he attributed the deviation in curve A, and similar ones for other plates, to
initial curvature of the plates. He found support for this cause in Massonnet's nonlinear study
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

of imperfect circular plates [15.11], which produced similar-shaped w 2 - P curves for initial
Copyright Wiley
deflections
Provided by IHS ofwith
Markit under license the order of the thickness of the
WILEY plate. Hence
Licensee=McDermott Lurie concluded
Inc - India/8215328006, that the VCT
User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Vibration Correlation Techniques {VCT) 1253

3000

2
2500

2000

w 1soo
"' ~\
~
\\
\
Curve A
--

\Theoretical curve

1000
\
\

500
\
\
\
0
0 50 100 150 200 300
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

p
Figure 15.9 GALCIT VCT experiments on plates: test results for thin square 24S-T duralumin plates,
with (hit) = 300, simply supported (steel knife edges in V-grooves) on four sides (from
[15.14])

"method does not seem to have a practical application to flat plates, either stiffened or
unstiffened."
Over a decade later, Chailleux et al. [8.82] were more successful and showed that with
careful testing, application of VCT to plates, made of AG5 duralumin and of aluminum
reinforced by boron fibers, was practical and yielded reliable results. The load-frequency
squared plot for a nearly square duralumin plate of (bl t) = 200 is shown in Figure 15.10.
Extrapolation of the test results to the ordinate yielded a buckling load P,, = 117.5 kg,
compared to Psouth = 117-118.5 kg, obtained from the corresponding Southwell plot (Figure
8.71b of Chapter 8), an agreement within 1 percent. The critical load attained by VCT was
about 4.4 percent above the calculated theoretical value P,, = 113.5 for this plate. Similar
results were obtained for the eight boron-reinforced aluminum plates tested. Though the
initial imperfections of the plates tested were not specified in the paper, it was emphasized
there that VCT results "are reliable only if the test specimens have weak initial imperfec-
tions." One may assume, therefore, that the Ecole des Mines investigators made sure that
their plates had initial deflections whose magnitude was small in comparison to their thick-
ness of about 1 mm. This probably led to the success of their application of VCT.
As for columns (see Figure 15.6), one can also distinguish three zones in the dynamical
j2 - P curves for plates (though not indicated in Figure 15.1 0), with a large linear zone,
zone II, on whose extent the accuracy of the VCT prediction depends. Chailleux and his

.. • Pd 10 :::1175kq

Figure 15.10 Ecole des Mines (ENSTA), Paris, VCT experiments


on plates: dynamical curve, load versus frequency
0
Copyright Wiley
s .. c~ squared, for a rectangular duraluminum plate, simply
Provided by IHS Markit under license with WILEY supported on four Inc
Licensee=McDermott edges (from [8.82])
- India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1254 Nondestructive Buckling Tests

colleagues found that if zone II extended about 75 percent of the dynamical curve, the VCT
results were within a few percent of the Southwell and theoretical predictions. For narrower
linear zones (zone II) P, ,, obtained by VCT, could be significantly larger and was not very
reliable.
In the Ecole des Mines test setup for plates (Figure 15.6b), uniformity of loading was
achieved by a compensation system, (7). The uniformity of loading was adjusted in this
system by 12 independent blocks on top and bottom of the plate, positioned by screws, and
was controlled by four groups of strain gages (2) near each loaded edge of the plate. These
eight measured strains were picked up on a galvanometric light spot recorder (8), and the
coincidence of the eight light spots indicated uniform loading. This was one example of
some simple but effective experimental devices employed by Chailleux et a!. in their tests.
Their 1975 paper, which also excells in the detailed reporting of their work, is well worth
reading.
At about the same time, Jubb et a!. at Cranfield Institute of Technology, England, carried
out another series of tests applying VCT to plates [15.21]. They tested 1225 X 305 mm box
columns made from four 4.8-mm thick mild-steel sheet, flame cut and then welded together
by submerged arc welding. The end faces were then milled flat and parallel, yielding four
welded rectangular plate test specimens of (bit) = 64 and aspect ratio 4. Two such built-up
columns were made and tested.
The columns were loaded in a 1.5 MN hydraulic compression testing machine. For vi-
bration an electromagnetic exciter, suspended from the top, was used for one column, but
due to current limitations the amplitudes at high frequencies were very small. Hence for the
second column a Goodman's exciter, mechanically connected to the column, was employed
instead. The vibration equipment also included a decade oscillator, a capacity vibration pick-
up and meter, a power amplifier, an accelerometer and amplifier, and an oscilloscope. Dis-
placements were measured by a system of brackets, light rods, piano wires and dial gages,
and on the second column pairs of strain gages were cemented at the centers of the plate
elements.
In the experiment, the columns were first vibrated as the axial load was raised in incre-

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
ments of 50 kN, within the elastic limit, the natural frequencies being determined at constant
load for each step. The modes of vibration of interest were those for which there was one
half-wave across the width of each plate. The adjacent faces were then out of phase with
each other, such that the corners remained perpendicular and the plate elements of the box
column behaved as if they were simply supported along their longitudinal edges. Of partic-
ular interest was the mode with four half-waves along the plate elements, since their aspect
ratio was 4 and therefore their buckling mode could be expected to be similar (m = 4, n =
1). The natural frequencies were, however, measured for m = 1, 2, 3 and 4 and n = 1 at
each load step, up to an axial load of 700 kN. Then the columns were retested to failure,
with the dial gages fitted, plate buckling being measured at 754 kN. Frequency checks during
the retests on the influence of the displacement measurement system indicated that the effect
of attachment of the dial gages on the vibrations was negligible.
The natural frequencies were determined with the aid of Lissajous figures on the oscil-
loscope. The Cranfield investigators discussed the use of Lissajous figures for various cases
in detail, and the interested reader may therefore wish to consult their paper [15.21].
The variation of natural frequencies with axial load is shown for box column no. 1 in
Figure 15.11. The frequencies were non-climensionalized by dividing by f 0 , the fundamental
frequency of a stress-free, simply supported square plate of width equal to that of the plate
elements of the box columns tested. Note that the frequency fo of the square plate is equal

(m = the aspect ratio, here 4). The frequency ratio squared (f,,
to the natural frequency of a stress-free simply supported plate element in the buckling mode
f 0 ) 2 is plotted in the figure
versus the load ratio (PIP,.,.), for m = 1-4. The best fit of the descending points form = 4
touched the abscissa at the buckling load, obtained from the corresponding static test, veri-
fying the VCT prediction.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Vibration Correlation Techniques (VCT) 1255

Aoplied oxialload (kN)


0 100 200 300 400 5(X) 600 700 800 900 1000 !100
I I
II'
II
i 0 ~ rI
Theoreticol
efostic
I
iI Actual buckling
loo<F
754
1

kN~Iooa•I070 k~
i
buc<linQ

0 51

I
I X
) m=4 X X
N_ c
"0
,E
~

"~
0

f
0

~ 0
"
(f)

05
X

X X
m•2
0

05
X
X
X
m=!
0
0 01 02 03 04 05 06 07 08 0·9 1·0
PI~,

Figure 15.11 Cranfield Institute of Technology VCT experiments on plate elements of steel box col-
umns: variation of natural frequencies squared with axial load-box column no. ! -
theoretical curves, X experimental points (from J. E. M. Jubb, I. G. Phillips and H.
Becker, "Interrelation of Structural Stability, Stiffness, Residual Stress and Natural Fre-
quency,'' Journal of Sound and Vibration 39, (1), copyright 1975 Academic Press Ltd.)

The similar lines form = 1-3 represent other modes than the buckling mode and therefore
did not cut the abscissa at the buckling load, but at higher ones. Only the m = 3 line would
yield a buckling load not much different from the m = 4 one, indicating that the VCT results
were not very sensitive to slight inaccuracies in the excited vibration mode. A similar in-
sensitivity to vibrations at the exact buckling mode was later observed in vibration correlation
application to stiffened shells, discussed below, which facilitated the use of VCT.
One may note in Figure 15.11 the increase in vibration frequency after the plate had
buckled, which was caused by the nonlinear effects of the postbuckling distortions. The
definite change in slope of the frequency squared versus load curves could represent another
criterion for buckling.
This definite change in slope, or the lowest point of the frequency squared versus in plane
load, was employed as the criterion for buckling in a series of experiments on vibrations
and buckling of radially loaded circular plates, carried out at the Technion, Haifa, in the late
seventies [15.22]. The lowest-point criterion, which was originally proposed by Massonnet
[15.11], gave consistent results for these circular plates, as can be seen in Figure 15.12 (a)
and (b).
The six 200-mm diameter, (Rit) = 94.3, circular plates in these experiments were made
of 7075-T6 aluminum alloy, cut on a lathe, with edges rounded to fit into V-grooves and
thus approximately simulate simple supports. The loading system was identical to that em-
ployed for buckling of annular plates (see Figures 8.47 and 8.48 and [8.63]), discussed in
Chapter 8, Subsection 8.2.8, consisting of 30 segments, with V-grooves. that compressed the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

circular boundary of the plate uniformly. The vibrations of the plates were excited by an
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1256 Nondestructive Buckling Tests

1.5

0.5

(a) (b)

Figure 15.12 Technion VCT experiments of radially loaded circular plates-frequency ratio squared
versus radial load ratio for 7075-T6 aluminum alloy plates (from [15.221): (a) plate no.
1. (b) plate no. 5

electromagnet, whose core was bonded to the plate, and the frequency was changed by an
oscillator and read on a digital counter. The vibration response was measured with an LVDT
and resonance was detected by Lissajous figures on an oscilloscope. The buckling loads
determined from the frequency squared-radial load curves (as in Figure 15.12) were well
defined but about 15 percent below the perfect plate linear theoretical ones, a significant
difference.
Though VCT has been employed successfully in plates for two decades (see also [15.23]
and [15.24]), it appears that some further study is warranted to define more clearly its bounds
of applicability. In Subsection 15.2.9 some more recent applications of different vibration
correlation methods to plates are discussed.

15.2.4 Correlation between Vibrations and Buckling tor Shells


While vibration correlation techniques have been applied to columns and plates for decades
and some of them have become nearly routine, the methods developed for shells have not
yet entered general use, though their potential benefit exceeds that for other structural ele-
ments.
The earliest application of vibration correlation to shells was apparently that of Okubo
and Whittier to spherical caps under external pressure, carried out in the mid-sixties [15.25].
They tested six nominally identical shallow spherical shells to determine the decrease in
resonant frequency of a particular mode of vibration with increasing external pressure, and
by extrapolation found the pressures at which the frequency vanished. Then they used these
pressures as nondestructive estimates of the buckling pressure, which were found to be within
a few percent of the buckling pressures in three of the shells that were also tested to destruc-
tive buckling.
The models were small integrally machined clamped edge spherical caps with heavy
clamping rings, with (Rit) = 476 and having a geometric shell parameter [as defined by Eq.
(9.21) of Chapter 9] A = 7.46. This value of A corresponded, according to Huang [9.317],
to a theoretical critical buckling mode with n = 3 (three circumferential waves). Hence, the
lowest resonant frequencies for the corresponding vibration mode, with three nodal diameters
(n = 3), were expected to decrease with external pressure and vanish at the buckling pressure.
The variation of this resonant frequency f., 1 with external pressure is shown in Figure 15.13
for shell 4, one of the spherical caps that was also buckled destructively. One can clearly
see that extrapolation of the solid experimental points would cut the abscissa very close to
Copyright Wiley
the
Markitactual measured
with WILEY buckling pressure, indicated by theInchollow pointUser=G,
in the figure. The
--`,`,`````,`,`

Provided by IHS under license Licensee=McDermott - India/8215328006, Boopathi


No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Vibration Correlation Techniques (VCT) 1257

12

10

>--
~ 4
L1J
r--
'
::::>
fa 2 ,'---+-----,---+ Figure 15.13 Aerospace Corporation, El Segundo, California,
cr early VCT experiments on 7075-T6 aluminum al-
LL
0 c___c___c_ loy spherical caps-variation of resonant fre-
0 10 20 30 40 50 quency j 11 (lowest frequency with n = 3) with
PRESSURE psi external pressure for shell 4 (from [15.25])

theoretical curve shown in Figure 15.13 was an approximate estimate obtained by using the
normalized shape of the curve from Archer and Famili's calculations for free vibrations of
perfect clamped shallow spherical shells subjected to external pressure [15.26], while taking
the magnitudes of the frequency and pressure intercepts from the work of Kalnins [15.27]
and Huang [9.317]. A curve through the experimental points would have a similar shape to
this approximate theoretical one, though it would be lower and different for each shell, due
to geometrical imperfections and residual stresses. Indeed, the experimental points for the
other five shells (presented in Figure 2 of [15.25]) show a similar behavior to that for shell
4 in Figure 15.13.
In the experiments the shells were mounted at the end of a pressure chamber and loaded
pneumatically, the pressure being measured with a mercury manometer. The vibrations were
excited by a small electromagnet close to the convex side of the shell and sensed by a small
microphone close to its concave side. The position of the microphone was adjustable in the
circumferential and radial directions, to facilitate the determination of the mode shape.
Okubo and Whittier's experiments therefore included the essential elements of a modern
VCT. They were successful because their spherical caps had a unique buckling mode, whose
corresponding free vibration mode could be readily excited (an uncommon situation in iso-
tropic shells), and due to their careful experimental work. It is surprising, however, that their
successful application of VCT to shells was not followed by other researchers for some
years, till the initiation of the extensive investigations on stiffened cylindrical shells at the
Technion Aerospace Structures Laboratory in the early seventies.
Unlike unstiffened shells, stiffened shells are usually characterized by unique buckling
modes and are therefore natural candidates for vibration correlation methods. This is con-
spicuous in particular in closely stiffened shells, whose design is mostly governed by general
instability of the entire shell.
As pointed out in Chapter 13, Section 13.1, for closely stiffened shells the effects of
imperfections on the buckling behavior are less pronounced and the reduction in predicted
buckling loads and experimental scatter is less severe, whereas the influence of boundary
conditions becomes prominent. This fact has motivated extensive studies of these boundary
effects (see also for example [13.30]-[13.32], [1.31] and [3.11]). For stiffened cylindrical
shells the boundary conditions have similar effects on the lower natural frequencies of vi-
brations whose shapes resemble the buckling modes (see for example [13.39] and [13.40]).
The similarities in the influence of various parameters on the buckling of stiffened shells
and their free vibrations motivated the extensive studies at Technion to correlate these two
phenomena and draw from the nondestructive vibrations conclusions about the destructive
buckling behavior of these shells (see for example [13.31], [13.32], [13.40], [13.41], [13.94],
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`

[13.96], [1.31] and [15.28]-[15.30]).


For theoretical predictions of vibration and buckling in VCT applications to stiffened
shells, smeared stiffener theory is usually used, based on "smearing" the stiffeners uniformly
Copyright Wiley
Providedover
by IHS the
Markit shell in awithmanner
under license WILEY that takes the eccentricity into account,
Licensee=McDermott as detailed
Inc - India/8215328006, in Boopathi
User=G, Chapter 13,
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1258 Nondestructive Buckling Tests

Section 13.1 (or see [2.41] and [9.299]). For closely spaced stiffening this model is satis-
factory since the discreteness effects are generally negligible. In correlation studies between
vibration and buckling of stiffened shells, linear theory based on Fltigge's equations is usually
employed (see [13.40] and [15.28]), which is formulated as a special-purpose program
VIBUL, having the capability to accommodate a wide range of boundary conditions, in-
cluding elastic boundary restraints.
The use of linear theory as an approximation is justified here since for stringer-stiffened
shells the nonlinear prebuckling deformations were shown to have a relatively small effect
for buckling, and even smaller for vibrations (see for example [13.32], except for heavily
stiffened shells or very short shells or in the case of significant eccentricity of loading, and
the nonlinear prebuckling deformations can therefore usually be neglected. For more details
on the general adequacy and bounds of validity of the linear smeared-stiffener theory the
reader is referred to Subsection 13.1.6 of Chapter 13.
In the presence of load eccentricity, or for practical boundary conditions that may include
unknown load eccentricities (see for example [13.94)]), a theory that considers nonlinear
prebuckling deformations must be used . A well-known multipurpose program, BOSOR 4
[2.62] that answers these requirements is therefore employed in the Technion investigations.
This program also uses smeared stiffener theory for stringers, but discrete rings can be
introduced. It has been extended to the case of axial elastic restraints combined with load
eccentricity (see [15.31] and [10.2]), the extension being based on a model proposed by
Bushnell, and has also been used for prediction of vibration frequencies and buckling loads
of spot-welded stringer-stiffened shells with axial elastic restraints.
The influence of boundary conditions, and in particular the in-plane boundary conditions,
and of prebuckling deformations on the buckling and vibrations of stiffened cylindrical shells
was discussed in Subsection 13.1.5 of Chapter 13. Some of the important features reviewed
there accentuate the motivation for application of VCT to stiffened shells.
The vibration correlation methods correlate the behavior of actual imperfect shells, mea-
sured in the vibration tests, with a theoretical model of a perfect shell. For the methods to
be valid, one has to show that initial imperfections have similar effects on buckling and
vibrations. This motivated theoretical studies on the influence of imperfections on vibrations
of cylindrical shells at the Technion (see for example [15.32]-[15.34]) and consideration of
other similar studies (for example [15.35]). It was found that imperfections have an influence
on the frequency of the vibrations similar to that which they have on buckling of cylindrical
shells, not only at high compressive loads but also at zero axial loads. Physically, the fre-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

quency reduction at zero axial load means that the small changes in geometry of the shell
implied by the initial imperfection reduce the stiffness appreciably for certain types of vi-
brations with no change in mass, the effect being largest for the imperfection and vibration
modes corresponding to those that also exhibit the strongest influence on the buckling load.
Extending the investigations to axially compressed stiffened cylindrical shells [ 15.33]
showed that, as in the case of buckling [3.7], the influence of the initial imperfections on
the vibrations of stiffened shells was indeed significantly less than on the corresponding
isotropic shells, though the shape of the curves was similar. The extensive parametric studies
carried out confirmed the similarity of the influence of initial imperfections on buckling and
vibrations, though the influence on vibration was somewhat less pronounced than that on
buckling.
Based on these theoretical considerations, practical vibration correlation techniques have
been developed for stitiened shells. The VCT for shells is again classified into two groups
according to their approach: (l) those for determination of boundary conditions, and (2)
those for direct determination of buckling loads.

15.2.5 VCT for Determination of Boundary Conditions in Shells


The VCT for determination of boundary conditions consists essentially of an experimental
determination of the lower natural frequencies for a loaded shell and evaluation of equivalent
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Vibration Correlation Techniques (VCT) 1259

elastic restraints representing the actual boundary conditions. It is based on the similarity
between the strong influence of axial and rotational restraints on free vibrations of stiffened
shells, in particular for the lower natural frequencies whose mode shapes resemble the buck-
ling modes, and that observed for buckling loads (see for example Figure 15.14, showing
the predicted influence of elastic axial restraint on the buckling and vibrations of a typical
simple-supported stringer-stiffened cylindrical shell R0-41, and Figure 15.15, showing the
variation of frequency squared with axial load for this shell and that for a nominally identical
shell R0-42 with nominally clamped boundary conditions).
Shells R0-41 and R0-42 were typical 7075-T65 aluminum alloy integrally machined
stringer-stiffened shells of one of the Technion test series, similar to the ones shown in
Figure 13.16, with geometrical properties (RI h) = 492, (L/ R) = I, (A 1I hh) = 0.72, (/ 11 I
bh1) = 3.1, stringer eccentricity (eJh) = -4.1 and twisting stiffness due to stringers 7], 1 =
9.1. Shell R0-41 was supported on laboratory-type approximate simple supports, a V-groove.
It was nominally simply supported, and the real boundary conditions can be represented by
an axial elastic restraint k1 between SS3 and SS4. Figure 15.14a (reproduced from [15.36])
shows the variation of buckling load P,. between SS3 and SS4 for this shell, as calculated
with VIBUL. The variation between SS3 and SS4 is obtained by the axial spring k1, the
stiffness of which is zero for SS3 and infinity for SS4 B.C.'s (the other B.C.'s are w = 0,
M, = 0, v = 0). In Figure 15.14b the influence of elastic axial restraints on the frequency
squared of the vibration mode m = 1, n = 12, at an axial load of 1700 kg, is shown in the
same manner as the buckling load in Figure 15.14a. This mode of vibration was chosen
because it represents the buckling mode here. The shape of the two curves is indeed similar.

ss'
~r-------------------~L------------------
Pcr

n:::ll, m"'l

0
(a)

0~0~------------------------L/
ss 4
_______________
1'.1o·• -------------~-------
-~~~~~-~-~-~-~-~-~-~-~-~-~-~-~--:--
2
{Hzl

Figure 15.14 Technion VCT experiments on stiffened cylindrical shells~infiuence of elastic axial
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
restraints on the buckling and vibrations of shell R0-41 (from [15.36]): (a) buckling
load, (b) vibrations (at P = 1700 kg, m = I, n = 12)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1260 Nondestructive Buckling Tests

EXPERIMENT
.& RO- 42
• RQ- 41

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
1700 3400 5100

EX PE RIME NT
• ;.((J --4~
3.0
e RO -~Z

I n :a ~
m•l j

2.0

1.0

s~; :~

L___ _ ___jl_ _ _ _ _ _ _, _ -- _ __j_ ---


0 1700 3400 pko 5100

Figure 15.15 Technion VCT experiments 011 stiffened cylindrical shells-frequency squared versus
axial load for shells R0-41 and R0-42 (from [15.36]): (a) m = I, n = 11, (b) m = 1,
n = 8.

Hence, by measuring the natural frequency at a relatively low load (1700 kg here is about
one-quarter to one-third of the expected buckling load), one can determine the effective
boundary conditions. For clamped boundary conditions a similar procedure can be employed
by introduction of a torsional spring k 4 between SS4 and C4 (see [13.32] and [13.40]).
The frequencies squared for shells R0-41 and R0-42, predicted with the VIBUL program,
are plotted versus the axial load in Figure 15.15, as are the corresponding experimental
points. These were obtained in a test setup (see for example Figure 2 of [13.40] or Figure
6 of [13.32]), which is similar to a more recent one shown in Figure 15.21.
The experimental technique consists of vibrating the loaded shell by an exciter, regulated
Copyright Wiley
byMarkit
Provided by IHS anunder
outside oscillator,
license with WILEY and then detecting resonance frequencies
Licensee=McDermott and mode
Inc - India/8215328006, shapes with an
User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Vibration Correlation Techniques (VCT} 1261

outside scanning microphone. In the earlier tests an acoustic driver installed inside the shell
was employed for excitation since it had been been found superior in terms of response and
avoided direct contact. In later tests, however, difficulties were encountered in identification
and separation of the resonant modes. Though these were partly eliminated by addition of
a narrow-band filter, direct-contact excitation was again tried.
A series of comparative tests was canied out (on a shell of the AB series-see [13.94]
and [13.95]) between excitation by an electromagnetic driver suspended on springs outside
the shell and acoustic excitation. In these tests detection of resonance by an accelerometer
instead of the microphone was also compared. The main conclusion that emerged was that
direct-contact excitation is superior, for clean vibration of the shell alone, to acoustic exci-
tation, which induces vibrations of the whole system of shell and end rings. It was also
found that the small mass of the contact leg of the electromagnetic shaker has only a neg-
ligible effect on the frequencies and mode shapes. Furthermore, the use of an accelerometer
for detection of resonance instead of the microphone was found not to be justified, since it
yields only a very slight improvement. The direct-contact electromagnetic driver was, there-
fore, employed in all the succeeding tests.
At each load step in a typical experiment, the excitation frequency is changed and reso-
nance is detected by Lissajous figures, or output amplification, on the oscilloscope. With
resonance identified, the mode of vibration is mapped by scanning the shell with the micro-
phone and plotting its reading versus its circumferential and axial position on X- Y recorders.
The load is applied with a screw jack and measured by a load cell, and the load distribution
is checked with an array of five pairs of uniformly spaced strain gages. In order to verify
the presence of load eccentricity, additional strips of five closely spaced strain gages were
bonded near the edges in some shells. The reader may find more detailed descriptions of the
test procedure in [13.32], [13.40], [13.94] and [13.95].
The experimental points in Figure 15.15 show a trend to a certain value of k1 for R0-41
and a certain k~ for R0-42. Note that the vibration mode in Figure l5.15(a) (m = 1, n =
11) is very close to the predicted buckling mode, but other vibration modes in the vicinity,
as, for example, Figure 15.15(b) (m = 1, n = 8), also show similar trends. This insensitivity
to the exact mode of vibration was confirmed in the many tests canied out and is important
for the reliability of the method because in practice one can only hope to excite vibration
modes close to that coJTesponding to the buckling mode, and not always the exact one
(especially because the experimental buckling mode usually has slightly fewer circumfer-
ential waves than the predicted one). However, in the presence of significant load eccentricity
the situation is different, as will be pointed out later.
From Figure 15.15(a), R0-41 appears to approach the SS4 curve and R0-42 that of k4 =
2000. Further study of Figure 15.15(b) and similar curves (see [15.36]) yielded k 1 = 53
(which is very close to SS4) for R0-41 and k4 = 2330 for R0-42. The predicted buckling
loads for these elastic restraints are compared with the experimental ones in Table 15.1 (from
[1.31]).
The technique was applied to shells of different Technion test series for axial compression
loading. Figure 15.16 (from [13.41]), presents the results for 35 shells and shows the sig-
nificant reduction in scatter as a result of the experimental determination of the boundary
conditions, pointed out in earlier studies and reviews. The scatter of the knock-down factor
is reduced from 0.6-1.3 to 0.6-0.9, the low values of p, 1, relating to clamped shells, probably
because of the additional imperfections introduced by clamping, which were mentioned in
earlier studies (see for example [10.2]).

Table 15.1 Vibration correlation technique: experimental and theoretical buckling loads and modes
Exp. SS3 SS4 C4
P,.,,, mode P," mode P_,,.; mode Pc.; mode P," Mode
Shell (kg) (n/m) (kg) (nlm) (kg) (nlm) (kg)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
(nlm) (kg) (nlm) p,,.
R0-41(SS) 4650 8/l 4418 II/] 5986 12/l 5830 12/1 0.80
CopyrightR0-42(C)
Wiley 4875 lOll 6036 12/l 8245 12/l 7615 12/l 0.64
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1262 Nondestructive Buckling Tests

0
1.2 0 0

P, Psp 1.0

,6
P Psp
SIMPLY ~

,4 SUP£0RTE0~9
CLAMPED 0 •'

.2

0 ,1 .2 .3 .4 .5 .6 ,7 .8 .9
A1/b h

Figure 15.16 Technion VCT experiments on stiffened cylindrical shells: knock-down factor of 35
simply- supported and clamped stringer-stiffened shells of RO, AB, KR, SN, RS, DUD
and DK series, corrected for experimentally (nondestructively) determined boundary con-
ditions. p,P' the corrected knock-down factor, is compared to the original PsS1 and Pc"
values (from [ 13.41])

15.2.6 Application of VCT to Practical Boundary Conditions and


Realistically Fabricated Shells
Having been established as a reliable technique in the laboratory, the vibration correlation
method is still of limited value unless it can also be applied to realistic boundary conditions.
Considerable effort has therefore been directed to extending its application to shells with
end supports, simulating joints employed in actual engineering construction. Such joints may
also have load eccentricity, which may not be well defined a priori and may depend on the
tolerances and behavior of the joints under load.
Now, whereas the lumping of the influence of boundary conditions and imperfections
inherent in the vibration correlation technique gives no reason for concern, since their be-
havior in vibrations and buckling is similar, the effects of load eccentricity differ for vibra-
tions and buckling, as was revealed by earlier vibration correlations studies on shells with
prescribed load eccentricity, and can therefore obscure the correlation.
Tests of the AB shells (see Figure 13.16) on practical boundary conditions [13.94] then
emphasized that the load eccentricity indeed depends on the tolerances and behavior of the
end connections under load. The Technion AB series consisted of six integrally stringer-
stiffened 7075-T6 aluminum alloy shells, similar to the earlier ones tested on laboratory-
type end rings, but riveted to end rings that simulated a typical missile clamp joint (turned
to the inside for experimental convenience; see Fig. 15.17t). Instead of the external circum-
ferentially tightened clamp that would appear in a missile, clamping of the end rings (riveted
to the shell) to the support rings (which sat in the grooves of the top and bottom baseplates)
was achieved by two relatively heavy annular plates, tightened together by a series of bolts.
Though the joint was designed without any nominal load eccentricity, such a load eccen-
tricity of different magnitude actually occurred in the AB shells under load, as was initially
only suspected from the soft buckling behavior of the shells. This suspicion was later verified
by significant bending strains in pairs of strain gages that were bonded (in special recesses
in the end rings) very close to the edges of some of the shells. This apparent eccentricity of
loading was expected also to affect the vibrations, as had been observed earlier in tests with
prescribed load eccentricity [15.29], and would thus obscure the vibration correlations unless
the eccentricity effects could be identified and separated.
A thorough study of the data of the AB shells eventually provided a solution to this
problem and revealed a salient property that distinguishes the vibrations in the presence of
significant load eccentricities. This important property is the large increase in the frequency
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

ratio squared f If ss 41J 2 with the number of circumferential waves n of the vibration pattern,
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Vibration Correlation Techniques (VCT} 1263

l
Figure 15.17 Technion VCT experiments on stiffened cylindrical shells-boundary conditions of shells
tested in the vibration con-elations studies: (a), (b), (c) RO and SN shells; (d) RO, ST,
KR and DUD (2 and 3) shells; (e) DUD (4-7) shells; (f) AB shells; (g) DK shells (from
[13.41])

which does not occur in the absence of load eccentricity. Figure 15.18 (reproduced from
[13.941) shows a plot of this ratio versus n for a typical stiffened shell AB-5, with the missile-
joint type boundary conditions shown in Figure 15.17f. The frequency ratio here is the ratio

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
of the experimentally observed frequency to that predicted by VIBUL for SS4 boundary
conditions. If one now plots in Figure 15.18 the corresponding theoretical frequency ratio
squared for some likely values of load eccentricity, where f ss 4 +;; is the frequency computed

1.8

1.6

1.4

1.2

1.0

0.8

0.6

0.4 SHELL AB- 5

0.2

noL-~--~--~~--~--~~---L--L-~
4 5 8 9 10 11 12 n

Figure 15.18 Technion VCT experiments on stiffened cylindrical shells-variation of frequency ratio
squared with circumferential wave number n, at P = 1000 kg-shell AB-5 (from [13.94])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1264 Nondestructive Buckling Tests

with BOSOR 4 (which considers nonlinear prebuckling deformations) for SS4 boundary
conditions in the presence of load eccentricity e,
the same property becomes evident. One
can easily find a load eccentricity (e/h) that has the same slope as the experimental one.
Here the apparent load eccentricity is (e I h) = 2.0. Note that the experimental frequency
ratios are lower than the corresponding theoretical ones since they also include imperfections
which are not taken into account in the theoretical predictions, as well as the difference in
boundary conditions from the SS4 ones considered in the comparison.
This frequency slope property therefore presents a tool for a nondestructive identification
of significant unknown load eccentricity in stiffened cylindrical shells. In [ 13.94] a detailed
modified vibration correlation method to define the boundaries, once the load eccentricity
has been identified, was presented. The main features of this modified method are now
recapitulated.
The frequency ratios are now all referred to SS4 boundary conditions since the modified
method tacitly assumes that the real boundary conditions in the presence of significant load
eccentricity are not far from SS4 and, hence, SS4 can be used for reference. When the actual
boundary conditions are far from SS4, the method is still effective but has a more empirical
flavor.
The procedure of the modified VCT is shown schematically in Figure 15.19. To a typical
plot of experimental frequency ratios squared versus n (like Figure 15.18), a theoretical curve
(calculated with BOSOR 4) is fitted for a load eccentricity (elh) 1 that has the same slope.
An appropriate value of n is chosen, say n 1 , which is usually taken as that at which buckling
is predicted, or a value slightly below it (based on the observation that the experimental n
in most tests on stiffened shells is just slightly below the theoretically predicted one). For
n 1, the effective theoretical frequency ratio squared, kA, and the corresponding experimental
one, km are read off the curves. Since the experimental values include the boundary effects
and also part of the influence of imperfections, the ratio kc = (k 8 / kA) represents an equivalent
boundary condition factor. With this k0 the procedure then follows the correlation method
in its regular form. An equivalent frequency squared j2 = kc · (f ss 41Y replaces the measured
frequency squared.
This modified VCT was applied to the AB shells and yielded consistent results that fit
very well those for shells without load eccentricity plotted in the summary diagram, Figure
15.16.
Since, however, the slope property is an essential element of the nondestructive method
for detection of unknown load eccentricities in the presence of other boundary effects, its
verification under controlled conditions is crucial. The results of earlier tests [15.29] of
similar shells on laboratory-type boundary conditions with prescribed load eccentricity (see
Figure 15.17a-c) were therefore analyzed in [13.94] in the same manner as the shells on the
practical boundary conditions and confirmed the expected positive and zero slopes in the
presence and absence of load eccentricity, respectively.
Additional tests were also carried out on seven SN shells [15.37] with prescribed load
eccentricity, under similar laboratory boundary conditions. In five shells outside load eccen-
tricity was achieved by applying the load through the tip of the stringer (Figure 15.17c) or
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 15.19 Technion VCT experiments on stiffened cylindrical


shells-determination of equivalent boundary con-
dition factor for eccentrically loaded shells (from
[13.94])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Vibration Correlation Techniques (VCT} 1265

the middle of the stringer (Figure 15.17b), and two had zero load eccentricity (Figure 15.17a).
The frequency ratio squared (f If ss 4L)2 for one pair of shells of very similar geometry, SN-
7, with prescribed outside load eccentricity (elh) = 6.28, and SN-4 with zero load eccen-
tricity, (el h) = 0, is plotted versus n in Figure 15.20. The essential slope property is again
clearly demonstrated. Having been confirmed for stringer-stiffened shells with different ge-
ometries and different boundary conditions, the method of identification of load eccentricity
appears to be fairly reliable.
The vibration correlation method for definition of boundary conditions has been applied
to a variety of boundary conditions shown in Figure 15.17 and discussed in [13.40], [13.94],
[15.26] and [15.36]-[15.39].
The tests on the DK shells [15.39] continued the study of problems of "lumping" of
different boundary effects, in particular of the necessity and means to isolate the influence
of load eccentricity, started on the AB shells. The DK shells are supported at top and bottom
on a type of aerospace joint (see Figure 15.17g), consisting essentially of two externally
flanged rings, whereas the type of joint employed for the AB shells was based on internal
flanges. The details and dimensions of the flanges differ for the different shells in order that
the influence of variations in stiffness of ends can be studied.
Shell DK-2, one of the seven shells tested in this series, is shown in Figure 15.21 posi-
tioned in the present setup for small shells under axial compression loading, which is a
slightly improved version of that employed in earlier tests. In the test rig the outer flanged
rings fit into grooves in two heavy end plates, which are then filled with Cerrobend (see
Figure 15.17g). After the axial force is applied by the screw jack, it is measured with a load
cell and the load distribution is again checked with an array of five pairs of uniformly spaced
strain gages. Compression and bending strains are read directly through a switching and
bridge circuit on a digital strain indicator.
As in earlier tests, the shell is excited by means of an electromagnetic shaker whose leg
is glued to the shell. The excitation frequency is changed by the regulating oscillator and

1
I f I
!-I
I, f SS4 I

/~
Th•ory 554,;; ( Bosor 4)- --~
2 0 SN- 7
0
o o
22

16r/
1,8

1.4

I 2
s~~~

I 0

0.8 SN-4

0.6r·
0 4

0 2

L----.L--o-------':---:':--':--- _l __ _ j _ _J__~_
9 10 II 12 n

Figure 15.20 Technion VCT experiments on stiffened cylindrical SN shells having laboratory-type
boundary conditions with prescribed load eccentricity: Variation of frequency ratio
squared with circumferential wave number n, at P = 1200 kg, for shells SN-7 (elh) =
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

6.28 and SN-4 (elh) = 0 (from [15.37])


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1266 Nondes truc tive Buckling Tests

Fij!UI'e 15.21 Technion vcr C\pcrimenh on 'ullcncd 9hndrical \hells-tc't \ClUfl l<>r recent n.'ial
comprc"ton test, !from [13.').1))

\\hen resonance i~ identtlicd. the frequency ts read on an electronic counler and the mode
of vibration is mapped by \Cannmg the shell wtth a nucrophone and plotting tl\ reading
versus its ci rcumferential and axial position o n an X-Y recorder.
Seven nomi nally identical OK shells were tested. Their geometrical properties arc pre-
\Cllled in ( 13.41) and [ 15.39 (. Sign ificant load eccentricities were identified from the vibra-
tiOn resu lts. T he predicted huci.Jing load~ for the cllccti'c boundary conditions obtained by
the modified vibration corrdatum technique P arc <.:omp:trcd with the expcnmcntal rt!sults
for live shells in Table 27.3 of ( 13.41) and for all the se\ en '>hells in Table 2 of ( 15.39]. The
result<. reconfirm that the 'thnttion correlation tc..:hmquc performs well in the dclimllon of
actual boundary condttton\.
Though VCT for dclimttun of boundary condittons had hccn applied to different t) pes of
boundary condi tions as shown in Figure 15. 17. all the shells tested were integrally mach ined
stringer-st iffened cylindrical shells of 240 nun diameter. To transform VCT into an industrial
tool. the techniq ue has more recent ly also been applied to two series of larg~r ~hells of
different construction. Both test serie., were part of a joint Tcchnion-RWTI I Aachen rc~earch
project. One. carried out ;It Hat fa [10.22] and [15.40]. u'ed 500-nml ( 19.7·inl dt;lm~tcr \pot-
"cldcd and riveted stringcnuffcncd ;hells (see f'tgurc 15.22). and the other. at Aachen

t'i~:u rc 15.22 Technion VCT c'pcrimcnts on fabricated sliil'cncd 'hcils-postbuckling jl<Jllcrns of typ-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

ical spot welded and rh eted aluminum alloy 'hells (from [ i 5.40]1
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Vibration Correlation Techniques (VCT} 1267

113.1-10), employed large 1.9-m (74.8-in) diameter !ltringcr-stiffened corrugated ~hells (sec
Figure 5 of [1.17] and Fi gure 2 1 of 113.140]).
The spot-welded and riveted specimens (Figure 15.21) were designed to be on as ~mall a
scale as is feasible with this type of construction. To fuci litate comparisons, their nom inal
geometrical parameters (Ri ll - 5 10. LI R = 1.33, A,I!Jil - 0.67 - 0.82, 111 /bh' = 7. 1-7.8,
e,llo = 3. 12 and -3.50) were intentionally similar to those of the smaller integrally stiff-
ened she lls. The Haifa teM series (see 110.221 and [ 15.401> included seven spot-welded and
two riveted L cross-section-type stringer-stiffened ~hells (who~c dimensions and geometrical
parameters were presented in Table I of [10.22] and Table I of [15.40)). The boundary
conditions of the first two •hells. AAC- 1 and AAC-2. were nontinal simple suppons. The
re\t of the shells. AAC-3 to AAC-7 and the two riveted stringer-stiffened shells. AACX-1
und AACX-2, were clamped tO the loading ring in a circular groove filled with Cerrobend.
in a manner that nominally rcprcsenls C4 boundary conditi ons.
The 50 ton MTS test system of the Technion Aircraft S tructures Laboratory was adapted
to these shells, and a special mu lt ipurpose scanning and measurements system was developed
I I 0.20]. A noncontact probe was used to measure bot h vibrations and imperfections of the
cyli ndrical shells. The mca!lurcments include natural frequencies. modes of vibration and
mapping of imperfections. All measurements and mapping were carried out by the same
probe inside the closed shell. and an electronic control system permitted automatic execution
of the different modes of operation. Here the vibration mode is described. The imperfection
mode is discussed in Chapter 10, Section 10. 7.
Figure 15.23 shows the complete system, (a) without the ~hell and (b) with the shell in
position for vibration tests. In the vibration mode the test procedure is essentially similar to

Figure 15.23 Technion VCT cxpcl"i oncnts on fabricated stiiTcncd shells-test setup for spot welded and
riveted shell s (from 115.40]): (a) noncontact scanning and measuring 'Y'tcm, (b) test
system in operation --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1268 Nondestructive Buckling Tests

that employed on the smaller, integrally stiffened shells discussed in the previous section.
The shell was again vibrated under a constant load over a range of natural frequencies, the
modes of which were in the neighborhood of the predicted buckling mode for the shell with
nominal boundary conditions. The natural frequencies were recorded and the corresponding
modes were plotted on an X- Y recorder. This information was later used to draw the fre-
quency versus load curves (for a given mode) from which the boundary conditions could be
defined. Though the actual operation of the system differs, the vibration plots obtained are
similar and the vibration correlation technique is applied in the same manner.
For example, the frequencies squared for the riveted shell AACX-1 (predicted with
VIBUL) for vibration modes m = 1, n =' 11 and m = 1, n = 12 are plotted versus axial
load in Figure 15.24. The experimental points obtained for this mode, which is close to the
predicted buckling mode, are also plotted in the figure and indicate effective boundary con-
ditions between SS4 and C4.
The effective elastic restraint, obtained as always in VCT from the measured frequency
squared with the aid of a plot of variation of frequency squared with elastic restraint, was
for this shell a rotational spring k 4 = 2680. The predicted buckling load for this effective
rotational restraint was P,, = 25,600 kg. Similar consistent results were found for the other
AAC and AACX shells.
The predicted and experimental buckling loads and modes for the AAC and AACX shells
are presented in Table 15.2 (reproduced in part from similar tables in [10.22] and [15.40],
where the test results of these shells are discussed in detail).
Note that the corresponding Table II of [ 10.21] and Table 3 of [ 15.40] include the predicted
buckling loads for the nominal boundary conditions Pss 3 , Pss 4 and Pc4 , whereas in Table

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
~
~-
AACX -1

a
woo 4000 6000 8000 10000

b
L---~,o~oo~--~,o~oo~---~.o~oo~--~.oo~o~--~~~oo~oo~~
P[kg]
Figure 15.24 Technion VCT experiments on fabricated stiffened shells-frequency squared versus axial
compression for riveted aluminum shell AACX-1 (from [15.40]): (a) n = 11, m = I, (b)
n = 12, m = I
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Vibration Correlation Techniques (VCT) 1269

Table 15.2 Experimental and predicted buckling loads and modes, and "knock down" factors for AAC
shells
Shell P,,, (kg) p"" (kg) Pss3 Pss4 Pc4 p\j) Prh tlp
AAC-1 16,930 (12) 7,380/ 11,290(10)* 0.562 0.340 0.436/0.667* 0.640 -0.03
AAC-2 18,300 (12) 6,800/ 11,740(11)* 0.484 0.313 0.372/0.642* 0.570 -0.07
AAC-3 25,600 (12) 15,000(10) 1.068 0.692 0.548 0.586 0.527 -0.06
AAC-4 23,850 (12) 12,000(10) 0.921 0.589 0.470 0.504 0.495 -0.01
AAC-5 24,650 (12) 12,150(10) 0.932 0.596 0.476 0.493 0.453 -0.04
AAC-6 27,200 (12) 13,900(10) 0.990 0.641 0.508 0.511 0.480 -0.03
AAC-7 24,650 (12) 13,000(9) 0.998 0.638 0.509 0.527 0.443 -0.11
AACX-1 25,600 (12) I 0,200(9-1 0) 0.726 0.470 0.373 0.400 0.350 -0.05
AACX-2 24,350 (12) 10,800(10-11)/ 0.769 0.498 0.395 0.444/0.510* 0.520 0.01
12,360(11)*
( ) = Circumferential wave number
*= "Equivalent experimental buckling load" based on panel buckling stress (see discussions for AAC-1, 2 and AACX-
2), upper number corresponds to overall experienced load at buckling/lower number represents the panel equivalent
expe1imental buckling load.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

15.2 these are replaced by P,", the VIBUL prediction for the elastic restraints obtained by
VCT. One may also note that (as pointed out in [10.22] and [15.40]) shells AAC-1 and
AAC-2 buckled initially at relatively low loads and with a single lobe, rather than a fully
developed buckled surface. Further study indicated that there was significant non-uniform
loading in these shells, which resulted in the regions of the single lobes buckling as a
stiffened panel subjected to intensified local stresses. Hence, an equivalent experimental
buckling load corresponding to the local panel buckling was considered for these two shells,
which is denoted by * in Table 15.2. A similar behavior, but with two initial lobes, was
observed for shell AACX-2, leading to a similar equivalent experimental buckling load also
for this shell.
The knock-down factors Pss 3 (P""/P 553 ) etc. appear in columns 4-7 of Table 15.2,
where
(15.8)
is the one that refers to the actual boundary conditions. Since the initial imperfections were
also scanned and recorded, theoretical imperfect buckling load ratios, or theoretical knock-
down factors
(15.9)
could be obtained from the reduced imperfection data, using the MIUTAM multi-mode anal-
ysis of [10.61], as discussed in Chapter 10 (or see [10.2] and [10.21]). In Eq. (15.9) P 553 , , ,
is the buckling load computed from the reduced measured initial imperfections by MIUTAM
and P553 is the classical linear smeared stiffener theory buckling load. This Prh appears in
the next-to-last column of Table 15.2 and is an indication of how imperfect the shell was
initially.
A comparison between Prh and p,1, carried out for 14 of the Technion integrally stiffened
shells, is shown in Table 15.3 (reproduced from [13.41]).
In general, the comparison is very good, although only fair for shells KR-1, DUD-2 and
DUD-3, which were clamped. The main reason for the discrepancy between VCT experiment
and imperfect shell theory is probably that the initial imperfections were measured in these
shells before being fixed in their final boundary conditions and the final clamping no doubt
produced additional imperfections not accounted for in p,".
For some of the shells having significant load eccentricity the comparison is also only
fair, and for some shells with large load eccentricity, not presented in Table 15.3, it is even
inferior, indicating the need for further improvement in prediction methods in the case of
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1270 Nondestructive Buckling Tests

Table 15.3 Comparison between buckling loads predicted from


measured initial imperfections and by VCT
Predicted Experimental
Shell Prh P.,p l:ip

AB-5" 0.87 0.76 0.11


AB-6 0.72 0.75 -0.03
KR-1 0.84 0.68 0.16
SN-1 0.95 0.91 0.04
SN-4 0.88 0.82 0.06
SN-6" 0.73 0.63 0.10
DUD-2 0.86 0.63 0.23
DUD-3 0.70 0.58 0.12
DUD-4" 0.88 0.80 0.08
RS-36 0.91 0.76 0.15
DK-1" 0.89 0.83 0.06
DK-3" 0.78 0.71 0.07
DK-4" 0.89 0.73 0.16
DK-5" 0.80 0.84 -0.04
"Significant load eccentricity.

large load eccentricity. For the larger spot-welded and riveted shells, the !::J..p values appear
in the last column of Table 15.2, and the comparison seems to be even slightly better than
that for the integrally stiffened shells.
To return now to shell AACX-1, which was considered as an example, it was observed
in Figure 41 of [15.40] that its circumferential axial strain distributions were fairly even and
that it buckled with a completely developed buckle surface with n = 9-10, m = 1. However,
it can be seen in Table 15.2 that p," = 0.400 for this shell is considerably lower than the
already relatively low values of the knock-down factors obtained for the spot-welded shells
AAC-1-7. Yet, as for these shells, this value is a little higher than the theoretical imperfect
p," = 0.350 corresponding to this shell. Hence, this low value of p,P' and in particular that
of the theoretical knock-down factor p," reflect the effect of the geometrical initial imper-
fections of the shell, which were more pronounced in shell AACX-1 than those measured
in the spot-welded shells. Also, as for the spot-welded shells, the difference between p," and
Prh• which is an indication of the success of VCT, !::J..p = -0.05 is small. This shows that for
this shell the actual boundary conditions were fairly well defined by the VCT and that the
lumped effect of imperfections in their definition was relatively small.
The similar small values of !::J..p (the differences between p,1, and p,") that appear in Table
15.2 and the earlier results of Table 15.3 add confidence to the application of VCT to
stiffened shells.

15.2.7 VCT for External Pressure and Combined Loading


The vibration correlation technique was originally developed for axial compression because
for this loading case the largest discrepancies occur between tests and predictions. Since
boundary conditions also significantly affect other loading cases, and in particular the im-
portant cases of external pressure loading and combinations of external pressure and axial
compression, the vibration correlation technique was also extended to them [ 13 .107]. This
test program involved steel ST shells and aluminum alloy KR shells, all integrally stringer-
stiffened cylindrical shells of geometries similar to the RO and AB shells, and established
the applicability and usefulness of the technique as a nondestructive method for improved
definition of boundary conditions also for external pressure and combined loading.
The results of the KR series were also used to evaluate theoretical interaction curves and
to obtain improved curves by referring to effective boundary conditions found with the aid
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Vibration Correlation Techniques (VCT) 1271

of vibration correlation. In order to obtain better interaction curves, two additional series of
tests, the DUD shells, with nominally clamped and simply supported integrally stringer-
stiffened shells, were carried out (see [15.38], [15.69], [13.107] and [9.273]).
The shells were tested in the setup shown in Figures 13.47 and 13.48 of Chapter 13, and
the loading and vibration procedure was described in Subsection 13.4.1. Typical results for
combined loading on shell DUD-10, frequency squared versus external pressure p, under
constant axial load P = 380 kg (n = 8, m = I) are shown in Figure 15.25. Interaction curves
for shell DUD-I 0 are presented in Figure 15.26, indicating the improvement in the interaction
curve obtained by the vibration correlation technique. It should be pointed out that this
improved interaction curve is obtained by a series of definitely nondestructive tests on the
same shell, at different combinations of axial compression and external pressure, care being
taken not to exceed about half the combined theoretical (linear theory) buckling load at each
load combination. This ensures that as many repeated vibration scans as required can be
performed without any damage or permanent deformation of the shell. From the vibration
scans the equivalent springs for each load combination were found, and these were then used
to compute the buckling load combinations, whose locus is the interaction curve marked
"INTER (from vibs. test)."
The second phase of the test on shell DUD-I 0, as on the other shells of this series, was
buckling of the shell. Special care was taken to detect buckling and to release the load
immediately after buckling. This permitted repeated buckling of the same shell with a rel-
atively small decrease in buckling load shown in Figure 15.26. Similar results were obtained
for the other shells in the series (see [ 15.38], [9.273] and [9.274], where the suitability of
repeated buckling as a means for deriving interaction curves, as well as the effect of sequence
of loading, are discussed).

15.2.8 VCT for Direct Prediction of Buckling Loads in Shells


The vibration correlation technique for determination of boundary conditions has also been
extended to direct prediction of buckling loads. Correlation using the same experimental
technique as for the definition of boundary conditions appeared promising for closely stiff-
ened shells since their low frequency vibration modes, observed in tests, were very similar
to their buckling modes (see for example [13.32], [15.28] and [15.41]).
The experimental curves of frequency squared versus axial load of a typical shell, when
the test is continued to buckling, exhibit a steepening at high loads with a rapid drop in
frequency before buckling (see for example Figure 15.27a for shell R0-34). The direct
prediction method is essentially a curve fitting to the experimental points, but using only

SHELL DUD-iO

Peons!"' 380 Kg

0 E,.;p1H1ment

0 0.03 0.06 0.09 0.12

Figure 15.25 Technion VCT experiments on integrally stringer-stiffened shells under combined load-
ing-frequency squared versus external pressure for shell DUD-10, in the presence of
constant axial load P,."'"' = 380 kg, when n = 8, m = 1 (from [9.273])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1272 Nondestructive Buckling Jests

0 UD 10

EXPERIMENT 6
SPRINGS 0
0.8

_SSL

0.6

0.4

0.2 O,L 0.6 0.8

Figure 15.26 Technion VCT experiments on integrally stringer-stiffened shells under combined load-
ing-interaction curves for shell DUD-10 (from [9.2731)

those points below, say, 50-60 percent of the buckling load, to make the method truly
nondestructive.
Already in the early attempts of correlation it became clear that curve fitting was easier
if the j2 curve was represented by a function
f 2 = (A - BPY:'./'1 where q > 2 (15.8)
Equation (15.8) can be written in another form:
f'! =A - BP, (15.9)
which has the advantage that the curve fitting is now for a straight line. Figure 15.27b shows
such a straight line for shell R0-34. It represents a least-square fit to the experimental points
to that optimal power, q"~'' for which the extrapolated line (P,.""'1,) yields the exact experi-
mental load (P,.,).
A parametric study was carried out [15.42] on all the integrally stringer-stiffened cylinders
tested at the Technion in axial compression (except those with significant load eccentricity
and those with insufficient experimental points in the relevant load range). The load range
for a truly nondestructive method was taken as 0.15 P, 1, - 0.6 P,f' for simple supports and
0.15 P,P - 0.5 P, for clamped ends, and the optimal exponent q was determined for all the
shells at the modes with the appropriate number of circumferential waves n. A study of the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

influence of the dominant geometric properties (the Batdorf parameter Z = (1 - u 2) 112 (LI
R? (R I h) and the stringer area ratio (A 1 I hh ), which were found to be the dominant properties
in earlier evaluations of buckling tests of stringer-stiffened shells) on the optimal exponent
of frequency q was then carried out.
From curves of optimal q versus Z(A bh) a functional relation for the exponent q was 1
/

found, for simple supports and for clamped ends,


q = a(ZA J bhY' (15.10)
where a and h are empirical constants. Figure 15.28 shows the actual use of the method as
a nondestructive technique for shell RS36, one of a test series specifically designed for
verification of the method. The direct prediction yielded P,.""'f' = 2820 kg, and the experi-
mental value was P, 1, = 2700 kg; hence, p,,1, 1, = 0.957.
In Figure 15.29 the buckling ratios for direct prediction Pnrmp of all the shells studied are
summarized and compared with the theoretical predictions for effective boundary conditions
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Vibration Correlation Techniques (VCT) 1273

RO- 3"
n = 11
m=l
o- EXP.

0
0
0 0

IOL00-------2-00L0_______3~0~00~-----4~00~0~i_p_K_g__
OOL________

m:l
q 0 P =2.9

0
0 1000 2000 3000 4000 Pt"xp2 Pe-xt

PKg

Figure 15.27 Technion VCT experiments on stiffened shells-frequency-load curves for development
of direct buckling load predictions (from [1.24] and [15.42]): (a) frequency squared
versus axial compression for shell R0-34 (n = 11, m = 1), (b) frequency to the power
q.," versus axial compression for the same shell R0-34 (n = 11, m = 1)

q -11
f xlO
[Hz]q

Figure 15.28 Technion VCT experiments on stiffened shells-direct buckling load prediction from
frequency to power q versus axial load for shell RS-36 (n = 10, m = 1, from [1.24]
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

and [15.42])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1274 Nondestructive Buckling Tests

Pextrop •
1.2 Psp 6

p
--- - - - - - - - - - - - - - - · - - - - - - - - - - - e--- - -

- = ----=-~-'----=·~~-=---=--
• ~ ~ -. =-=-- -. ~ =-= =-=--=---=
"' '!!'
0.8 t."" "'
"' "'"' "' "'"' "'

Figure 15.29
·1 100
·----

I
200
I
400
"' -b-

I
600
-
"'
-- -

800 1000
-- -

1200

Technion VCT experiments on stiffened shells-buckling load ratios for prediction of


"'
-

1400 z

buckling load with direct method and comparison with results from VCT determining
boundary conditions (from [1.24] and [15.42])

found with the vibration correlation technique p,r The scatter of Pwmp is found to be about
two-thirds that of p,". Similarly, in Table 27.6 of [13.41] Pnrmp for some more recent tests
is compared with the corresponding p,", yielding results fitting into the same pattern. How-
ever, since the functional relation used, Eq. (l5.l0), is empirical, many more tests in a
broader range of geometries are required for wider applicability of the method.
Similar VCT studies were carried out on Hostaphan cylindrical shells under axial com-
pression and external pressure by Radhakrishnan at the Indian Institute of Technology, Mad-
ras, in the early seventies (see [15.43]). The variation of frequency squared with load that
was plotted was not linear, but the extrapolated buckling loads agreed well with those ob-
tained from static tests.

15.2.9 Other Recent Vibration Correlation Methods


A good example of a more recent application of a modified VCT to plates is the experiments
carried out by Segall and G.S. Springer at Stanford University in the mid-eighties [8.83].
Their dynamic method for determining the buckling load requires vibrational excitation of
the plates, but no application of in-plane loads as needed in the usual VCT. The method is
an extension of an earlier dynamical method developed by Segall and Baruch [15.44] for
determination of the critical load of elastic columns. ·
That dynamic method differed from the usual VCT in requiring only experimental inputs
of the unloaded column (frequencies, mode shapes and masses) and therefore needing no
axial loading of the column. The change was achieved by an integral equation representation
of the elastic stability of a column proposed by Baruch [15.45], which was shown to apply
to the most general column and boundary conditions. For lateral vibrations, the influence
function G(x, g) of the column (which denotes the deflection at point x due to a unit lateral
load at point g) could be approximated by a converging series of measured normal vibration
modes of the column and its generalized mass. It was stated in [15.44] that in calculations
for the cases of simply supported, clamped-free and clamped-simply supported uniform col-
umns, two theoretical modes led to an approximation within 2 percent of the theoretical
buckling loads. This indicated that a few measured resonances and modes should suffice for
practical results, as was indeed found in the experiments.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,

The tests (see [15.44]) included one 900-mm long, constant cross-section 6061-T6 alu-
minum column in two configurations: (A) with soft springs at both ends, and (B) clamped
at one end and with a medium spring at the other, as well as one 910-mm long 2024-T3
aluminum column with linearly varying width and medium springs at both ends, configu-
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Vibration Correlation Techniques (VCT} 1275

ration (C). Each configuration was vibrated in the horizontal position with no axial load.
The first four resonances were determined and their corresponding mode shapes were
mapped, though the results showed that three resonances would have sufficed in these cases.
For comparison, the columns were also vibrated under incrementally increased axial load,
with the frame positioned vertically, in the usual VCT procedure. The results are compared
in Table 15.4 (reproduced from [15.44]), which also includes the predictions of two non-
destructive static methods, the Southwell method (see Section 4.5 and [4.12] and [4.40]) and
the lateral force method (see Section 15.3 and [15.46]).
For configurations A and B of the uniform column, the results of the two dynamic methods
in the table agree within 8 percent, and the two static methods also yield close results (except
the Southwell prediction for configuration B). However, for the nonuniform column, config-
uration C, the differences between the methods, even between the dynamic methods, are
significant! y larger.
The more recent application of the modified vibration correlation method (the normal
mode dynamic method) to flat plates [8.83] was more comprehensive. It consisted of exten-
sive numerical studies and six tests on two aluminum and one composite (graphite-epoxy)
square plates. Segall and Springer's analysis was essentially a two-dimensional expansion of
the earlier one for columns, and their numerical calculations centered upon a comparison
between the buckling loads computed by the normal mode dynamic method and those ob-
tained by classical plate theory for a range of plate configurations. The comparison (Table
1 in [8.83]) indicated that the predictions by the dynamic method were within a few percent
of the exact values, even if only one or two mode shapes were used.
The discussion here, however, focuses on the experiments. A schematic diagram of the
rigid steel test frame in which the plates were mounted is shown in Figure 15.30. Two steel
strips were installed into each vertical (unloaded) side of the frame. The test plate could be
clamped between these strips, approximating clamped boundary conditions, but it could also
be supported by the sharp knife edges of these strips, approximating simply supported bound-
ary conditions.
The lower and upper horizontal supports were made of aluminum and contained V-
grooves, which approximate simply supported boundary conditions. The upper support was
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
not fixed but could move vertically. Lateral motion of the upper support was prevented by
horizontal bars placed along each side of the support. The load was applied by a beam
attached to the frame by two bolts. The movement of the beam was transmitted to the upper
support through a 130-mm long aluminum bar inserted between the beam and the support.
This bar was in three-point bending during application of the load. A strain gage bridge was
mounted on the bar, from which the applied load could be determined. Two strain gages
were attached back to back along the centerline of each plate, about 50 mm below the upper
edge.
Lateral displacements of the plate were measured by an MTI model KD100 Fotonic sensor
with a KDllOl plug-in unit. The output of the sensor was fed into an HP 5423A structural-
dynamic analyzer. As can be seen in Figure 15.30, the optical probe could be positioned at

Table 15.4 Buckling loads (in kg) for columns calculated by


different nondestructive methods
Configuration Static Dynamic
Southwell Lateral force Normal
method method VCT mode
[4.12] [15.47] [15.14] [15.45]

A 33.6 35.6 36.5 33.7


B 75.4 87.9 90 96.5
c 87-93 119 97.2

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1276 Nondestructive Buckling Tests

STRAIN GA.UGE

LOAD MEASURING
SUPDQRT BAR ;
BAR

SLIDE
COLUMNS

0.3m

SIDE
SUPPORTS

Figure 15.30 Stanford University modified VCT (normal mode dynamic method) experiments on
square plates-test frame. schematic (from [8.83])

any point on the plate with the aid of a two-directional slide system. The advantages of
optical sensors are that they do not affect either the rigidity or mass distribution of the plate.
To determine the resonance frequencies,
the plate was impacted by a rubber-tipped stick (pencil) and the lateral deflections were measured
at several points across the plate by the optical probe. The output of the probe, processed by the
HP signal analyzer, provided directly the resonant frequencies. The plate was then excited by a
loudspeaker tuned to the required resonance frequency. Only those frequencies were used which
resulted in one-half sine-wave shapes in the horizontal (y, unloaded) direction of the plate. Fre-
quencies corresponding to other shapes were not used since, at buckling, the deflection in the y
direction is a half sine wave.
With the plate continuously excited by the loudspeaker, the vibration amplitudes (mode shapes)
were measured at predetermined grid points. In the present tests there were eleven and seven grid
points in the x (loaded) and y (unloaded) directions, respectively (m = 11, n = 7). Since mea-
surements were not made at points on the boundaries, the deflections were measured at 45 points.
The tests were repeated for up to three resonant frequencies (k = 1-3). The outputs of the two
strain gages mounted on the plates were monitored throughout each test to ascertain that the
conditions did not change during the test.

The HP analyzer could provide directly both the resonant frequencies and the correspond-
ing mode shapes if the exciting forces were known. But since in the present tests these forces
were unknown, a two-step procedure was necessary.
The only difficulties encountered . . . stemmed from technical problems associated with the
hardware. First, there were irregularities in the edge supports, causing changes in frequencies
measured during repeated installations of the same plate, and drift in the resonance frequency during
each test. This drift was limited to one percent to two percent during any particular test. Second,
the optical probe's sensitivity was affected by the plate's reflectivity, influencing the measured
mode shapes.

The buckling loads, determined by the normal mode dynamic method, are presented in
Table 15.5 (reproduced from [8.83]). To evaluate the results, these are compared in the table
with buckling loads of test nos. 1-6 determined statically by the Southwell method. The
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

plates were all in uniform compression, with the loaded edges approximately simply sup-
ported (V-grooves). Test nos. 3 and 4 differed in the configuration of the loaded edges of
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Vibration Correlation Techniques (VCT) 1277

Table 15.5 Buckling loads of 300 mm long and 290 mm wide plates measured by the normal mode
dynamic and by the Southwell method
Test Material Thickness Unloaded Frequencies Buckling load per unit Dynamic/
no. (mm) edge used (Hz) length (Nim) Southwell
support
Dynamic Southwell
I aluminum 0.8 S.S. 45.75; 112; 227 1,470 I ,450-1,560 1.01-0.94
2 aluminum 0.8 C. 48.8; 108.5 1,650 1,470-1,550 1.12-1.06
3 aluminum 1.6 S.S. 93.8; 226 13,100 12,200-13,000 1.07-1.01
4 aluminum 1.6 S.S. 105; 260 16,500 15,800 1.04
5 composite 0.8 C. 53.2; 143; 315 1,260 I, 160-1,180 1.09-1.07
6 composite 0.8 s.s. 45; 137.5 915 1,000-1,030 0.92-0.89

that plate, which were rounded in no. 3 and fiat (causing more rotational restraint) in no. 4.
The anisotropic plate of test nos. 5 and 6 was made of a cross-ply [0/90/0/0/90/0] graphite-
epoxy Fiberite T300/976. Note also that the Southwell buckling loads per unit length (and
therefore also the load ratios in the last column of Table 15.5) exhibit ranges that indicate
the spread in the measured static data.
The comparison in Table 15.5 shows that there was very good agreement, within about
10 percent, between the buckling loads obtained by the normal mode dynamic method and
by the static Southwell method, the dynamic prediction usually being higher. The modified
VCT, or normal mode dynamic method, therefore appears to have great potential as a non-
destructive technique for plates but still needs more experimental evidence. Future research
may perhaps also find it to be suitable for shells.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
A different vibration correlation approach was developed by Horton and his students at
Georgia Tech in the seventies [9.213]. They used the relation between the local dynamic
stiffness, or dynamic mass (ratio of shaker force to resultant acceleration) and axial com-
pression in the areas of weakness, identified by minimum dynamic stiffness (dynamic mass).
The method was first applied to an aluminum alloy column, in which the excitation input
acted at the column midspan. The shaking force and the resulting acceleration of the test
specimen were both measured at the excitation point, here and in the other tests. A plot of
dynamic mass variation with increasing axial load lead eventually to a linear relationship
between resonant frequencies and axial load, as in the usual VCT procedure. Extrapolation
of this line to the load axis yields the predicted buckling load.
Then the method was applied to an unstiffened rectangular panel, as well as to an unstif-
fened acrylic circular cylindrical shell and an unstiffened acrylic elliptical one (with a rec-
tangular cutout). To permit observations at a large number of stations in these specimens,
the shaker system was preloaded to maintain contact throughout the excitation cycle. By
comparison of the magnitudes of the dynamic mass, as the specimen was pointwise excited,
the weak areas of minimum dynamic mass could be identified.
When the minimum values of dynamic mass for a given frequency were plotted against
the corresponding axial loads, linear relationships resulted, whose extrapolations intersected
the load axis at virtually the same point-the predicted buckling load. Such a local minimum
dynamic mass versus axial compression plot for an acrylic unstiffened cylindrical shell is
shown in Figure 15.31. The extrapolations of the lines yield the predicted buckling load
5049-5293 N, which is 1-5 percent below the experimental one of 5316 N. The dynamic
mass predictions for the other tests were 5.3 percent above the actual buckling load for the
column, 2.3 percent below for the rectangular panel and 7.5 percent below for the unstiffened
acrylic elliptical shell with the rectangular cutout.
The results appear to be very satisfactory, but the reliability of the local minimum dynamic
stiffness method has, however, still to be verified by more extensive testing.
Before we leave this section, we should point out that a number of studies have been
carried out at different research centers, aiming at improved evaluation or reduction of the
measured VCT data (for example [15.47]-[15.52]).
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1278 Nondestructive Buckling 7i~sts

0.3

....
z"'
r>
.,., 0.2

...2
~ BUCKLING LOAD ( N)

...:::E
z ACTUAL
>- 5305
0
Q.j

...J
...
<J
0
...J

AXIAL COMPRESSION P( N)

Figure 15.31 Georgia Tech vibration correlation experiments-local minimum dynamic mass versus
axial compression plot for an acrylic unstiffened cylindrical shell (from [9.213])

For example, Souza et al. [15.47] and later Souza and Assaid [15.50] and Souza [15.51],
suggested a different straight-line technique than that employed in Subsection 15.2.8. Instead
of plotting the frequency to the appropriate power q; versus P and fitting the best straight
line, as employed there (see Figures 15.27 and 15.28), they proposed plotting, as shown in
Figure 15.32b for a stiffened shell R0-34 (of [13.32]), (1 - pf versus (I - j4), where
p = (PIP,,), f = (f,, j 0 ), P = the applied axial load, P,, = the critical load of the corre-
sponding perfect shell, !, = the measured natural frequency at a certain load and fo = the
natural frequency of the unloaded perfect shell. The critical load of the perfect shell P,,
could be calculated from the elastic and geometric properties of the shell with its correct
boundary conditions (which could correspond to P,1, of Subsection 15.2.5) or obtained from
a best-fit straight line to the squares of the natural frequencies at low load levels, as in Figure
15.32a.
Souza and his co-workers employed the latter alternative and obtained slightly high values
for P,,. For example, for Technion shell R0-34 their P,, = 6409 kg, whereas the former
alternative of P,, = P,P yielded 5288 kg (see [13.32]); or similarly, for shell R0-32 their
P, = 7857 kg, whereas P,., = P, = 6220 kg, a difference of 21-26 percent.
The procedure proposed by Souza et a!. then consisted of measuring the natural frequency
!, at various values of applied load P and plotting (1 - p? versus (1 - j4) as shown in
Fig. 15.32(b). The parametric form of (1 - p) 2 versus (1 - f 4 ) originated in the represen-
tation of the cylindrical shell by a simplified mode, consisting of a strut and arch connected
at their center and loaded along the axis of the strut, employed in an earlier analysis by
Souza and Walker [15.52]. The value of (1 - p) 2 corresponding to (1 - j4) = 1 (i.e. when
r --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley = 0) would represent the square of the drop of the load carrying capacity,
Provided by IHS Markit under license with WILEY
due to the e.
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Vibration Correlation Techniques (VCn 1279

CJ
10
EXPE~IMENT
e RO 34

06

06

0.4

0.2

(a) 0
0.1 0.2 0.3 0.4 O.S 0.6 07 O.B 0.9 1.0
p

1.0

e2

0~

(b) ( 1-1 4 )
0.5 1.0

Figure 15.32 Souza's straight-line technique for nondestructive prediction of buckling loads applied to
data of Technion stiffened shell R034 (from [ 15.47]): (a) estimation of P, from best-fit
straight line to squares of natural frequencies at low load levels, (b) plot of (I - p) 2
versus (I - J"'), yielding the drop of load carrying capacity of the imperfect shell e.
from which the buckling load parameter p, is obtained

initial imperfections. For shell R0-34, Figure 15.32b yields = 0.09 and therefore the e
buckling load parameter p" = 1 - V0.09 = 0.70. Thus, the predicted buckling load would
be P;, = 0.70 P,., = 0.70 X 6409 = 4486 kg, if Souza's value of P" was used. Since the
actual buckling load observed in the test was P,, 1, = 4220 kg, the buckling load ratio was
p1, , " = (4220/4486) = 0.94.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

If, however, the calculated value of Per = P,1, (which includes the vibration correlated
boundary conditions) were used, the predicted buckling load would have been P" = 0.70 X
5288 = 3702 kg, and the corresponding buckling load ratio p1,e" = (4220/3708) = 1.14.
For comparison, Singer's extrapolation straight-line technique (of Subsection 15.2.8 or
[1.24] and [15.42]) yielded for shell R034 Pn""'' = 4232 kg, and their corresponding buck-
ling load ratio was Pntmp = 0.997.
A similar application of the Souza et a!. straight-line technique to Technion shell R0-32
(see Figure 15.33 ), yielded a drop of load-carrying capacity = 0.086 and therefore P~> = e
1 - \10.086 = 0.707. With Souza's value of P,,, P" = 0.707 X 7857 = 5555 kg. Since the
buckling load observed in the test was P"l' = 4 700 kg, the buckling load ratio was p1,""1 =
(4700/5555) = 0.846. If, instead, the calculated value of P". = P,l' = 6220 kg were used,
Ppm!= (4700/4400) = 1.07. For the same shell Singer's extrapolation straight-line technique

yielded P,, 1, " = 5159 kg and hence p"''"'" (4700/5159) = 0.911.


Souza and his colleagues at University College, London, and at the Catholic University
of Rio de Janeiro used the data of only two of the many Technion shells and unfortunately
did not carry out any further tests. Hence, extensive additional studies are required to bolster
their suggested VCT method.
Another example of improved evaluation and reduction of measured frequency for non-
destructive determination of buckling loads is the recent investigation of Plaut and Virgin at
Virginia Polytechnic Institute [15.49]. Their aim was to extend Singer's extrapolation pro-
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1280 Nondestructive Buckling Tests

fl
Experiment f2 : 1·05 • 10'6 IHzJ 2
J.O • R0-32 pcr =7857kg
08

0·6

Q-1.

0·2

(a) 0
0 0·2 0·6 0·8 1·0

{1 -~1 1
1·0

OS

(b) L--------'o--- - - - - - "_.,... - 11-f'J


5 10

Figure 15.33 Souza's VCT applied to data of Technion stiffened shell R0-32 (from [15.47]): (a)
estimation of P from squares of natural frequencies at low load levels, (b) plot of (1
0
.

- pf versus (1 - f"), from which drop of load-carrying capacity t' and buckling load
of parameter p" are obtained.

cedure for buckling load predictions [1.24] and [13.41] in order to facilitate also the esti-
mation of lower and upper bounds.
Based on an analysis of a shallow elastic arch with pinned ends, they suggested the
following procedure. For each measured frequency !,, one should compute and plot f'~
versus p (the frequency ratio, f = f,lf 0 , to exponent q versus the axial load ratio p = PI
Pc,.), for q = 1, 2, 3 . . . . To keep the method nondestructive, values of f, should be
measured for loads up to not more than half the roughly estimated buckling load. Then one
should connect curves through the plotted points and extrapolate linearly (i.e. draw the
tangent at the last data point) till f = 0. Examining the connecting curves, f 2 versus p, one
should characterize them as being either concave or convex (with respect to the origin). If
the curve is concave, the value of p at f == 0 can be expected to be an upper bound on p",
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

and if convex, p at f = 0 will probably be a lower bound, as shown schematically in Figure


15.34, where the predicted p 1 and p 2 bracket the actual buckling load parameter p".
In Figure 15.35 Plaut and Virgin's suggested procedure is carried out, with q = 1, 2, 3,
4 and 5, for the vibration data of shell R0-34 (from [15.35], in mode n = 11, m = 1). The
values off, are taken up to P = 2400 kg, which is nearly 112 Per, and the expected trend
does indeed appear. Extrapolations for q = 4 and q = 5, which are convex, yield lower
bounds to the experimental buckling load of 4220 kg. The other extrapolations result in
upper bounds, though only for q = 1 can the curve be clearly identified as concave.
A somewhat similar assessment was carried out in the early stages of the Technion VCT
studies for direct buckling load prediction, when (in Figure 13 of [13.32]) curves of the
experimental values of f, for shell R0-34 raised to different powers were plotted. The
intersections of the f'~ extrapolations with the abscissa resulted there in a similar bracketing
of the experimental buckling load by upper- and lower-bound predictions.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Vibration Correlation Techniques (VCT) 1281

Figure 15.34 Plaut and Virgin's procedure schematically: the curves of f" = f 0 )'1 with q 1 are
concave (with respect to origin), yielding upper bounds like p 1 , whereas those with q,
(f,,
are convex, yielding lower bounds like p 2

Very few nondestructive VCT buckling experiments for plates and shells have been carried
out in industry. An example for shells is the tests carried out by Schneider et al. at Mc-
Donnell-Douglas Space Systems Company, Huntington Beach, California [15.53]-[15.55]).
They conducted experiments on buckling under external pressure, vibration and the inter-
action of external pressure and vibration on a sixth-scale Lexan (a polycarbonate plastic)
model of a Titan IV launcher nose cone fairing.
Previous experience at McDonnell-Douglas with Lexan models had shown that repeatable
buckling loads could be obtained from a single model, provided that postbuckling defor-
mations were restricted to prevent yielding of the material. This was assured here in all tests
by the use of an internal constraint fixture (a mandrel) limiting the deformations. Hence,
repeated buckling was possible, and the model could be buckled first and then vibrated for
modal survey.
The model experiments had three objectives: (1) to assess the analytical methods used in
the design, (2) to investigate experimentally important design features like the influence of
discontinuous rings (due to the explosive separation system), and (3) to experimentally study
a proposed vibration-based diagnostic technique, their version of VCT. Naturally the dis-
cussion here will focus on the third objective.
The test setup for the 1081-mm high scale model is shown in Figure 15.36. The nose
cone is a 15/25° biconic shell. For the vibration tests, to determine the shell mode shapes
and frequencies, 42 accelerometers were placed at equal distances around the circumference

fq
1.0

-- - - - -
.5
------ - -
.........

--
.........
.........
......... -........
.........
......... -........
.........
......... .........
.........
..................
.........
.........
1000 2000 3000 4000 t p(kg)
Pexp
Figure 15.35
Virgin's procedure, plotting frequency ratio f =
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
(f,,
Technion VCT experiments on stiffened cylindrical shells-application of Plaut and
f 0 ), to the power q versus axial
compression, for shell R0-34 with q = 1, 2, 3, 4 and 5 (from data of [15.36], n = 11,
Copyright Wiley m = 1), and extrapolating linearly
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1282 Nondestructive Buckling Tests

Figure 15.36 McDonnell· Douglas. Huntington Beach, VCT experimellls on a one-sixth-scale Lexan
model of the Titan IV nose cone fairing- test setup (from [15.53])

to identify the number of circumferent ial waves (N) and 7 accelerometers at equal meridional
distance along the !5° conical shell to determi ne the axial waves. For excitation, a 20 N
force shaker was suspended on a bungee cord and auached to a load cell through a 0.8-mm
diameter stinger, the input excitation force being provided by a random noise generator.
Mode shapes and frequencies were measured first for the unloaded condition and then at
pressure increments.
The frequency data obtained were presented as frequency squared versus external pressure
curves for different mode shapes (actually different numbers of circumferential waves N) .
The experimental data were compared with pred ictions by NASTRAN (finite elements-
[2.98)) and BOSOR 4 (fini te differences-[2.62)) codes and good agreement was obtajned.
The J;., versus p curves (l ike Figure 15.37) were then evaluated as a nondestructive tech-
nique for prediction of the buckling load. actuall y VCT. or as Schneider and his colleagues
called it. " vibration diagnostic technique." They observed (see Figure 15.37) that the ex-
perimental frequency squared versus pressure data presented linear relatjons as long as the
applied load remained below 80 percent of the buckling load. Beyond that, the interaction

15 .000
-e- N=J
~N =4
--+- N=5
-e- N• 6
"';:;
!.
N.,_Z
10,000
0
"''"
.,;
~
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

6-
<J>
g s.ooo
~
.."'
!

Prtssurt. p (pol}

Figure 15.37 McDonnell-Douglas . Huntington Beach. VCT experiments on a one-sixth-scale Lexan


model of the Titan IV nose cone fairing- experimental frequency squared-pressure
Copyright Wiley
curves for continuous Z-ring stiffening (from [15.53], courtesy of Mr. M.H. Schneider)
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Vibration Correlation Techniques (VCT) 1283

curves very rapidly rolled away from the linear ones, presumably due to geometric imper-
fections of the shell, and intercepted the pressure axis below the linearly predicted buckling
pressure. One should recall that the most important curve in Figure 15.37 was that for N =
5, which was the buckling mode shape.
The McDonnell-Douglas investigators therefore concluded, as observed earlier by the
Technion researchers (see Subsection 15.2.8), that extrapolating linearly only the squares of
the natural frequencies taken at low loads would result in an unconservative buckling load
prediction. They pointed out, however, that one can benefit from the linear frequency
squared-pressure curves of the model to establish a stop-test criteria, which applies also to
the corresponding full-scale test. Such "criteria were successfully applied to the Titan (nose
cone) fairing prototype test, which utilized the vibration diagnostic technique (VCT) to alert
the test direction if general instability failure were imminent. At each load step the (full
scale) structure was excited by a small shaker while the pressure was held constant."
Furthermore, one may obtain from the vibration behavior useful information on the influ-
ence of design changes, like the cutting of the Z-rings for the explosive separation, though
the frequencies are not as sensitive to design features as is the buckling behavior. For ex-
ample, here, in the nose cone models, the frequencies for the discontinuous rings were only
5-l 0 percent below those for the continuous rings, whereas the buckling load was reduced
by 30 percent. Additional studies might perhaps yield better methods for correlation between
vibrations and buckling as indicators for the effect of design changes.
A further experimental study on the vibrations of ring-stiffened conical shells under ex-
ternal pressure (in a water tank) was carried out at the University of Portsmouth [15.57].
Though the study was not a proper VCT experiment, but only a series of vibration tests
under destabilizing load, it included the development of a variable inductance noncontact
transducer for monitoring the vibration characteristics, which operated satisfactorily. The
Portsmouth tests may therefore provide useful information for future VCT investigations.

15.2.10 The Status of VCT


In view of their importance and potential, it may be worthwhile to summarize briefly the
status of vibration correlation techniques, particularly with respect to plates and shells.
For plates, both the usual VCT (discussed in Subsection 15.2.3) and the more recent
modified VCT, or normal mode dynamic method (discussed in Subsection 15.2.9), appear
to be working well. Though some additional tests have been reported (as for example in
[15.56]), further experimental studies are warranted to bolster confidence in these methods.
For shells, the extensive Technion studies on the correlation between vibration and buck-
ling of stiffened shells (discussed in Subsections 15.2.4-15.2.7) have yielded a nondestructive
method for improved definition of boundary conditions. The vibration correlation technique
(VCT) has been extensively tested and proven a reliable tool on a laboratory scale. The
technique has been extended to different types of practical boundary conditions and modified
to detect the presence of significant load eccentricity and account for it once it has been
identified. It has also been applied successfully to larger shells of ditierent constructions.
VCT has also been extended to other loading cases for stiffened shells, external pressure
and combinations of external pressure and axial compression, and the applicability of the
method has been verified in a substantial test program. A nondestructive method for deri-
vation of interaction curves has also been developed.
Measurement of initial imperfections and calculation of their degrading effect on the buck-
ling load (discussed in Chapter 10) is another approach to improved predictions. When
combined with more precise definition of the boundary conditions by the vibration correlation
technique, it has been shown to yield good agreement between prediction and test for shells
as well.
The experimental technique of vibration correlation has also been extended to direct pre-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
diction of buckling loads from vibration tests, yielding a semi-empirical method that has
proven reliable within the range of geometries tested (discussed in Subsection 15.2.8).
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1284 Nondestructive Buckling Tests

VCT can therefore be considered a practical tool for improved buckling predictions of
stiffened shells. However, most studies on buckling and vibrations of shells have been con-
cerned with uniform loads, whereas in practice shells are usually subjected to nonuniform
loading conditions. Research should therefore be directed to developing the nondestructive
test methods, and in particular VCT, to shells with nonuniform loading and in particular with
large openings. For the case of nonuniform loads, VCT could perhaps be complemented by
static NDT methods.
Most of the recent VCT studies for shells (discussed in Subsection 15.2.9) used earlier
data and did not produce additional tests that would verify and reinforce the VCT methods.
These methods indeed have great potential, but they still require some further experimental
support for widespread use by industry.

15.3 Static Nondestructive Methods


15.3.1 Experimental Determination of End Fixity
As mentioned in Section 15.1, the earliest practical NDT for buckling was the Southwell
plot, which is a static method. Thus, it also represents the first expedient static nondestructive
technique for determination of the buckling load of columns. The Southwell method is
discussed in Chapter 4 for columns (including also its NDT uses), while its applications to
plates and shells are detailed in Chapters 8 and 9, respectively.
The end fixity of a column, or its corresponding effective length, is one of the major
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

design parameters that determine the buckling load. The correct assessment of its actual
value for an imperfect column has therefore been of great concern to structural engineers
since the beginning of the 20th century. For example, in 1921 E.H. Salmon in his monu-
mental thesis on columns [4.3] wrote: "The most pressing point for future research on the
subject of columns is undoubtedly the degree of imperfection common in practical fixed
ends; ... at present the designer has no real data whatsoever regarding practical end con-
ditions."
Hence, considerable efforts have been devoted to experimental determination of end fixity.
Empirical data have been accumulated from tests on realistic structures, and application of
the Southwell method to these tests assisted in the estimates of the end fixity. Indirectly, the
Southwell plot could also be used to assess the end fixity of a particular test setup, but its
nondestructiveness was not always ensured. Safer nondestructive methods were therefore
sought.
For example, in the forties Schuette and Roy at NACA Langley [15.58] developed a
method for the experimental determination of the effective length of a column by establishing
the points of zero curvature from readings of strain gages along the length of the column.
They measured the curvatures (obtained from the strain differences divided by the distance
between the gages) from pairs of wire strain gages, placed opposite each other or at other
convenient locations on the same cross-section, and found the points of inflection of the
column from the plots of curvatures along the length, for different axial loads.
Columns with four typical types of cross-section were tested. Satisfactory results were
obtained, even when there was considerable scatter in the strain gage readings, and the
buckling loads calculated with the experimentally determined effective lengths were within
5 percent of the test values. However, in order to have measurable curvatures and eliminate
excessive scatter, the strain measurements had to be carried out at loads near the maximum
(from 92-95 percent on). This raises doubts about the nondestructiveness of the method.
In the sixties and seventies different approaches were taken to develop static nondestructive
test methods to obtain end fixity and buckling loads. For example, Horton and his co-workers
at Georgia Tech [15.59] obtained empirical correspondence rules between the critical axial
load of slender beams and the deflection due to a lateral midpoint load. They analyzed a
variety of columns with different end fixities and also carried out a series of tests, in which
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Static Nondestructive Methods 1285

the end restraints could be varied without removing the column from its fixture. However,
though it worked well in some cases, the method does not have sufficient general applica-
bility.
Another nondestructive method, developed by Baruch [15.46], also employs a lateral load,
but is applicable to elastic columns with all types of end restraints. Using this method for a
beam of constant stiffness EI, one applies to it a lateral load Q = 1 (at x = a and z = b)
and measures the support deflections v0 , v L and rotations ¢ 0 , ¢ 1 (see Figure 15.38). With
these measured quantities the support reactions A0 , A 1. and end moments M0 , M 1. can be
computed from Eq. (15.11):
1 1 0
a -b -1
(15.11)

r (a 2 /2)
(a' I 6)
W/2)
-(b 3 /6)

Note that repetition of the measurements of v0 , v v ¢ 0 , and ¢ 1 for different values of Q and
averaging them for Q = 1 will yield more reliable data.
With these reactions and end moments, the spring constants can be evaluated, since
a 0 = (M0 / ¢ 0 ) a 1. = (M1l c/JL)
(15.12)
/30 = (A 0 /v 0 ) /31 = (A 1lvJ
The buckling load for the column with the elastic restraints of Eq. (15.12) is the smallest
P for which the following determinant vanishes.
a 0k ElF
(a 1k cos kL - Elk2 sin kL) -(a 1k sin kL + Elk2 cos kL)
= 0
0 f3o
-f3L sin kL -f3L cos kL
(15.13)

where (PI El) = F.

Figure 15.38 Deflections, rotations and reactions of a spring constrained beam, loaded by a lateral
force Q = I --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1286 Nondestructive Buckling Tests

The method is nondestructive since the values of Q can be kept well below those that
would result in inelastic stresses. Examples of buckling load calculations with this lateral
force method for two uniform columns are presented in Table 15.4 of Subsection 15.2.9,
where the results are compared both with those obtained by the static Southwell method and
by two dynamic ones, the usual VCT (of [15.14]) and the normal mode technique (of
[15.44]).
The method was extended to columns with an unknown bending stiffness El (see [15.60]),
when an additional measurement, say the deflection u, at point 3, was needed, as well as to
alternative locations of measurements, away from the support, say at points 1, 2, 3, 4 and 5
in Figure 15.38.
As mentioned in Subsection 15.2.9, another approach was proposed by Baruch at Technion
in the early seventies, an integral equation representation of the elastic stability of columns
and plates [15.45]. In this method the kernel of the relevant integral equation is found by
measuring the deflections and rotations of the beam or plate, caused by a unit force at some
given point of the structure. As indicated earlier, a typical static application of the method
to define the elastic end restraints, or end fixity, of a column (lateral force method) was
briefly described in [15.44]. There also the corresponding dynamic application was shown
to permit modification for VCT to the normal node dynamic method, in which only exper-
imental inputs of the unloaded column were needed, as pointed out in Subsection 15.2.9. A
similar dynamic application of the method to a plate was also discussed in that subsection.
A different approach was proposed by Becker in the late sixties [15.61], in which the
change of transverse bending stiffness (due to a lateral force Q) with axial compression,
measured at relatively low loads, was studied and extrapolated to the load at which the
stiffness vanished. The proposed procedure was applied to the data of an early Caltech
GALCIT bending instability test of a metal cylindrical shell [15.62].
From an analysis of a long cylinder, loaded in axial compression and subjected to a ring
of uniformly distributed radial inward forces, Q per unit circumference, a linear relation was
obtained between stiffness squared and axial load,

(15.14)
where S = (Q/ 8) and S0 is the stitiness when P = 0. This relation was assumed to hold
also for bending, i.e.
(15.15)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

The experimental data are plotted in Figure 15.39, and indeed for this test the buckling

2 20 • EXPERIMENTAL DATA FROM


4 GALCIT TEST (RE'f. 15.62 )
( f)•10-
(lb /in )2 15

AVERAGE
/CURVE
10

·~ "(0

0~--~~--~--~--~~--~~--~--~~·~~-
0 so 100 150 200 250 300
Mx 10- 3 (in-lb)
...

Figure 15.39 Bending general instability of a stiffened cylindrical shell, stiffness squared versus applied
and moment for experimental data from an early Caltech GALCIT test (from [15.61])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Static Nondestructive Methods 1287

moment obtained from extrapolation agrees very well with the experimental failure moment.
One should, however, point out that it is not safe to conclude from one test that the relations
(15.14) or (15.15) apply more generally, as has been erroneously assumed in the literature!
The variation of lateral stiffness with compressive axial loading in cylindrical shells was
also studied by the researchers at Georgia Tech in the seventies [9.213] and [9.219]. The
static stiffness method developed by them was an extension of the lateral stiffness method
used earlier on columns (for example [ 15.59]). It related the variation in the shell wall lateral
(or normal) stiffness as a function of the applied compressive force to the buckling load.
Lateral stiffness was determined at growing levels of axial load by a stiffness probe applying
small normal force, and the circumferential distribution of this stiffness was plotted at a
number of axial stations.
Experimental measurement of lateral stiffness was made with the apparatus shown in
Figure 15.40. Its primary function was to apply a continuously adjustable and measured
point force normal to the shell wall and to determine the resulting deflection at the point of
application. A small blunt probe was attached to a traverse table, and the load transmitted
through it to the shell wall was measured by a sensitive force transducer. The traverse table
was positioned by a micrometer drive, and the motion relative to the shell axis was measured
by an LVDT.
Since the force transducer had a compliance (flexibility) of the same order of magnitude
as the shell wall, it could not be neglected; but because it was constant, it could be isolated
from the compliance of the shell, whose measure could thus be corrected automatically. The
system therefore produced plots of the lateral point force Q versus the deflection 8, at spec-
ified axial loads, from which stiffnesses or compliances (fiexibilities) could be computed.
Weak regions were identified and finally lateral stiffness-axial compression plots, at
regions of least stiffness, were extrapolated to yield the predicted buckling load. Craig and
Duggan [9.219] obtained an excellent prediction, within less than l percent of the experi-
mental buckling load, for a 284-mm (11.2-in.) diameter acrylic (methyl methacrylate) iso-
tropic cylindrical shell (R/t = 187) they tested. But a single specimen is obviously insuffi-
cient for such an essentially empirical method.
Horton et al. [9.213] tested six cylindrical acrylic shells of different construction (three
unstiffened and three stiffened, four circular and two elliptic, with R It ~ 180-280). Five of
the six shells yielded predictions within 6 percent of the experimental buckling load, but for
one, a spirally stiffened cylindrical shell, the prediction was about 40 percent too high.
Hence, though it is promising, more tests are required for the method to be acceptable.
A lateral stiffness survey was also adopted by Wilkins and Love [14.203] to define low-
stiffness areas in their graphite-epoxy and boron-epoxy cylinders. They automated their data
processing to yield the stiffness surveys, but did not employ them for buckling prediction.
They then reverted to load-strain curves, using what is essentially a strain-reversal method,
for indication of imminent collapse. They also automated their load-strain plots and used

Shell Wall

Micrometer Feed

Displaceme::r.t
Transducer ( LVDT )
--`,`,`````,`,````,,``,``,,`-`-`

Figure 15.40 Georgia Tech static-stiffness method applied to a cylindrical shell-sketch of lateral
stiffness measurement apparatus (from [9.219])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1288 Nondestructive Buckling Tests

them successfully for buckling load prediction, permitting the repeated buckling required for
load-interaction curves. But since the load has to approach the buckling failure closely, as
in a Southwell technique, their method cannot be considered a true nondestructive technique.

15.3.2 Force I Stiffness Techniques


The most widely used static nondestructive method is probably the force/ stiffness technique
developed by R.E. Jones and B.E. Greene in the mid-seventies [4.29]. This technique is

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
essentially similar to the 1938 Donnell plot (see Figure 4.14d), which he proposed [4.26] as
an alternative interpretation of the Southwell plot, or the basic hyperbolic law.
To rederive the characteristics of the force/ stiffness (F IS) plot, one may follow Jones and
Greene [4.29] and consider a slender column, simple supported at both ends. If the column
has a small initial curvature, which results in an initial lateral displacement 80 at the center
of the column, the additional center deflection 8 due to an axial load P can be written
8 = 8J[(Pe,J P) - I] (15.16)
where P,, is the Euler load. Equation (15.16) can be rewritten as
(P/8) = (Pe,- P)l8o (15.17)
Hence, a plot of (PI 8) versus P, the F IS plot, is a straight line of negative slope intersecting
the P axis at Pa (see Figure 15.41). Note that the F/S plot is actually the Donnell plot (of
Figure 4.14d in Vol. I) with ordinate and abscissa interchanged.
The F IS plot is usually obtained from strain gage data. The bending moment at the
midspan of the column is given by
(15.18)
and the corresponding bending strain is
t:" = kM = kP( 8 + 80 ) (15.19)
where k is a constant, which depends on the material and cross-sectional properties of the
column. From Eq. (15.16) one obtains the total center displacement
(8 + 8o) == 80 /[1 -(PIPe,)] (15.20)
and substitution in Eq. (15.19) yields
(Pis")= [I - (PIP,.,.)]Ik8 0 = -(Pik80 P") + (l/k80 ) (15.21)
Again the F IS plot, here P le" versus P, is a straight line, intersecting the load axis at P,.,
(see Figure 15.41). One should recall that s 1, is the bending strain, such as obtained from
the difference between back-to-back strain gages.

\--~classical buckling
\ behavior

\
V \
\
'yekol "''"'"'
point of test

\ predicted
\ buckling
----~load
p PCR

Figure 15.41 Force/ stiffness plot for general instability. P 18 versus P or PIs, versus P
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Static Nondestructive Methods 1289

If maximum total strain (such as that obtained from a single strain gage) is used, the total
strain is given by:
s = kP(o + 80 ) + (PIA£) (15.22)
Substituting for the total center displacement from Eq. (15.20) yields (after a few steps) the
FIS form

(Pis)= [ l - (PIP,,.)]I{k8 0 + {[l- (PIP,,.)]IAE}} (15.23)


The resulting F IS plot is not linear anymore, but a curve, concave downward, that again
intersects the load axis at P One may therefore conclude that also in more complex buck-
0
.•

ling problems an F IS plot that is inclined downward to the right is an indication of ap-
proaching buckling.
In a structural component the failure may occur either by buckling or by local failure,
which could be material failure or a local crippling instability. Supposing that, either ana-
lytically or by tests of coupon specimens, a strain level has been determined at which this
local failure will occur, this limiting strain s" can be represented conveniently on an FIS
plot as a straight line, with slope II s", passing through the origin (as shown in Figure
15.42). In a test, failure is indicated by the intersection of the FIS plot and the limiting
strain line (point A in Figure 15 .42).
Since local failure is determined by total strain, the F IS plot representing buckling be-
havior is curved, as was apparent in Eq. (15.23). However, the plot can usually be extrap-
olated to its intersection with the limiting strain line to yield a reasonably accurate estimate
of the local failure load.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

When local failure occurs under combined loading, as in the case of a structural panel
loaded in compression, bending and shear, one has to monitor a general biaxial strain instead
of the unidirectional one. The strain in the F IS plot has then to be a generalized strain
variable, derived from the local failure criteria, which are often expressed as strain interaction
equations. These must be based on adequate test data, obtained from material coupons, local
buckling specimens or crippling specimens.
Though essentially similar to the Southwell plot, or Donnell's version of it, the application
of the force stiffness plot is more convenient in those cases when the F IS plot, or the
corresponding Southwell plot, becomes significantly nonlinear, since the extrapolation of the
FIS plot is easier. This is exemplified by Figure 15.43 (a simplified version of Figure 3 of
[4.29]). which shows a typical set of such test data. Points A-E indicate possible load levels
at which the test can be stopped. The classical buckling load is predicted by extrapolation
of the straight line portion of the F IS plot. After point C the plot curves down, usually at a
faster rate as the load increases. Extrapolations from the higher load levels D orE, by tangent
straight lines, or preferably curved continuations, will yield successively better predictions

PREDICTED
FAILURE

Figure 15.42 Limiting strain line, representing local failure, on a force/stiffness plot. The strain here
is the total strain
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1290 Nondestructive Buckling Tests

2.0

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
8

1.0
c

0.5

OL-----~------L-~~~~--~~~
o p 10 flo p0 15 f!:
Figure 15.43 Force/ stiffness plot with significant nonlinearity-the process of successive extrapolation
for prediction of the buckling load

of the failure load ( Pn or PE), as is clear]y seen in the figure. One can also easily program
a code that will present to the test engineer successive real-time estimates of the failure load.
It should be noted, however, that for good predictions the test has to be continued to load
levels close to the actual failure, which might endanger the nondestructiveness of the method,
as in the case of the Southwell plot.
Jones and Green also studied the PIS signatures of different behavior types (see [4.29])
and indicated how they could assist in data interpretation of tests of structures whose buckling
behavior was more complex. They also applied the F IS plot in extensive design tests of
many beaded and tubular panels, both full-scale (about 1 X 1 m) and smaller specimens,
tests whose purpose was to guide the choice of structural configuration. The full-scale panels
were each tested in 10 different load combinations to obtain nondestructive failure load
predictions, both for local and general instability.
Figure 15.44 shows a typical example of a beaded compression panel for which the pre-
dicted general and local instabilities occurred at about the same load levels. The stiffness
axis in this and similar figures is identified by F I D, where D represents bending or total
strain. The predicted failure loads appear in parentheses, with local instability occurring at
the intersection of the curve of stiffness based on total strain (at location 4) with the limiting
strain line, while general instability appears where the extrapolation of the curve of stiffness
based on the bending strain (at location 2) cuts the load axis. Both predictions were very
close to the actual final measured failure loads.
An example of a tubular panel, loaded in combined compression and shear and tested to
destruction, is shown in Figure 15.45. Here the local instability was the critical failure load,
and again the FIS prediction was very close to the final measured failure load.

15.3.3 Further Application of Force/Stiffness Methods


In the last two decades the FIS plot has been widely applied as a nondestructive test method
both in industry and in research centers (see for example, [ 15.63]-[15.67]).
The tests carried out at the Centre d'e>.sais aeronautique, Toulouse, France, on a Mirage
2000 wing [15.63] are an example of an application of FIS plots to a practical aerospace
structure in the late seventies. The method worked well, though the nondestructive buckling
load predictions for the relatively stiff wing structure were found to be somewhat higher
than the finally observed failure loads. The author of the paper, Finance. attributed these
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Static Nondestructive Methods 1291

20

- rJf2B ~
~SG~~;\_) '--:J
SG2A
STRAIN GAGE LOCATIONS
16

14

12

F/D
10

FAILURE LOADS:
F/S 36 K
ACTUAL 35.9K

10 15 2~ 25 30 35
F- KIPS

Figure 15.44 Boeing Aerospace Company nondestructive experiments on aluminum panels using F IS
plots-a beaded compression panel for which the predicted general and local instabilities
were at about the same load levels (from [4.29])

differences to local moments and stresses already deep in the plastic region, which occurred
close to the failure load and accelerated the growth of deformation, that lead to failure.
Another example is the buckling tests carried out on an integrally stiffened, integral wing
fuel tank by Weissberg and Baruch at Israel Aircraft Industries at about the same time
[15.64]. They concurred with Jones and Green that F/S plots were more convenient than

350
LOAD CONDITION
Nxy"' 1/3 Nx

FAILURE LOAD:
300
ANALYSIS 197 KIPS
F/S 202 KIPS
ACTUAL 210 KIPS

250

FID
R •1.34''

1~5"
200

150

100

50

0
o~--~5~0----~I~00~--~1~50~----200~----~250
COMPRESSION LOAD, F- KIPS

Figure 15.45 Boeing Aerospace Company nondestructive experiments on aluminum panels using F IS
plots-local buckling F IS plot for a tubular reinforced panel loaded in combined com-
pression and shear and tested to destruction (from [4.29])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1292 Nondestructive Buckling Tests

the usual Southwell plot and employed them successfully for prediction of both local and
general instability. A typical F IS plot for a local buckling test is shown in Figure 15.46.
(Note that Figures 15.46 and 15.47 arc copies of the graphs obtained from the computer in
real time during the tests.) In this test the specimen was loaded only longitudinally without
any lateral restraint. The predicted local buckling stress, obtained from the F IS plot, was
a, = 24.5 kglmm 2 . Figure 15.47 shows an FIS plot for general instability under combined
bending and pressure loading. The extrapolation of the F IS plot here predicted general
buckling at a reference compressive stress of 31.7 kglmm 2 , whereas in the final destructive
test general instability occurred at a reference stress of 30 kglmm 2 , an acceptably small
difference.
One may note in Figure 15.47 the change in buckling mode at a reference stress of 20
kglmm2 . The designations mode I and mode III in the figure refer to predicted mode shapes
from finite-element (NASTRAN) calculations. In the final destructive test the structure indeed
tended initially to buckle in the predicted mode III, which would have resulted in a critical
compressive stress of 100 kglmm 2 if the mode had not changed to one that permitted buck-
ling at a much lower stress.
A more recent example are the nondestructive buckling tests on Rene 41 tubular panels
for a hypersonic aircraft wing, carried om by Ko and his co-workers at NASA Dryden Flight
Research Facility, Edwards, California, in the mid-eighties [15.65] and [15.66].
One of the concepts for future hypersonic aircraft was an aerodynamically acceptable wavy
heat shield made of a heat-resistant material (such as Rene 41, a nickel alloy) designed to
limit the structural temperature to about 730"C ( ~ 1350"F), a so-called hot structure. Due to
the combined effects of thermal stresses and air loads, such hot structures are prone to
buckling, as discussed in Chapter 19. Because curved shell sections exhibit high local buck-
ling strength, most of the hot structural panel concepts employ curved surfaces. Two of these
promising concepts were beaded and tubular panels, encountered in the previous subsection.
To characterize the buckling behavior of tubular panels in a realistically heated wing
structure, five Rene 41 noncircular tubular panels (see Figure 15.48) were attached to the
wing root region of the NASA hypersonic wing test structure (HWTS) for extensive non-
destructive buckling tests under different combined loading conditions (axial compression,
bending under lateral pressure and shear) and different temperature environments of 21 "C,
288"C and 538"C (70"F, 550"F and lOOO"F).

WING TEST BOX


8007 36~36
F/D
I, ~00

3 DO

$ (i 18

2.00

\
1.00
\
X
xx \
lo.-oo----,-5.oo_____z~o~-o--,7~\+~-~~--3o~o-o----~~.oo-
OOL-----5o_o_____
BucKuHCi STRESS " c;.: ?J. S )HJ/rr~m"

Figure 15.46 Israel Aircraft Industries buckling experiments on an integrally stiffened. integral fuel
tank-F/S plot for a local buckling test, with the specimen loaded only in the longitu-
dinal direction (from [15.64]): here F is the applied (reference) compressive stress and
D represents the bending strain e36-e38.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Static Nondestructive Methods 1293

WING TEST BOX


6007 Z3·25
Flo
04 s.u 23 SG :15

OJ

TRANSITION fR f1
MOOE Ill TO HODE I
~~' /-
A

02

Dk=-~k
Sli23

\ MOD£ I fXTR;pOlAI!ON
I

20.00
117
BUCkLING

Figure 15.47 Israel Aircraft Industries buckling experiments on an integrally stiffened, integral fuel
tank-F IS plot for a general instability test, under combined bending and pressure (from
[15.64]): here F is the applied (reference) compressive stress and D represents the bend-
ing strain £23-£25. Note also the change in the buckling mode

Each panel was fabricated from two sheets of Rene 41 alloy, seam-welded together to
form five fiat regions (double sheets) and four noncircular regions (flattened tubes formed
by the union of two circular arcs), as can be seen in Figure 15.48. The five tubular panels
were installed in the lower wing root zone of the HWTS (which was 7.9 m 2 in area), since
this was the most critically compression-loaded area, and were subjected to mechanical
loading, applied by 20 channels of closed-loop electrohydraulic equipment, combined with
simulated aerodynamic heating. The radiant heat flux for this heating was provided by a
system of water-cooled infrared quartz lamps.
The five panels were simultaneously and repeatedly tested for their bucking characteristics.
Nondestructive buckling tests were carried out under different combined loading conditions
and temperature environments, using the force/ stiffness technique for prediction of buckling
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

t =
42.900

0000~
- 2tls + I)
t = - - - = 0.037
I

Figure 15.48 NASA Dryden hypersonic wing panel tests-geometry of tubular panel tested, with
dimensions given in inches (from [ 15.66], courtesy of Dr. W.L. Ko, NASA Dryden Flight
Research Center, Edwards)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1294 Nondestructive Buckling Tests

loads. A typical F IS plot for tubular panel no. 3 under combined compression and shear
loading (Nj Nx = 1.22) at room temperature is shown in Figure 15.49. The loading was
stopped at 55 percent of the critical axial compression, and the intersection of the extrapo-
lated curve (obtained by a least squares fit) with the limit (or limiting) strain line yielded
the predicted buckling load F". The force at load point 703 (associated with a particular
jack) was arbitrarily chosen as the reference compression force F. The F IS plot, up to F =
0.6 F,, for the same panel no. 3 under combined loading at room temperature, but here with
lateral pressure and a higher shear to compression ratio, N,) N, = 3.21, is shown in Figure
15.50. One may note the different shapes of the F IS plots, which both yield satisfactory
predictions of the buckling load. In cases like Figure 15.50 it was found that often more
accurate buckling load predictions were obtained by curve-fitting only the higher load region
while ignoring the lower load one. In both Figures 15.49 and 15.50, and in the other cases
presented in [15.65], the FIS plot was used to predict local buckling failure.
Ko [15.66] applied Southwell plots to some of the tubular panel tests and compared the
buckling load predictions from the F IS plots with those from the Southwell plots. He found
that the F IS technique here gave more accurate predictions than the Southwell plots and that
in the case of pure compression the F IS plot yielded reasonably accurate buckling loads
even with the load cutoff point as low as F = 0.5 F, ,, whereas the Southwell method required
a much higher load cutoff point.
Another example of the application of F IS plots is the nondestructive tests carried out at
the 629th Research Institute of the Ministry of Aeronautical Industry, China in the mid-
eighties [15.67]. There the technique was applied to integrally stiffened panels and to a plate
with a hole and was found to be successful. The accuracy of the buckling load predictions
was checked by comparison with destructive tests (in which strain gages and the shadow
moire technique were employed) and found to be very satisfactory.
Before we close this chapter, we should point out that, in addition to the methods dis-
cussed, other techniques have been studied (see for example the 1989 review of Mujundar
and Suryanarayan [15.68]). The dynam[c vibration correlation techniques and the static

I I I I I ....
zoooo

l-
--
" ~
IBooo
j f
l
l
. .:..
iii
~
~
~
~
I • 5 ~
I ~
I
l t
I ~
JoOoo l v t
~ I t I I t ....
;<
~ eOClO - ~517'

OQOO
"- SGSI6
6<>'0

4000

3ooo +400o 5000


F(7o3). lb ~,... :3GI7'

Figure 15.49 NASA Dryden hypersonic wing panel tests-F IS plot for tubular panel no. 3 under
combined loading (compression and shear with N, IN, = 1.22) at room temperature and
p = 0 (from [15.66], courtesy of Dr. W.L. Ko, NASA Dryden Flight Research Center,
Edwards) --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1295

8000
-rssT .q. • 4 · G (t-J,.;./>-.~,.• 3.41) -~
'PAt-.IJ;O:L.. S
! " ' 7"""F. P - . 75 1"'51. l t
STIO!C.... Io-1 Gi ..... G;!Oi'S SIWIT I
l t

..• .a
CSG S l<i>
C"SG, 517
RSG 6.:Z.3
~~~ 63.3 l ~
~
j ~
5~ ~
5000
l t
~ l I t
£ l t
.1 3000 l t
I
j '-' ~
I
I
I
t t t l l Nx

1000
I ~517
Fc,::::~9~1 I
0$00

IOOO
\: SG:SIG

F(7o3), lb.

Figure 15.50 NASA Dryden hypersonic wing panel tests-F IS plot for tubular panel no. 3 under
combined loading (compression, shear and lateral pressure, with Nj N, = 3.21) at room
temperature (from [15.661. courtesy of Dr. W.L. Ko, NASA Dryden Flight Research
Center, Edwards)

F IS plots have, however, been applied most extensively in practice and show the greatest
promise for the future. Since VCT deals only with general instability, while F IS techniques
can also treat local buckling, combined application of both approaches may develop into the
practical nondestructive test methods of the future.

References
15.1 McGonnagle, W.J., Nondestructive Testing, 2d ed., Gordon & Breach Science Publishers, New
York, London, Paris, 1975.
15.2 McMaster, R.C., Mcintire, P., Bryant, L.E. and Mester, M.L., eds, Nondestructive Testing Hand-
book, 2d Ed., American Society for Nondestructive Testing, Columbus, Ohio, and American
Society for Metals, Metals Park, Ohio, 1982.
15.3 Bray, D.E. and Stanley, R.K., Nondestructive Evaluation, A Tool for Design, Manufacturing and
Service, McGraw-Hill, New York, 1989.
15.4 American Society for Testing Materials, 1990 Annual Book of ASTM Standards, vol. 3.03, Non-
destructive Testing, ASTM, Philadelphia, 1990.
15.5 Birchon, D., Non-Destructive Testing, Engineering Design Guides 09, published for the (U.K.)
Design Council by Oxford University Press, Oxford, 1975.
15.6 Sommerfeld, A, Eine einfache Vorrichtung zur Veranschaulichung des Knickungsvorganges, Zeit-
schrift des Verein deutscher lngenieure (ZVDJ), 1905, 1320-1323.
15.7 Foppl, L., Bestimrnung der Knicklast eines Stabes aus Schwingungsversuchen, in: Beitrage zur
Technischen Mechanik und Technischen Physik, August Foppl Festschrift, Julius Springer, Berlin,
1924, 82-88.
15.8 Hohenemser, K. and Prager, W., Dynamik der Stabwerke, Springer-Verlag, Berlin, 1933.
15.9 Stephans, B.C., Natural Vibration Frequencies of Structural Members as an Indication of End
Fixity and Magnitude of Stress, Journal of the Aeronautical Sciences, 4, (2), 1936, 54-60.
15.10 Massonnet, C., Les relations entre les modes normaux de vibration et la stabilite des systemes
elastiques, Bulletin des cours et des laboratoires d'essais des constructions du genie civil et
d'hydraulique fiuviale, vol. 1, nos. 1 and 2, Brussels, 1940.
15.11 Massonnet, C., Le Voilement des Plaques Planes Sollicitees dans leur Plan, Final Report of the
Third Congress of the International Association for Bridge and Structural Engineering, Liege,
Belgium, Sept. 1948, 291-300.
15.12 Lurie, H., Effective End Restraint of Columns by Frequency Measurements, Journal of the
Aeronautical Sciences, 18, 1951, 566-567.
15.13 Chu, T.H., Determination of Buckling Loads by Frequency Measurements, Thesis, California
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Institute of Technology, Pasadena, Calif, 1949.


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1296 Nondestructive Buckling Tests

15.14 Lurie, H., Lateral Vibrations as Related to Structural Stability, Journal of Applied Mechanics,
ASME, 19, June 1952, 195-204.
15.15 Davidson, T., and Meier, J.H., Impact on Prismatical Bars, Proceedings of the Society for Ex-
perimental Stress Analysis, 4, (1 ), 1946, 88-111, App. 3.
15.16 Johnson, E.E. and Goldhammer, B.F.. The Determination of the Critical Load of a Column or
Stiffened Panel in Compression by the Vibration Method, Proceedings of the Society for Exper-
imental Stress Analysis, 11, {1), 1953, 221-232.
15.17 Meier, J.H., Discussion of the paper entitled "The Determination of the Critical Load of a
Column or Stiffened Panel in Compression by the Vibration Method," in: Proceedings of the
Society for Experimental Stress Analysis, 11, 1953, 233-234.
15.18 Klein, B., Determinations of E1Iective End Fixity of Columns with Unequal Rotational End
Restraints by Means of Vibration Test Data, Journal of the Royal Aeronautical Society, 61, (554),
1957, 131-132.
15.19 Burgreen, D., End Fixity Effect on Vibration and Stability, Journal of the Engineering Mechanics
Division, ASCE, 86, 1960, 13-28.
15.20 Jacobson, M.J. and Wenner, M.L., Predicting Buckling Loads from Vibration Data, Experimental
Mechanics, 8, (10), Oct. 1968. 35N-38N.
15.21 Jubb, J.E.M., Phillips, I. G. and Becker, H., Interrelation of Structural Stability, Stiffness, Residual
Stress and Natural Frequency, Journal of Sound and Vibration, 39. (1), 1975, 121-134.
15.22 Abramovich, H., Gil, J., Grunwald, A. and Rosen, A., Vibrations and Buckling of Radially
Loaded Circular Plates, TAE Report 332, Dept. of Aeronautical Engineering, Technion-Israel
Institute of Technology, Haifa, Israel. July 1975.
15.23 Kielb, R.E. and Han, L.S., Vibrations and Buckling of Rectangular Plates under Inplane Hydro-
static Loading, Journal of Sound and Vibration, 70, (2), 1980, 543-555.
15.24 White, R.G. and Teh, C.E., Dynamic Behaviour of Isotropic Plates under Combined Acoustic
Excitation and Static Inp1ane Compression, Journal of Sound and Vibration, 75, (4), 1981, 527-
547.
15.25 Okubo, S. and Whittier, J.S., A Note on Buckling and Vibrations of an Externally Pressurized
Shallow Spherical Shell, Journal of Applied Mechanics, ASME, 34, 1967, 1032-1034.
15.26 Archer, R.R. and Famili, J., On the Vibrations and Stability of Finitely Deformed Shallow Spher-
ical Shells, Journal of Applied Mechanics, 32, (1), March 1965, 116-120.
15.27 Kalnins, A., Free Non-Symmetric Vibration of Shallow Spherical Shells, in: Proceedings of the
4th U.S. National Congress of Applied Mechanics, ASME, New York, 1962, 225-233.
15.28 Rosen, A. and Singer, J., Vibration of Axially Loaded Stiffened Cylindrical Shells with Elastic
Restraints, International Journal of Solids and Structures, 12, 1976, 577-588.
15.29 Rosen, A. and Singer, J., Vibrations and Buckling of Eccentrically Loaded Stiffened Cylindrical
Shells, Experimental Mechanics, 16, 1976, 88-94.
15.30 Singer, J., Vibration Correlation Techniques for Improved Buckling Prediction of Imperfect Stiff-
ened Shells, in: Buckling of Shells in Off-Shore Structures, I.E. Harding, P.J. Dowling and N.
Agelidis, eds., Granada, London, 1982 . 285-329.
15.31 Singer, J., Abramovich, H. and Yaffe, R., Application of the BOSOR 4 Program to Stitiened
Shells on Elastic Restraints and Eccentric Supports, TAE Report 342, Department of Aeronautical
Engineering, Technion-Israel Institute of Technology, Haifa, Israel, June 1979.
15.32 Rosen, A. and Singer, J., The Effect of Axisymmetric Initial Imperfections on the Vibrations of
Cylindrical Shells under Axial Compression, AIAA Journal, 12, (7), July 1974, 995-997.
15.33 Rosen, A. and Singer, J., Influence of Asymmetric Imperfections on the Vibrations of Axially
Compressed Cylindrical Shells, Proceedings, 18th Israel Annual Conference on Aviation and
Astronautics, Israel Journal of Technology, 14, 1976, 23-36.
15.34 Singer, J. and Prucz, J., Influence of Imperfections on the Vibrations of Stiffened Cylindrical
Shells, Journal of Sound and Vibration, 80, 1982, 117-143.
15.35 Lipovskii, D.E. and Tokarenko, V.M., Effect of Initial Imperfection on Free Oscillation Fre-
quencies of Cylindrical Shells. All-Union Conference on the Theory of Shells and Plates, 6th
Baku Azerbaidzhan SSR, Sept. 15-20, 1966, Transactions, Moscow Izdatel'stvo Nauka, 1966,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`

542-547 (in Russian).


15.36 Rosen, A. and Singer, J., Further Experimental Studies of Vibrations of Axially Loaded Stiffened
Cylindrical Shells, TAE Report 210, Department of Aeronautical Engineering, Technion-Israel
Institute of Technology, Haifa, Israel, 1975.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1297

15.37 Singer, J. and Segal, Y., Further Experimental Studies on Vibrations and Buckling of Eccentri-
cally Loaded Stiffened Cylindrical Shells, TAE Report 331, Department of Aeronautical Engi-
neering, Technion-Israel Institute of Technology, Haifa, Israel, 1978.
15.38 Abramovich, H., Singer, J. and Grunwald, A., Nondestructive Determination of Interaction
Curves for Buckling of Stiffened Shells, TAE Report 341, Dept. of Aeronautical Engineering,
Technion-Israel Institute of Technology, Haifa, Israel, Dec. 1981.
15.39 Weller, T., Singer, J. and Koss, D., Further Vibration Correlation Studies on Stiffened Shells with
Practical Boundary Conditions, TAE Report 499, Department of Aeronautical Engineering, Tech-
nion-Israel Institute of Technology, Haifa, Israel, April 1987.
15.40 Weller, T., Abramovich, H. and Singer, J., Correlation between Vibrations and Buckling of
Cylindrical Shells Stiffened by Spot-Welded and Riveted Stringers, TAE Report 537, Department
of Aeronautical Engineering, Technion-Israel Institute of Technology, Haifa, Israel, 1985.
15.41 Ross, C.T.F., Vibration and Instability of Ring-Reinforced Circular Cylindrical and Conical
Shells, Journal of Ship Research, 20, (1), March 1976, 22-31.
15.42 Segal, Y., Prediction of Buckling Load and Loading Conditions of Stiffened Shells from Vibration
Tests, M.Sc. thesis, Faculty of Aeronautical Engineering, Technion-Israel Institute of Technol-
ogy, Haifa, Israel, Feb. 1980 (in Hebrew).
15.43 Radhakrishnan, R., Prediction of Buckling Strength of Cylindrical Shells from Their Natural
Frequencies, Earthquake Engineering and Structural Dynamics, 2, (2), 1973, 107-115.
15.44 Segall, A. and Baruch, M., A Nondestructive Dynamic Method for the Determination of the
Critical Load of Elastic Columns, Experimental Mechanics, 20, (8), August 1980. 285-288.
15.45 Baruch, M., Integral Equations for Nondestructive Determination of Buckling Loads for Elastic
Plates and Bars, Proceedings of the 15th Israel Annual Conference on Aviation and Astronautics,
Israel Journal of Technology, 11, (1-2), March 1973, 1-8.
15.46 Baruch, M., Undestructive Determination of the Buckling Load of an Elastic Bar, AIAA Journal,
8, (12), Dec. 1970, 2274-2276.
15.47 Souza, M.A., Fok, W.C. and Walker, A.C., Review of Experimental Techniques for Thin-Walled
Structures Liable to Buckling, Part !-Neutral and Unstable Buckling, Experimental Techniques,
7, Sept. 1983, 21-25.
15.48 Souza, M.A., The Effects of Initial Imperfection and Changing Support Conditions on the Vi-
bration of Structural Elements Liable to Buckling, Thin-Walled Structures. 5, 1987, 411-423.
15.49 Plaut, R.H. and Virgin, L.N., Use of Frequency Data to Predict Buckling, ASCE Journal of
Engineering Mechanics, 116, (10), Oct. 1990, 2330-2334.
15.50 Souza, M.A. and Assaid, L.M.B., A New Technique for the Prediction of Buckling Loads from
Nondestructive Vibration Tests, Experimental Mechanics, 31, June 1991, 93-97.
15.51 Souza, M.A., Coupled Dynamic Instability of Thin-Walled Structural Systems, Thin-Walled
Structures, 20, 1994, 139-149.
15.52 Souza, M.A. and Walker, A.C., Vibration Characteristics of Buckled Cylinders, in: Brasil Off-
shore 81, International Symposium on Offshore Engineering, Sept. 1981, Pentech Press, London,
1982, 382-394.
15.53 Schneider, M.H .. Snell, R.F., Tracy, J.J. and Powers, D.R., Buckling and Vibration of Externally
Pressurized Conical Shells with Continuous and Discontinuous Rings, AIAA Journal, 29, (9),
Sept. 1991, 1515-1522.
15.54 Schneider, M.H. and Tracy, J.J., The Use of Modal Parameters in the Prediction of Buckling of
Externally Pressurized Shells: Part 1, Analytical Considerations, in: Proceedings of the 7th In-
ternational Modal Analysis Conference, Las Vegas, Jan. 1989, 1319-1328.
15.55 Powers, D.R. and Driskill, P.H., The Use of Modal Parameters in the Prediction of Buckling of
Externally Pressurized Shells: Part 2, Experimental Description, in: Proceedings of the 7th In-
ternational Modal Analysis Conference, Las Vegas, Jan. 1989, 1329-1335.
15.56 Ilanko, S. and Dickinson, S.M., The Vibration and Post-Buckling of Geometrically Imperfect,
Simply Supported, Rectangular Plates under Uni-Axial Loading, Part II: Experimental Investi-
gation, Journal of Sound and Vibration, 118, 1987, 337-351.
15.57 Ross, C.T.F. and Seers, A., Non-Linear Vibrations of Ring-Stiffened Conical Shells under Ex-
ternal Water Pressure, in: Developments in Computer Aided Design and Modelling for Structural
Engineering, B.H.V. Topping, ed., Civil-Comp Press, Edinburgh, 1995.
15.58 Schuette, E.H. and Roy, J.A., The Determination of Effective Column Length from Strain Mea-
surements, NACA Wartime Report, Advanced Restricted Report (ARR) L4F24, June 1944.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1298 Nondestructive Buckling Tests

15.59 Horton, W.H., Craig, J.I. and Struble. D.E., A Simple, Practical Method for the Experimental
Determination of the End Fixity of a Column, in: Proceedings of the 8th International Symposium
on Space Technology and Science, Tokyo, 1969, 269-280.
15.60 Baruch, M., Determination of the Stiffness of an Elastic Bar, AIAA Journal, 9, (8), August 1971,
1637-1639.
15.61 Becker, H., Nondestructive Testing for Structural Stability, Journal of Ship Research, 13, Dec.
1969, 272-275.
15.62 Some Investigations of the General Instability of Stiffened Metal Cylinders, III-Continua-
tion of Tests of Wire-Braced Specimens and Preliminary Tests of Sheet-Covered Specimens,
GALCIT. California Institute of Technology, Pasadena, Calif., NACA TN-907, 1943.
15.63 Finance, I.P.A., Methodes Experimentales Contribuant a Ia Prevision des Ruptures lors des Essais
de Resistance sur Structures Complexes, L'aeronautique et /'astronautique, (78), 1979, 67-74.
15.64 Weissberg, V. and Baruch, M., Nondestructive Buckling Test for an Integrally Stiffened Structure,
Journal of Aircraft, 18, (9), Sept. 1981, 780-785.
15.65 Ko, W.L., Shideler, J.L. and Fields, R.A., Buckling Behavior of Rene 41 Tubular Panels for a
Hypersonic Aircraft Wing, in: Collection of Papers, AIAA/ASMEIASCE/AHS 27th Structures,
Structural Dynamics and Materials Conference, San Antonio, Tex., May 1986, 517-540 (AIAA
86-0978).
15.66 Ko, W.L., Accuracies of Southwell and Force/Stiffness Methods in the Prediction of Buckling
Strength of Hypersonic Aircraft Wing Tubular Panels, NASA TM 88295, Nov. 1987.
15.67 Yu, W., Zhang, Y., Zhang, C. and Ding, H., Nondestructive Buckling Testing Technique of Panels,
in: Proceedings, International Conference on Experimental Mechanics, Beijing, Oct. 1985,661-
667.
15.68 Mujundar, P.M. and Suryanarayan, S., Nondestructive Techniques for Prediction of Buckling
Loads-A Review, Journal of the Aeronautical Society of India, 41, 1989, 205-223.
15.69 Weller, T. and Abramovich, H., Effect of Sequence of Loading on Interaction Curves for Buckling
of Stiffened Shells, TAE Report 536, Dept. of Aeronautical Engineering, Technion-Israel In-

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
stitute of Technology, Haifa, Israel, Sept. 1984.

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
16
Plastic Buckling
Experiments

16.1 Plastic Buckling Phenomena


16.1.1 Introduction
Inelastic behavior appears in the postbuckling of most practical thin-walled structural ele-
ments, as has been pointed out in the previous chapters. When the walls of the structure are
not very thin, fully plastic buckling may also occur, in which significant parts of the structure
yield before the onset of buckling. In the usual elasto-plastic buckling, plasticity appears just
before buckling, or immediately after it, and is often taken into account in analysis and
design by a plasticity factor (see for example [2.2], [2.3], [2.7], [2.10], [2.13], [6.3], [16.1]
and [ 16.2]) or by inclusion in the analysis of the plastic deformation that accompanies the
large deflections that precede failure (see for example [16.3], [16.4] and [16.5]).
Though the plastic buckling of columns has been studied for more than a century, the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

analysis of inelastic buckling behavior of other structural elements was initiated only in the
forties and fifties, after extensive testing of plates and some of shells at that time. Plastic
bifurcation and postbifurcation theory has been placed on a firm mathematical foundation in
the last decades (see for example [2.79]-[2.81], [16.4] and [16.6]); models that lead to a
better understanding of bifurcation and postbifurcation behavior have been developed (see
for example [16.4] and [ 16.7] and [ 16.8]); an asymptotic theory of initial postbuckling be-
havior in the plastic range has been developed (see for example [16.4] and [16.7]-[16.11]);
numerical solutions have been obtained for plastic postbuckling behavior (for example
[ 16.9]-[ 16.11 ]); and many additional analytical and numerical studies are in progress.
Unfmtunately, the enormous theoretical efforts in recent years have not been adequately
matched by extensive experimental work, and hence there is a need for further careful plastic
buckling and postbuckling tests. However, as has been pointed out in previous chapters, in
many cases the test setups and techniques employed for elastic buckling have also been used
for the elasto-plastic part of the tests. In this chapter the emphasis will therefore be on
experiments devoted primarily to plastic buckling.
The study of plastic buckling has been troubled by two widely discussed controversies,
or paradoxes. One, the column paradox, was settled in the late forties, whereas the other,
the flow theory versus deformation theory paradox, or plastic buckling paradox, is still dis-
puted today. Both controversies are briefly discussed below.

16.1.2 Inelastic Column Theory


As outlined in Chapter 2, Subsection 2.1.13, two models competed for representation of
plastic column behavior: the reduced modulus theory, which included strain reversal, and
Copyrightthe
Wiley tangent modulus theory, with no strain reversal.
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
Buckling
No reproduction Experiments:
or networking permittedExperimental Methods
in Buckling of Thin-Walled
without license from IHS Structures:
Not for Resale, Shells,
02/13/2019 Built-Up
01:32:58 MST Structures, Composites
and Additional Topics – Volume 2. J. Singer, J. Arbocz and T. Weller Copyright © 2002 John Wiley & Sons, Inc.
1300 Plastic Buckling Experiments

The history of plastic column theory may be concisely summarized, following Sewell
[2.81]:
In 1889 Engesser [16.12] first proposed the tangent modulus load to replace the Euler load in the
plastic stress range, but on the erroneous basis that it permitted adjacent equilibrium positions.
Considere [2.74] in 1891 first established the reduced (or double) modulus load as the lowest load
permitting deflection under no change of load, and consequently involved unloading over part of
the column (strain reversal). Engesser [2.75] acknowledged his error in 1895, and the reduced
modulus became firmly established, especially after von Karman's [4.41 use of it in 1910, until
1947. Then Shanley [16.13] and [2.11] pointed out that quasi-static deflection could indeed begin
at the tangent modulus value (explaining the significance of Engesser's 1889 tangent modulus load),
provided the load were simultaneously increased sufficiently to permit continued plastic loading
over the whole cross-section (no strain reversal). Von Karman agreed with this and clearly expressed
the idea of (inelastic) bifurcation [ 16.14].

Many other investigators participated in the inelastic column controversy, some also after
Shanley in 1947 explained the significance of column bifurcation at the tangent modulus
load (see for example [16.15]-[ 16.23]). Though the debate continued into the fifties and
beyond, it was von Karman's discussion of Shanley's 1947 paper [16.14] that actually settled
the controversy, by clarifying that the tangent modulus load and the reduced modulus load
are answers to two different equations. Or, in von Karman's formulation (italics added):
Both Engesser's and my own analyses of the problem were based on the assumption that the
equilibrium of the straight column becomes unstable when there are equilibrium positions infini-
tesimally near to the straight equilibrium position under the same axial load. The correct answer
to this question is given by replacing, in Euler's equation, Young's modulus by the so-called
reduced modulus. Mr. Shanley's analysis represents a generalization of the question. His procedure
can be formulated as follows: What is the smallest value of the axial load at which a bifurcation
of the equilibrium positions can occur, regardless of whether or not the transition to the bent
position requires an increase of the axial load') The answer to this question is that the first equi-
librium bifurcation from the straight equilibrium configuration occurs at a load given by the Euler
formula when the Young's modulus is replaced by the tangent modulus .
. . . Two aspects of the question are worthy of mention:
(a) My original analysis (leading to the reduced modulus load), and also Engesser's, is a gen-
eralization of the reasoning used in the theory of elastic buckling.

(b) Although the Euler formula with the tangent modulus does not, in general, give the maximum
load to which the column can be subjected without large deflection, it is conservative and therefore
advisable to use this formula for practical computation of column loads.

[It] is difficult to determine the actual peak of the axial load. It will certainly be between the
two values that correspond to the tangent and the reduced moduli. These two values can be re-
spectively designated as the lower and upper limits of the critical load.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Typical experimental results, obtained in the late thirties in tests of solid circular aluminum
alloy columns (see Figure 16.1, from [6.46] with test data from [16.24]), or in tests of I-
section aluminum alloy columns (see Figure 16.2, from [16.15] with test data from [6.11]),
were in satisfactory agreement with the tangent modulus theory. For the higher buckling
stresses some of the test data were slightly above the tangent modulus curve, but still well
below the reduced modulus curve, reconfirming the last two sentences quoted above from
von Karman's short discussion of Shanley's paper.
The lively column controversy not only resulted in many theoretical studies (like [ 16.15]-
[16.22]) but also motivated in the fifties very careful experiments, as well as related analytical
studies, of the actual behavior of columns in a testing machine, primarily by Hoff at the
Polytechnic Institute of Brooklyn (see [16.22] and [16.23]). It may also be pointed out that
the tangent modulus approach was generalized by Hill [2.80] for bifurcation analysis of
three-dimensional elastoplastic bodies.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Phenomena 1301

60

\ Alv~invm Afloy ,7Lr


'\
\ I
\£educed, odvlu; \
\
so
~ \
Euler I .o!;~if'i~a<t/:/d
~~--
b..,---
\~
I l I
40
Tar. qent Aodu/t.),f
~
I
II
30 ~' I

20
~I Ii
~I I

10
I
I
I I ~ ~I
I

0 /0 /0 30 -10 so 60 -,o &0 90


i
100

J/enderne" Ratio ~

Figure 16.1 Experimental results obtained in ALCOA tests on solid circular 2017-T aluminum alloy
columns in the late thirties show satisfactory agreement with the tangent modulus theory
for short columns (from [6.46] with test data from [16.24]). For the shortest columns some
of the test data are slightly above the tangent modulus curve, but still well below the
reduced modulus one

16.1.3 The Flow Theory versus Deformation Theory Paradox- The


Cruciform Column
In the late forties and early fifties it was noted that bifurcation loads calculated using the
simplest smooth yield surface incremental, or flow theory (J 2 flow theory) consistently over-
estimated the experimental buckling loads of plates and shells. On the other hand, calcula-
tions based on the corresponding deformation theory, which is actually not a physically
acceptable plasticity theory, gave much better agreement with tests results. This is the so-
called "plastic buckling paradox," which has been comprehensively discussed by Hutchinson
[16.4], as well as by Tvergaard and Needleman [16.6]. Part of the presentation here follows
the discussions in these papers.
The cruciform column, which has played a central role in the controversy, is probably the
best example demonstrating the discrepancy between the predictions of flow theory and

50,000 . - - - . - - - . . . - - - - - - - - - - - - - - - - , . - - - - ,

1-- ' v; --- -1


--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

~l.
r11
z
:>
:0

8 1qooo r----+----+ [liLER'S LOI<G COLUMN rORMULA---...__

20 40 60 80 100 120 140 160


EFH:CliVE SLENDERNCSS RATIO. lclp

Figure 16.2 Experimental results obtained in ALCOA tests on I-section 2024-T aluminum alloy col-
umns in the late thirties show satisfactory agreement with tangent modulus theory for
shot1 columns (from [16.15] with test data from [1.29])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1302 Plastic Buckling Experiments

dcfonnation theory. The buckling of a crucifonn column under axial compre~;ion (see Fig·
ure~ 16.3 and 16.4) was originally studied hy StO\\CII at NACA Langley in the late fonies
(6.25(- (6.27]. If the column is not too long. its bud.ling mode is pure torsion about it> axi\.
Only the effective shear modulus enters into the fom1ula for the bifurcation stress.
ln the clastic range the cri tical compressive stress (at bifurcation) is
u, = G(ti!J)l ( 16. 1)
where G is the clastic shear modulus, 1 is the thickness, and b is the width of the flange
plates (sec Figure 16.4). In the plastic range (Eq. ( 16.1 )) still holds accordi ng to J2 flow
theory. This is so because th is plasticity theory has a smooth yield surface and therefore the
increments in the relevant components of shear stress and strain following uniaxit~ l com-
pression arc related by the clastic shear modulus.
On the other hand, for any deformation theory. for an initially isotropic material. the
relevant instantaneous shear modulus G (often denoted G,). following uniaxial cornprcs>ion.
that replaces the elastic shear modulus in Eq. ( I6. I) is given by
- G
( 16.2)
G =I + 3G((I/£,) (I/£))
where E. ull: is the current secant modulu s. The defonnation theory (and consequently
ah.o sli p theory) therefore predicts a critical ;.trc>s
u, = ccttbl' ( 16.3)
Thus, the ratio of the deformation theory pred iction to that of the si mple flow theory is
(GIG). For a Poisson's ratio,, = 0.5 , G "' (/;' /2( 1 + ,,JI = {£/3), and using Eq. ( 16.2) this

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
ratio (G/G) becomes(£,/£).
In Figure 16.4 experimental results. obtained from NACA Langley tests on 2024-T4 al·
uminium alloy crucifonn columns and taken from Figure 5 of [2.13] (some of these points
appeared also in Figure I of (16.26]). are plotted in the fom1 of buckling stress u , nonnalitcd
by G(llb)' versus u ,. The discrepancy between the two theories for the cruciform column
i~ indeed very pronounced. The dcfonnation thcol) predictions. (b) in Figure 16.4. arc in
good agreement with the experimental re,ults, wherca~ the predictions of J, flow theory. (a)
in the figure. widely diverge from them.

Fi~urc 16.3 ' lb rsiunat buckling of a 2024·T4 aluminum alloy cruciform column tested at NACA L!lng·
Icy in the late forties (from [ 16.27 J)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Phenomena 1303

"c
G{t/b)2.6

%·~~~10~~20~~3~0_L_4~0~~50~~60
"c I k" I

Figure 16.4 Theoretical and experimental results for the plastic buckling of a cruciform column (from
[16.4]). The test data are from Stowell's NACA tests on 2024-T4 aluminum alloy cruci-
form sections [16.27], as reported in [2.13]. Curve (a) is the prediction of J, flow theory
with a smooth yield surface, while curve (b) is the prediction of any deformation theory
with v = (l /2), showing satisfactory agreement with test data

The attempts to resolve this discrepancy preceded in two directions: one is the unavoidable
impeifections approach, and the other is the corner theories approach.
The first direction was assertively presented by Onat and Drucker [16.28], who carried
out a detailed, yet approximate, calculation of the maximum support load based on J2 flow
theory and found that extremely small initial imperfections could reduce the maximum load
to the level predicted by deformation theory and observed experimentally. However, the
imperfections required to bring about this reduction were so small that one should expect

10-·-~-~--

6
G(tf
0.8

0.6

_ -------~ ''Blunt Corner ..


~ "Sharp Corner"
0.4 Deformation Theory

0.2
x Max1mum Load

0.1 0.2 03 0.4 0.5


Buckl1ng Mode Amplitude

Figure 16.5 Influence of a small imperfection on the buckling of a cruciform column for various
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

constitutive models. including two comer theories (from [16.6))


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1304 Plastic Buckling Experiments

no significant scatter in the buckling loads, whereas the experimental results in Figure 16.4
show appreciable scatter. A later study by Hutchinson and Budiansky [16.29] confirmed the
Onat-Drucker conclusion of the feasibility of a very large influence of exceedingly small
imperfections, provided the strain-hardening was sufficiently low. But for materials with
relatively high strain-hardening the required imperfection levels were not small anymore.
The other direction, the corner theories approach, which was initiated by Batdorf [16.30],
searched for a justification of the deformation theory predictions (and experimental results)
in terms of a sophisticated flow theory that permitted the development of a vertex on the
yield surface. Though the occurrence of vertices on yield surfaces is supported by models
of the elastic-plastic behavior of polycrystalline metals based on a single crystal slip (see
for example [16.31] or [16.32]), the experimental evidence is contradictory.
Incremental theories, whose bifurcation predictions for nearly proportional loading paths
are similar to those given by deformation theory, have been generated, based on a system
of linear loading functions. In the late seventies a phenomenological corner theory, called J2
corner theory, was developed by Christoffersen and Hutchinson [16.33]. In this theory (as
summarized in [16.6]) the instantaneous moduli for nearly proportional loading are chosen
equal to the J 2 deformation theory moduli, and for increasing deviation from proportional
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

loading the moduli stiffen monotonically until they coincide with the elastic moduli for stress
increments directed along or within the corner of the yield surface. The aim of this corner
theory was to provide a phenomenological theory embodying features of physical theories,
while being tractable enough for analyses of postbifurcation and imperfection sensitivity.
For example, this J2 corner theory was employed by Needleman and Tvergaard in their
study of the imperfection sensitivity of the cruciform column [16.34]. Their results, for a
case when the ratio of the corner theory bifurcation stress to the smooth yield surface flow
theory bifurcation stress of Eq. (16.1) is 112, are shown in Figure 16.5. Each case in the
figure has an identical uniaxial stress-strain curve, as well as an identical initial imperfection
amplitude. For the sharp corner in Figure 16.5, the maximum load is only slightly higher
than that given by deformation theory, while for the blunter corner the maximum load is
about halfway between the values of deformation theory and those of flow theory. Also, for
all the other cases considered in [16.34] the corner theory predictions fall between the 12
deformation theory and J 2 flow theory predictions.
The corner theory suffers from a severe drawback, namely that although one can fit the
appropriate corner to observed results and thus obtain good agreement, one cannot estimate
the appropriate corner a priori. Hence, corner theory can assist in the explanation of exper-
imental results, but it cannot as yet serve as a design tool. This is also pointed out later in
the discussion of plastic buckling experiments on cylindrical shells subjected to combined
loading in Subsection 16.2.4.
Here one should perhaps reiterate Hutchinson's authoritative statements in favor of the
"less respectable" deformation theory (see [16.4]) that "this example (the cruciform column)
does lend further credibility to the usc of bifurcation load predictions of deformation theory
for engineering purposes"; and that "in the meantime there seems to be little doubt that for
engineering purposes bifurcation predictions based on deformation theory should be favored
over those based on incremental theories with smooth yield surfaces." But he added a warn-
ing that "where deformation histories do depart from total loading, as may be the case in
the post-bifurcation regime, for example, deformation theory predictions must obviously be
regarded with suspicion."
Studies by Bodner and Rubin [16.35] and [16.36] attack the plastic buckling paradox by
a "proper treatment of realistic material behavior upon an abrupt change of straining direc-
tion," and "modification of the plastic flow law" with the aid of a generalization of the
Bodner-Partom unified elastic-viscoplastic theory. Though in their detailed discussion of the
Stowell cruciform columns (see [16.36]), their theoretical predictions compare reasonably
well with the experimental values, and their analysis significantly contributes to the allevi-
ation of the plastic buckling paradox, one cannot yet consider the paradox as settled. Hence,
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1305

for the time being, Hutchinson's recommendation in favor of the deformation theory bifur-
cation predictions still stands.
Further results from plastic buckling experiments on plates and shells, discussed in Sub-
sections 16.2.2, 16.2.4, 16.2.6 and 16.3.3, confirm that the plastic buckling paradox is still
not resolved.
Before we leave this section, we should note that the discrepancy between the predictions
of flow theory and deformation theory has been clearly demonstrated for plastic buckling of
many structural elements, such as clamped circular plate under radial compression, or a thin
spherical shell under uniform external pressure in [16.4] (though in both these cases without
reference to experimental results), or thin cylindrical shells under axial compression in
[16.37] (see Figure 16.6). The discrepancy is, however, usually not as dramatic as for the
cruciform column.
But one example, clamped annular plates in pure shear, studied recently by Ore and
Durban, exhibits an even greater discrepancy between the predictions of J2 flow theory and
deformation theory than that of the infamous cruciform column (see [16.39]). As can be
seen in Figure 16.7, the predictions of deformation theory agree well with experimental
results obtained by Bauer (the details of the tests are discussed in Subsection 16.2.2 and in
[16.40]-[16.42]), whereas those of the J2 flow theory are considerably above the measured
values. The clamped annular plates in pure shear therefore represent another strong confir-
mation of the plastic buckling paradox. It and the other examples provide additional support
for the recommended use of deformation theory in plastic buckling.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
16.2 Plastic Buckling Experiments
16.2.1 Plastic Buckling of Columns
Plastic buckling experiments on columns were discussed in Chapter 4, where von Karman's
"classical" 1910 experiments were presented, and in Chapter 6, where experimental column
curves, test fixtures and test procedures (which include the plastic range) were considered
in detail. Hence here only a few additional short column tests that demonstrate instrumen-
tation and techniques of particular interest will be examined.

800001--r---
a,;,[psi]
THEORETICAL
Af>t:.At:.. EXPERIMENTAL
70000 ( R•L 16.38)

60000

50000

••
40000

30000
..
0 20 40 60 80 100 120
Rlh

Figure 16.6 Plastic buckling of axially compressed 2024-T4 aluminum alloy, simply supported, circular
cylindrical shells-comparison of ern predicted by 12 flow and deformation theories with
the experimental results of Batterman, as a function of the radius to thickness ratio R I h
(from [16.37] with data from [16.38])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1306 Plastic Buckling Experiments

18
Mer. th
-M--16
cr. exp
i:J. f" "OW THEORY
14 0 D~FORMATIUN I rl::Of-1Y

12

10

"' "' "'


1 ----~--:~"§-------'"--~--8--s-
1

20 40 60 80 100 120 140 160


b-a
-~-

Figure 16.7 Elastoplastic buckling of annular plates in pure shear-the ratio of the theoretical critical
torsion moment M"'"" for 12 flow and deformation theories, to the experimental critical
torsion moment M"""' measured by Bauer [16.42] on three groups of annular plates made
of AI 98.7 W aluminum, 1403 >.tee! and CuZn 30 brass (from [16.39]). Note that whereas
the deformation theory predictions agree well with the experimental values, the flow theory
ones are considerably above them

Some of the experimental studies of plastic buckling of aluminum columns sponsored by


NACA in the thirties and forties are of interest even today. For instance, the tests on extruded
aluminum alloy H-sections at the U.S. National Bureau of Standards and the Aluminum
Company of America in the late thirties [ 1.29], mentioned in Chapters l and 6, excelled in
their careful execution and clear specification and actual test conditions of round ends (pinned
supports) and fiat ends (clamped ends). Employing the effective lengths (L0 = L for round
ends and L0 = 0.5L for fiat ends) meant that the results could be conveniently plotted in the
same figure, facilitating direct comparison of tests with the two types of end conditions.
Most of the specimens in these tests were not very long and failed by plastic buckling, as
is evident in Figure 16.8.
Another series of NACA-sponsored tests on thick tubes elastically restrained against ro-
tation at the ends, carried out at the Bureau of Standards in the same period [16.43], rep-
resents another example worth noting. The series included 200 specimens, with (Rit) = 6-39
and slenderness ratios 16-150, the majority being in the inelastic range and made of chro-
mium-molybdenum steel, heat-treated chromium-molybdenum steel, stainless steel and du-
ralumin. The experimental program was aimed at providing data for a design method for
compression members elastically restrained at the ends.
Special fixtures were developed for procuring predetermined elastic restraints at the ends.
The lower fixture is shown in Figure 16.9. It consisted essentially of a carrier with a knife
edge, which bore on a seat on a stationary support that was clamped to the weighing table
(lower platen) of the testing machine. The end of the specimen was held in position on the
carrier and could be moved horizontally under low loads in a direction perpendicular to the
knife edge. Rotation of the carrier about the knife edge was restrained by the two helical
springs shown in the figure. The degree of restraint could be varied by changing the active
length of the springs. The rotation of the carrier about the knife edge could be measured by
a dial gage (not shown in the figure). Wing nuts on the ends of the rods, extending through
the springs made it possible to compress the springs, so that they would remain in com-
pression during rotation of the carrier (since the springs could not be used in tension). The
upper end fixture was essentially the same as the lower one.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1307

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
(a) (b)

Figure 16.8 U.S. National Bureau of Standard' and ALCOA tc't' on exlrudcd aluminum-alloy H-
SCCtJon columns in the late lhinie~ 'pecimeos or one or lhe two cro;;-scctioos aflcr
te,ting (from [ 1.29)): (a) tested with round ends. (b) tc\tcd with nat ends

Before a series of columns with elasti cally restrai ned ends could be tested. the desired
degree of rotational restraint had to be provided by adj usting the active length of the springs.
namely by compressing them with the wing nuts. The restraint provided was accurately
determined by hanging a series of known weights on one of the hangers at the ends of the
carrier (see Figure 16.9). This caused the carrier to rotate about the knife edge. and by
correspond ing rotation readings of the dial gage (not shown in the figure) being noted. the

Figure 16.9 U.S. National Bureau of Standards test< nn thick tubes claqically restrained against rotation
at the ends in the late thirties--{!ingram ~howing the lower end fixture providing elastic
rc<troint (from )16.4 3))
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1308 Plastic Buckling Experiments

moment per unit angular rotation could be accurately determined. One had to remember to
center the column under load with springs out of action, and only then to bring the springs
into action. The fixture appears to be a simple, yet accurate means to apply a desired pre-
determined rotational end restraint to columns.
Two decades later the plastic buckling of aluminum columns became a major concern for
civil engineers, which led to further studies of the phenomenon, some of them experimental.
One of these, carried out by Chilver and Brivtec in the Engineering Laboratory of Cambridge
University [16.44], focused on buckling and postbuckling behavior in the transition from
elastic to fully plastic conditions and pointed out that the buckling mechanism changed
during this transition. The study was accompanied by very careful tests on ball-ended alu-
minum alloy columns of !-section, with slenderness ratios (L/ p) = 20-65. At (L/ p) = 60
the initial buckling of the column was elastic. As the slenderness ratio was reduced, plasticity
effects became important, and in the range of (LIp) ~ 40 buckling was observed to be
dynamic in nature (unstable postbuckling). As the slenderness ratio was still further reduced,
buckling became wholly plastic, but its behavior reverted to static (stable postbuckling).
The columns were tested in compression with specially designed adjustable ball-end fit-
tings (Figure 16.10). Each fitting consisted of a hardened steel ball (a), which was centrally
seated in an aluminun alloy block (b). On the upper surface of this block were two ground
steel plates (c) and (d), plate (d) being attached to the block (b) plate (c) being held in a
rectangular frame (e) that could be adjusted to any lateral position by means of the differ-
ential screw device (f). This device consisted of a screw having a right-hand thread with a
pitch of 30 threads per in. working in conjunction with a left-hand thread of 31 threads per
in. One complete turn of the screw gave a displacement of about 0.001 in. of the frame (e)
along the axis of the screw. Hence, if the end of a column was held in frame (e), the relative
positions of the centroidal axis of the column and the centerline of the steel ball could be
accurately adjusted (in the direction of the axis of the screw, which lay in the plane of lateral
buckling of the column). During the adjustment plate (c), which carried the column, slid
over plate (d).

e b c d

Figure 16.10 Cambridge University tests on ball-ended aluminum alloy columns-adjustable end fit-
ting (from [16.44]): (a) hardened steel ball, (b) aluminum-alloy block, (c) and (d) ground
steel plates, the latter being attached to block (b) while the former is held in a rectangular
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
frame (e), (f) differential screw device, (g) set screws
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1309

For testing, the !-section column was fitted into frames (e) of the top and bottom fittings
and held tightly in a central position by screws (g). With both end fittings clamped to the
column and with the hardened steel balls in their seating, the specimen was transferred to a
compression testing machine and lined up as centrally as possible. The steel balls were
positioned on hardened-steel pads attached to the testing machine. By repeated loading within
the elastic range and adjustment of the screw device (f) while the column was under a small
compressive load of 0.5 tons, the initial eccentricity of loading could be minimized. These
adjustable end fittings were essentially a development of the much earlier ( 191 0) von Karman
column end fixtures discussed in Chapter 4 (see Figure 4.2).
Most of the later experiments on plastic buckling of columns have dealt with built-up
columns, primarily those with cross-sections such as Z-, H- or channel sections or square
tubes, which are essentially an assembly of plates. When the columns are very thin-walled,
their local buckling is elastic, crippling, which is discussed in Section 6.2 of Chapter 6.
When the walls are not so thin, the local buckling of the plate elements is plastic and is
considered in the next subsection.

16.2.2 Plastic Buckling of Plates


While the analysis of plastic buckling of columns dates back to the end of the 19th century
and the early decades of the 20th, viable theoretical solutions for plastic buckling of plates
commenced only in the late thirties and forties. Hence, when Gerard presented his empirical
secant modulus method for inelastic buckling of plates in 1946 [16.45], he stated that "an
applicable theoretical solution for buckling of thin plates above the proportional limit has
not been found, although attempted solutions have been published." At the time, Gerard was
probably unaware of Bijlaard's earlier work in Delft, The Netherlands (see for example
[16.46]-[16.49]), which, though first published in the late thirties, became known in the
United States only in the late forties (see for example [16.52]).
In developing his secant modulus method [ 16.45] Gerard started by defining the secant
modulus
(16.4)
as the function that relates stress and strain directly in the yield region. Obviously in the
elastic range E, = E. Then he made the assumption that the critical stress and critical strain
are implicitly related by the stress-strain curve of the material. With this assumption, the
plate buckling equation in the inelastic range (above the proportional limit) could be written
as
(16.5)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Here t is the plate thickness, b its width and


K = k[7T 2 /12(1 - v 2 )], (16.6)
where k is the usual elastic plate buckling coefficient, which is solely a function of the
geometry of the plate for a given material.
In terms of critical stress Eq. (16.5) becomes
(16.7)
or if one defines here the plasticity coefficient
7J = (EJE) (16.8)
Eq. (16.7) can be written as
(aj7J) = KE(tlbf (16.9)
Gerard carried out a test program on Z- and channel section aluminum alloy columns to
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1310 Plastic Buckling Experiments

validate the use of the secant modulus and to substantiate the assumption that the critical
stress and critical strain are implicitly related by the stress-strain curve of the material.
As indicated in Section 8.2 of Chapter 8, simply supported edge conditions in buckling
tests of uniaxially compressed rectangular plates were difficult to realize (and hence moti-
vated many sophisticated test arrangements, discussed in Subsections 8.2.2-8.2.5). This was
true also when the critical stress occurred in the plastic range. Gerard pointed out that the
use of Z- or channel sections obviated this difficulty along the unloaded edges (whose bound-
ary conditions are predominant in the determination of the buckling coefficient) since the
geometry of the cross-section provided well-defined support for the fiat plate elements, com-
prising the cross section. Obviously, this fact extends to other similar cross-sections, like H-
sections and rectangular-tube sections. The use of such sections therefore offered a conven-
ient experimental method for duplicating boundary conditions specified by theory.
The tests included two sets of specimens: one set formed from 2024-0 and one from
Alclad 7075-0. Both sets were subsequently heat-treated and aged to the T temper. The web
had the same width in all the cross-sections, while the flanges varied. All specimens were
designed to have the flanges buckle first. Two strain gages were cemented back to back at
the center of the leg with the larger flange (taking into account the slight variations in flange
dimensions resulting from machining tolerances) on each specimen. The specimens were
only 203 mm (8 in.) long to prevent premature column failure, and all were tested flat ended.
During a typical test the stress (the applied load divided by the cross-sectional area) was
plotted against the difference in measured strain (which is a function of the bending stress
in the flange when buckling initiates), the buckling stress being taken as the point at which
the bending strain became large with small increase in axial load. The critical average strain
was determined in a similar manner.
The test values of critical stress and critical strain for the Z- and channel specimens are
plotted in Figure 16.11. The compression stress-strain curves of the two aluminum alloys,
2024- T and Ale lad 7075-T, are also included in the figure with scatter bands in the plastic
range. Within experimental accuracy, Figure 16.11 confirms the basic assumption of the
secant modulus method that the critical stress and critical strain are implicitly related by the
stress-strain curve. For a given critical stress, most of the measured critical strain values

60
;i
0 I

50

~
;:o
""
o;:,
I
Ico I
TEST DATA
20
- STRESS-STRAIN
CJRVE
I
10 -

: I ,
___ [~j__j_~_:
0 .004 .006 .012 .016

E:cr

Figure 16.11 Gerard's tests for validation of 1~he secant modulus method-agreement between critical
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
stress-critical strain test data and stress-strain curve of the material (from [ 16.45])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1311

were somewhat less than those predicted from the stress-strain curve. This difference was
attributed to experimental difficulties in plastic strain measurements. Since the cement that
bonded the strain gages had a tendency to creep at large strains, the critical strains determined
by the strain gages were effectively smaller than their true values, which caused the apparent
slip to the left of the test data.
In Figure 16.12 the predictions of the secant modulus method are compared with the test
data for 2024-T aluminum alloy. The prediction chart, of ucr as a function of KE(tlb) 2 , was
obtained from Eq. (16.9) by multiplying the abscissa of the relevant stress-strain curve of
Figure 6.11 by the modulus of elasticity and then using the stress-strain curve to get the
corresponding critical stress. For example, the point 0.008 on the abscissa of Figure 6.11
became 85.6 kips/in. 2 , and the corresponding ordinate ucr = 49.2 kips/in. 2 was plotted at
that abscissa of 85.6 kips/in. 2 • In this manner the curve (marked Eq. 16.9) in Figure 16.12
that represents the secant modulus method was obtained. The agreement with test data for
all the 2024-T specimens was excellent. The agreement for Alclad 7075-T (shown in Figure
9b of [16.45]) was not as good, since the cladding introduced an indeterminacy that was
difficult to account for.
Though performed over half a century ago, Gerard's 1946 tests have been discussed here
in detail since they exhibit his practical way of thinking and represent simple yet careful
experimentation.

16.2.3 NACA Langley Tests in the Forties


During World War II extensive experimental studies on column and plate compressive
strength of aircraft structural materials were carried out at NACA Langley (see Figure 16.13)
and appeared as Wartime Reports (see for example [8.91], [16.51]-[16.54]). These tests
focused on local buckling and were actually plastic buckling tests on the plate elements of
the columns. The results of these studies, and of additional later ones as well as the secant
modulus method, were embodied in plate compressive curves, or plate buckling curves like
Figure 16.14, upon which the design procedures of the fifties and sixties were based (see
for example [16.56]).
All these tests were made in hydraulic testing machines, accurate within 0.75 percent. The
specimens were basically H-section extrusions of different web widths, in some of which
the flanges were removed to make Z- or channel sections, as desired. The flange widths of


~ 3of + .
12024·~
0 '-

J 20 I
I e CHANNEL
-THEORY

101--- -- f

0
0 40
·-'---:':---'---::-:--J 120

Figure 16.12 Gerard's tests for validation of the secant modulus method-agreement between test data
and secant modulus prediction for 2024-T aluminum alloy (from [16.45])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1312 Plastic Buckling Experiments

the sections were varied by milling off parts of the Hanges. As detailed in [6.18], '' the lengths
of the columns were chosen in such a way that the test results were essentially independent
of the length, that is, sufficiently long so that the increase in strength due to very short
lengths was avoided and at the same time not so long as to incur !he possibility of an Euler
of column failure. The length was also selccced so as 10 obtain a convenient chree-half-wave-
buckling pallern which gave a pronounced cross-sectional disconion al !he mid- length po-
si tion " (see Figure 16. 13 and [16.57]).
The columns were cested wich their machined flat ends bearing directly agai nst the heads
of the testing machine. The experimental critical local inscabi lity stress was taken as the
stress a1 the point near the lop of the fence of the stress-distortion curve at which a marked
increase in the distortion first occurred wich small increase in stress, the "top-of-che-knee
melhod" (see Figure 8.7 and Subseccion 8.3.1 in Chapler 8, or [2.15]). The cross-sectional
distortion was measured by various methods, and in che later tests it was recorded autograph-
ically together with !he load.
The idea of obtai ning wel l-defined unloaded boundaries of unidirectionally compressed
plates by using relatively short H-, Z- or channel scccion columns. employed by the NACA
Langley investigators and Gerard, was utilized in a different manner by Pride and Heimerl
at Langley, who used seamless square lubes (see [ 16.55]). During the buckling defonnalion
the adjacent faces of lhese tubes were oul of phase wich each ocher, leaving the corners
perpendicular and providing the longitudinal edges of !he four plate elements wi lh the un-
reslrained rocation required for si mple supporcs.
The specimens were cui from five differenl cross-sectional sizes of drawn square tube of
20 I4-T6 alumi num alloy. all having the same nominal thickness. Thus. plate elements of
(blr) = 20.8-42.5 and aspecl ratio (length lo width ratio) (Lib) = 4-6 were tested. Actually,
the program also included a few shorter tubes, down 10 (Lib) = I , butlhe minimum aspect
ratio to ensure test conditions of long simply supported places (avoiding the increase in
scrcnglh due 10 short plate lcnglh) was found co be (Lib) 2: 4.
lltc lalcral clisplaccmenls of the plate elements were measured by means of surface bearing
bars, held in direct contact with che plate sur face by flexible cancilevers (see Figure 16.15,
showing a typical square tube under lese). The critical compressive stress was again deler-
mined by the top-of-lhe-knee method, from autograph ic load displacemenl curves. obtained
from two adjacenl plate elemenls in each test specimen.

Figure 16.13 NACA Langley place compressive strength tests in the forties -local instabilily for an
H-section under test (from ( 16.52])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1313

These NACA tests by Heimerl and others, and in particular the later ones like [ 16.55],
also served to evaluate the plastic buckling theories for plates that were developed in late
forties [16.26], [16.27], [16.58] and [16.59], as well as those of Bijlaard and Gerard [16.50]
and [ 16.45]. Figure 16.16 presents a comparison of the experimental critical compressive
stresses for square tubes [16.55] with the predictions of the theories available at the time.
The test results were in excellent agreement with Stowell's unified theory for the plastic
buckling of columns and plates [16.26]. They also showed good agreement with Bijlaard's
[16.50] and Ilyushin's [16.58] theories, all three being deformation-type theories. The ex-
perimental results agreed also quite well with Gerard's secant modulus method [16.45]. But
there was significant disagreement with the predictions of Handelman and Prager [16.59],
which were based on a flow-type theory of plasticity-again a typical manifestation of the
plastic buckling paradox.
Stowell also verified his unified plastic buckling theory with experiments on cruciform
columns (see Figure 16.3 and [16.27], which were mentioned in Subsection 16.1.3 and
became the crucial example for the plastic buckling paradox. The cruciform column repre-
sents four identical hinged flanges that, if equally loaded, will twist at the same time without
restraint to each other. It therefore simulates ideal hinged flanges, provided the column is
short enough to trigger torsional buckling instead of the Euler type. The buckling and max-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

imum load were measured for each specimen, as well as its mid-span rotation. The experi-
mental points followed the relevant 2024-T4 aluminum alloy stress-strain curve, and there-
fore the appropriate reduced modulus was concluded to be the secant modulus. The
agreement between Stowell's theory and the experimental points for 2024-T4 cruciform
columns was found to be good (see for example Figure 7 of [16.27]), and was shown to
apply equally well to H-sections of other aluminum alloys.

16.2.4 More Recent Plastic Buckling Tests on Plates


The use of square tubes to represent long plates with well-defined unloaded edges in buckling
experiments is a method of simulation that has been considered a very useful and convenient
test method for many years. For example, Jubb et a!. employed welded mild-steel box col-
umns in their Cranfield vibration correlation tests in 1975 (see Subsection 15.2.3 of Chapter
15 and [15.21]), noting again that the plate elements of the column behaved as if they were
simply supported along their unloaded edges.

I
H-
-I
I-
Cc~, ks~o~ -+ J - ! ~ ' - _,
~-~~--- T~~
20:- - !0 f-c -I ~ --
J
j·-t; Ci ___'__ ---; -,.~,. , , ,
L__-~---"·---•--- "'\"'""'~"'"'""'"'
0 20 40 60 50

Figure 16.14 NACA Langley plate compressive strength tests in the forties-typical plate buckling
curve for extruded 7075-T aluminum alloy, obtained from tests of H-, Z- and channel
section columns (from [16.52]). Here h,, t, are the web width and thickness and k, is
the plate buckling coefficient corresponding to the web of the particular cross-section
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1314 Plastic Buckling Experiments

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Figure 16.15 NACA Langley plate comprc~<ivc ~trength tests in the la1c fonics-IC>I setup for square
tube; representing four stmpl) ;upponcd long plate,. 'howing the <urface beanng bars
employed for measurement of lateral d"placemenl> (from [16.55))

More recentl y, in 1993. Inoue at the University of Tsukuba. Japan. used welded steel box-
section subcolumns to simulate simply ~upported plates in hi~ plastic buckling experimenL~
[16.60). The mild-steel specimens were fonned by panial penetration welds at the comers.
The ends of the stub columns were mil led plane and perpendicular to the longi tudinal axis
and then annealed. Thus, relatively thick recwngular plate clements of (bit) = 10-30. and
aspect ratio (L/ b) = 1.5 and 3.0 were created. Thc'e were tel-led in compression. The ~imple
test ~etup employed is shown in Figure 16.17. The load was centrally applied by a 200-ton
screw-type testing machi ne. The square specimen>. were placed on a thick steel plate, des-
ignated A in Figure 16.17. wh ich wa~ supported by a rigid steel base. Load wa~ appl ied
through a bearing plate 8. while plate' C were auached to it and to the bouom plate A in
order to facilitate measurement of axial shortenmg by four dial gages. positioned o n the
extension of the diagonals. as shown in the figure. This modern simple test arrangement is
very si mi lar to thut employed 50 yean. earlier at NACA Langley (see for example (16.55]).
but it was apparently quite ~atisfactory. The te~t re~uhs were found to be in good agreement
with Inoue's analysis, especially in the early pla;tic 1.one.
Return ing for a moment to the late fonies. one sho uld nme that the cri tical shear stress
of plates in the plastic r:angc was also studied then theoretically and cxperiment:tlly, for
example by Gerard [16.61]. In this >tudy Gerard c~tendcd the ~ecant modulus method pre-
viously employed for compression to an equivalent formulation for shear, using affinity
relat io nships (si milar to the o nes proposed by Swng ct al. in (9.26 1] for torsion of tubes)
for approximate determination of the ~hear stress-strain curve for the material from its axial
load stress-strain curve.
Gerard carried out a series of tests to conlim1 the usc of the shear-secant modu lus for
plastic huckling of plates suhjcctccl to shear. The test techn iq ue employed was introduction
of shear to a Hat panel in a loadi ng fmme. The shear loading frame design (see Figure 16. 18)
was similar to a much earlier one used by Coker and Filon m Cambridge for their photo-
elasticity studie• [ 16.62]. In the simple frame (Figure 16.18) the aluminum alloy sheet spec-
imens were cla mped between rigid steel strips along the long loading edges and were sup-
ported alo ng the sho rt un loaded edge' by clamp> that only prevented rotation. Thu~. the
shear distribution was undisturbed whtlc the clamps provided torsional '>upport.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1315

(1:1~1. 16.58 )
( R~rf , '16.46)
Sit est,
(R•f. 16.26 )
kol

40

.004 .008
Stroin

Figure 16.16 NACA Langley square lUbe tests representing long plates in the late forties-correlation
of adjusted test results for the ct1tical stress u" with calculated plate buckli ng curves for
long, simply supported, nat plates of 2014-T6 aluminum alloy, employing the plastic
buckling theories available at the time (from r16.55])

Theoretical considerations of the opposing effects of nonuni formity of stress distr ibution
and that of the aspect ratio on the critical shear strain of the panel produced an overall
cotTection factor for aspect ratio. Calculations then indicated that for aspect ratios larger than
6 this correction factor was less than 3 percent and cou ld therefore be neglected. Hence, all
test panels (as shown in Figure 16.18. the frame accommodated two panels simultaneously)
were made wi th aspect ratios of 6 or larger. This also minimized the bending stresses im·
posed.
Though theoretical calculations predicted that the supporting strips should provide full
clampi ng to the plates, and therefore a shear instability coefficient K, = 8.1, calibration of
the shear loading frame indi cated that the average coefficient was K, = 6.7. It appears,
therefore. that the torsional restraint was less than that required for complete clamping. Also,

Figure 16.17 University of Tsukuba, .Japan, plastic buckling experiments on rectangular steel plates-
test seiUp for welded box-secLion stub-columns that each simulate four simply-supported
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

plates (from [16.601)


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1316 Plastic Buckling Experiments

Figure 16.18 Gerard's shear loading frame for plates-schematic (from


[16.61])

when vibration correlation techniques were employed (primarily in shells) to determine the
actual boundary conditions (see Chapter 15), nominal clamped supports were found to pro-
vide only incomplete clamping.
The test specimens were 2024-0 aluminum alloy plates of nominal thickness 0.064 in.
(1.63 mm) and (bit) = 45-103, with aspect ratios 6-10. The material was used in its an-
nealed condition because then the plastic range was large, which facilitated evaluation of the
shear-secant-modulus method. The critical stress and strain were determined from the bend-
ing strain, obtained as usual from the difference in the readings of the strain gages attached
on each side (at the center) of both panels.
The test results verified the empirical shear-secant-modulus method for determination of
the critical shear stress for plates in the plastic region. Though the method was checked only
for one aluminum alloy 2024-0, Gerard recommended its use also for other aluminum alloys
and other materials with similar stress-strain curves, taking into account the earlier extensive
validation for axial compression (for example in [16.45] and [16.55]).
The study of plastic buckling of plates under shear was also a pivotal topic in more recent
experiments on clamped annular plates in pure shear. These experiments, mentioned in Sub-
section 16.1.3 as a pronounced example of the plastic buckling paradox, were carried out
by Bauer at the Institut fUr Umformtechnik of the University of Stuttgart, Germany, in the
late eighties [16.40]-[16.42]. Their motivation was the development of a nonconventional
test technique for determination of the stress-strain curves of thin sheet metal, the plane
torsion test.
This plane torsion test, which evolved in the seventies and eighties (see for example
[16.63]-[16.66]), permits determination of stress-strain curves of thin sheet over a wide range
of strains, as occur in deep drawing. For testing, the specimen, which does not have to be
circular, is cut out of the sheet to be tested and is clamped as shown in Figure 16.19. The
pair of ring-shaped outer clamps and the concentric pair of circular inner clamps define an
annular plate specimen. The application of a torque to the inner or outer pair of clamps,
while the other pair remains stationary, produces a hyperbolic radial shear stress distribution
in the annular unsupported portion of the specimen. Circumferentially, the shear stress dis-
tribution can be considered to be uniform. After yield is reached, the maximum stresses at
the inner edge of the annular plate specimen cause plastic deformation (in simple shear). As
the relative rotation of the clamps continues, the plastic region spreads radially in a manner
depending on the work-hardening properties of the material tested. Finally fracture occurs
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

at the inner edge, where the maximum stress appears.


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1317

Figure 16.19 Plane torsion test for determination of stress-strain curves of thin sheet-schematic di-
agram of clamping arrangement (from [16.41], by permission of the Society of Manu-
facturing Engineers)

The torsional testing of the plates is carried out in a torsion test device shown in Figure
16.20. The specimen is centered in it by a pin and is clamped separately by the inner (4)
and outer clamps (5). Setting a pressure gage adjusts the hydraulic pressure, which can exert
a maximum clamping load of 80 kN. Thea a geared motor, with a maximum torque of 1570
Nm, rotates the left outer clamp cylinder slowly at 2 rpm via a frictional coupling that can
be adjusted to slip at a predetermined torque. The torque is measured on both sides of the
specimen underneath the inner clamps by two quartz load washers (3), which also gage the

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
axial clamping load.
During the test two angles of rotations are continuously measured: that of the outer clamps
relative to the stationary inner clamps and the angle of rotation of the specimen at a radius

Figure 16.20 Stuttgart University plane torsion testing device: (a) schematic diagram of the device, (b)
working section showing the main features (from [16.42]): (1) specimen, (2) pretension-
ing bolt, (3) two component quartz load washers (for measurement of torque and axial
clamping load), (4) inner clamp jaw, (5) outer clamp jaw, (6) engager plate, (7) drum for
direct measurement of the angle of rotation
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1318 Plastic Buckling Experiments

slightly larger than the inner boundary. The former is recorded on a potentiometer shaft
encoder, which is driven by a friction wheel rolling on the left end of the outer cylinder;
whereas the latter angle is registered by a special drum (7) having an engager plate (6) with
pins gripping directly onto the specimen. That angle is transmitted by a bevel gear train to
a second potentiometer shaft encoder. Both records of angles are digitized, stored and re-
duced on a PC after the completion of the test.
The stress-strain curves can be obtained from the test data by two methods: spiral eval-
uation and the continuous torque-angle of twist method. The former method transforms the
point of the distorted radial lines (the spirals) into points on the stress-strain curves, whereas
the latter evaluates M-e curves at a given radius. The second method, the continuous torque-
angle of twist method, appears to be preferable since it requires less effort and lends itself
to automated data acquisition and reduction with a PC.
Since the shear stresses in a plane torsion test of thin sheet cause a principal tensile-
compressive stress state in the unsupported area of the specimen, buckling may sometimes
occur. In this case, however, the plane stress state assumed is no longer valid and stress-
strain curves can no longer be obtained from this test.
Buckling of the specimen under shear (or wrinkling, as it is sometimes called in manu-
facturing), which limits the applicability of the plane torsion test, has therefore been inves-
tigated theoretically and experimentally [16.40]-[16.42]. An energy method was employed
for the buckling analysis, which, through introduction of the reduced modulus, also included
plastic, or partially plastic, buckling. Recalling that the determination of the stress-strain
curves presumed the sheet to be in the plastic regime, one notes that elastic buckling is
really not relevant here to the limit of applicability of the plane torsion test, but only plastic
buckling is. Calculations on typical sheet materials and thicknesses have indeed shown that
yield always preceded buckling.
The buckling tests, which are of primary interest here, are carried out in the same testing
device as the plane torsion test (Figure 16.20). The postbuckling shapes of typical thin
annular plates tested in torsion are shown in Figure 16.21. Whereas in elastic buckling the
maximal deflection would occur at about the middle between the inner and outer boundaries
(r 1 and r 2 , or a and b), in partial plastic buckling the buckling wave appears primarily in the
plastic region (near the inner boundary), whose stiffness is lower and where the maximum
stresses occur.
The buckling tests presented some practical problems. First, the serrated jaws of the clamps
that press into the specimen to make the transmission of torque possible produced a plastic
deformation in the specimen that affected buckling, usually resulting in a small but undefined
reduction of the critical torsion moment. To minimize this effect, the buckling tests were
carried out with the smallest possible clamping force, though this caused increased relative
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

motion under the inner jaws.


Secondly, since the initiation of buckling could not be observed visually during the tests,
the load on several specimens was increased stepwise with the aid of the adjustable frictional
coupling. After the tests, four different conditions of the specimens were ascertained: (1) no
buckling, (2) initiation of buckling, (3) buckling and (4) shear failure without buckling waves.
The results, classified in this manner, are plotted in diagrams like Figure 16.22. Complete
flatness was required for the category "no buckling," whereas the first signs of buckling
initiation (minute unevenness) classified the test results as "initiation of buckling." The
buckles usually appeared one following the other, and from a certain level of torque the
number of buckling waves remained constant till failure. The category "buckling" repre-
sented what is usually referred to as "postbuckling." The last category, "shear failure without
buckling waves," did not actually represent a relevant classification, since the shear failure
strongly depended on the notch effects of the inner clamps. It can, however, be used as an
indication of the trend towards buckling i:nitiation since it defines a torque up to which no
buckling has occurred. By its definition, "initiation of buckling" was the primary yardstick
for comparison with small-deflection theory.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1319

b • 3 5 mm

~ 80~ m

------

a ll.')_ml&l

10 •m

figure 16.21 Stuugan Uni,crsit) plane torsion tc;b-initial p<l\tbuciJmg shape; or tlun annular 98.7
W aluminum plate; (courtesy or Dr. M. Bauer)

2$01:: f'.OP•''""•n\.
l 'ii f'Qr IQI urf' • 'lhOVI
" t"i"&ouon or C..I( O. Itrtg Du~ o l"9 •O'Vf'S
• bucktirg o r.o bwcO. n9 'I

... .,T
~ '"'110\iOr cr b\IC.cllr.g
r•
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

60 .. Jso

.o
z"


~
~
u
roo

~l~l ~~==··'
0
.!?
olvn .nul'!\ AI 88,7
v so 1: 1.0 mm
'a•lS rr:m

0 ~~.--~.

(a)
S I~ ·a 12,5 t$
radius of inner clom p ri
11,5 mm 72.5
( b)
• 1,S tO ll.S 15 1~'5 "'~'~"~ 2'2,5
rod us of inner clomp 'I

F igure 16.22 Stuttgart University plano torsion buckling tests- experimental and cnlculatcd critical
torque versus radiu' or inner clamp r, for tWO group< nr aluminum AI 98.7 specimen<
1rnmt 116.42]): (>) 1 0.5 mm ond r 35 mm. (b) 1 1.0 mm and r, ; 35 mm
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1320 Plastic Buckling Experiments

The buckling experiments included specimens made of three materials, aluminum AI 98.7,
steel St 1403 and brass CuZn 36, of two thicknesses, t = 0.5 mm and 1.0 mm. The outer
radii were r, = 35 mm and 180 mm, and the inner ones were r1 = 7.5, 11.5 and 20 mm.
The experimental results, critical torque versus radius of inner clamp, according to the
four categories, are shown in Figure 16.22 for two groups of three aluminum specimens,
together with Bauer's theoretical predictions for partial plastic buckling. The tests were re-
peated with increased applied torques, as can be seen by the number of symbols for each
specimen. Bauer found that the experimental results could be reproduced rather well. Buck-
ling was observed for all the aluminum specimens of 0.5 mm thickness (Figure 16.22a),
including those not shown here (which appear in Figure 67b of [16.42]). For the thicker
aluminum specimens oft = 1.0 mm (shown in Figure 16.22b), no buckling was observed
for the smaller inner radii, r 1 = 7.5 and 11.5 mm, but shear failure occurred, as classified
in the fourth category. Buckles appeared only for the larger inner radius, r1 = 20 mm. Again
this was also observed in the specimens with the larger outer radius, ra = 80 mm, not shown
here (see Figure 68b of [16.42]). Similar results were presented in [16.42] for the steel and
brass specimens.
It may be recalled, that, as pointed out in Subsection 16.1.3, Ore and Durban [16.39]
compared their calculations with Bauer's experimental results. They found good agreement
with their deformation theory predictions but considerable overestimation with J2 flow theory,
which they considered to be additional support for the use of deformation theory in plastic
buckling.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Before closing this section, we should note that many of the plate buckling experiments
discussed in Chapter 8 also extended into the plastic range, when the strength of the plates
or their failure were investigated, as was the case in columns. Many of the points discussed
in Sections 8.2 and 8.4 therefore also apply to plastic buckling of plates. For example, the
important Cambridge University plate buckling experiments (see Subsection 8.2.4) and the
related elasto-plastic large deflection analyses, or the University College, Carditi, strength
tests on plate girders under shear (see Subsection 8.4.3) and the related failure sway mech-
anism, are relevant here too. Hence they, and many others discussed in Chapter 8, should
be consulted by the reader of the present section.

16.2.5 Plastic Mechanisms in the Buckling of Thin-Walled Steel


Structures
Thin-walled steel structures of approximately prismatic form, which carry axial or bending
loads, fail by first developing local buckles which eventually change into local plastic mech-
anisms (see for example Figure 16.23). Whereas solid structural elements (like structural
frameworks) develop only global plastic mechanisms at failure, thin-walled structures differ
by developing local plastic mechanisms prior to failure or by failing with combinations of
local and global plastic mechanisms.
The postbuckling or postcollapse behavior of a structure determines its ability to withstand
overloads, the effects of initial imperfections and those of load eccentricities. In order to
evaluate this ability of a structure, sometimes called its "toughness," one has to study its
postcollapse behavior, a task that can be conveniently done in an approximate manner by
rigid-plastic theory. For thin-walled structures, local buckling and postbuckling strongly in-
fluence their collapse behavior and toughness and therefore the relevant local plastic mech-
anisms have to be included in the investigations.
In the seventies and eighties the widening usc of thin-walled structures, primarily in cold-
formed thin-walled steel sections, motivated extensive study of their local plastic mecha-
nisms, in particular by Murray and his co-workers at Monash University in Melbourne (see
for example [16.67]-[16.72] and [4.19]).
If one derives the rigid-plastic curve !that approximates the postcollapse behavior of a
structure, it indicates the toughness and ductility of that structure. When the curve is plotted
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1321

( f)

Figure 16.23 Some observed plastic mechanisms in a sctics of tests on channe l-columns. (a)-(e) (from
[16.67)). (f) a view of a CF I mechanism (courtesy of Professor N.W. Murray)

together with the corresponding one fo r elastic behavior, the combination also permits a
rough esti mate of the failure load, as can be seen in the examples shown in Figure I 6.24
for a simple pi n-ended column (a) and fo r a plate simply supported along its unloaded edges
(b).
In Figure J6.24a, the elast ic curve is plo ued for an ini tial imperfection D;- The observed
behavior of an actual column (with the same imperfection) is seen to depart from the elastic
curve at B, when the first fiber reaches the yield stress (the load at which B occurs can easily
be calculated). The colu mn reaches its maximum load P1 at C, after which the load carried
decreases and the observed behavior curve drops and approaches the theoretical rigid-plastic
curve asymptotically. It should be noted that the rigid-plastic curve determines the limit of
load carrying capacity and complements the clastic curve. Their intersection A represents an
upper bound to the failure load and indicates approximately the deflection at which it occurs.
The rigid-plastic cun•e (whi ch closely approximates the o bserved behavior after the collapse
load) is a g raph of load-catTyi ng capaci ty and therefore indicates the suddenness of col lapse.
The load that the colu mn can sustain beyond the col lapse load C, say at D is represented
by DE. The applied load FE is, however, larger and therefore an out-of-balance force FD
appears that accelerates the load. Hence, if the rigid-plastic curve (which approximates the
actual one) droops steeply, the accelerating force increases rapidly and collapse will be
abrupt: but if the rigid-pl asti c curve droops Jess steeply, collapse will be more gradua l.
The behavior in the case of the plate in Figu re I 6.24b is simi lar. Though the elastic
postbuckl ing behavior is stahle. yielding begins at B. and the rigid-plastic curve agai n de-
termines the maximum load carrying capacity at the fai lure ai and beyond. as well as the
character of the collapse.
S ince the rigid-plastic cu rve appears to provide a reliable indication of the character of
the collapse of a structure, Murray and his co-workers set out to derive plastic collapse
curves for practical types of th in-walled structures. When such structures are tested to failure
in the laboratory. they seem at first to collapse by a confusing variety of plastic mechanisms,
as for example the five mechanisms observed for thin-walled channel colu mns, shown in
Figure 16.23. Because the minor axis bendi ng of this type of structure has a nonsymmetric
cross-section, it can be expected that two columns that fail in different directions will exh ibit
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1322 Plastic Buckling Experiments

0
<!
0
-'
-'
<!
X
<!

(a) ~6~ CENTRAL DEFLECTION 6

{(! .----Elastic th~ory

<
Ri91d- plastiC
fr1eory .ior perfec~ plcte

I
"'- ,~Eiast1c
Vl "'-· // 1heory for imperfect
Vl 11
plcte
w
0:: Oj
f-
Vl

-'
~
"" Ot:served
X behaviour 1
<!
I
I
I
I
E
(b) CENTF?AL DEFLECTION 6

Figure 16.24 The observed behavior of (a) an imperfect column and (b) an imperfect unidirectionally
compressed plate is closely related to the corresponding theoretical elastic and rigid-
plastic curves (from [16.67] with some changes). After the maximum capacity of the
structure is passed at C, the out-of-balance force or stress FD is an acceleration force

different collapse behavior. This indeed is the source of the difference between the first two
plastic mechanisms of Figure 16.23 (CWl and CW2) and the last three (CFl, CF2 and CF3).
In the mechanisms CWl and CW2 the web is in compression, whereas the flanges eventually
carry some tension, which results in the plastic mechanism to be located predominantly in
the web. On the other hand, in CFl, CF2 and CF3 the plastic mechanism is mainly positioned
in the flanges. A photograph of a typical CFl mechanism is shown in Figure 16.23f.
The Monash University investigators showed (see [16.67]) that the observed mechanisms
for channel columns and other thin-walled structures can be analyzed by combinations of
eight basic mechanisms and three fully plastic zones that can be made to satisfy compatibility
and equilibrium. This alternative approach is rational though approximate but also much
simpler and significantly less costly than the elasto-plastic finite-element analysis into the
large-deflection nonlinear range, which it aims to supplant for design purposes. The impor-
tance of this approach is that it represents an alternative (and relative simple) approach,
based on the dominant physical behavior of thin-walled steel structures at collapse, as ex-
hibited in the laboratory.
Tests on many thin-walled structures, such as box girders, box columns, !-columns and
channel columns, led Murray and his co-workers to the conclusion that parts of the plastic
mechanisms observed in quite different structures had essentially the same shape. Hence (see
for example [16.70]) the whole plastic mechanism at collapse of a thin-walled structure could
be considered an assembly of "basic mechanisms, which fit together and deform in a manner
compatible with the deformation of one another, and, of course, with the deformation of the
plastic mechanism as a whole." --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1323

The first step in the development of the analysis was the derivation and cataloguing of
the load-deflection relationships of the basic mechanisms (see [16.67]). The eight basic mech-
anisms shown in Figure 16.25 appeared to suffice for the analysis of most plastic collapse
mechanisms. The second step in the analysis was the combination of the necessary number
and types of basic mechanisms to form the complete plastic mechanism observed in the
various structures, which usually involved considerations of spatial geometry. It was often
found that the basic mechanisms in themselves were not sufficient to form complete collapse
mechanisms but required the introduction of regions of continuous yield deformation, regions
that changed their shape to allow the complete mechanism to deform. An example of this
is shown in Figure 16.26b, representing the plastic collapse deformation of a stiffened plate.
The region ACED in the figure, which was originally square, is such a yield zone that had
to undergo an in-plane deformation to become a rhombus. There also has to be a corre-
sponding tension yield zone FEHG and a compression yield zone ABEF in the stiffener,
with a pivot line at their common boundary EF, to allow the formation of the pattern. Such
a mechanism is referred to as a quasi-mechanism, because in addition to the assembly of
basic mechanisms it also requires these yield zones. Collapse mechanisms that can be made
by simply folding along hinge lines, without the additional yield zones, are called true
mechanisms, as for example the thin-angle strut shown in Figure 16.26a. A true mechanism
is developable from the flat sheets that make up the cross-section of the column or beam.
The load-deflection relationships of the eight basic mechanisms, which are true mecha-
nisms, are listed in Figure 16.25 and have been derived by Murray and Khoo in [16.67] (the

(P,nn~ Prnned~ ~
P • 0"0 HB~-
2
-e]
41 lie 0 e = B/2

A P • CJOHBL-Jt.~.f+t - ~J
2 ~I ! I~ 0 e = B/2

3 41 \,sie/
~/ 11. vA P• ~~~~)2+1-~+iSL'ltnH2"'t+I+~J]
2 K;H K' H 26 Jo;H 1K'1H

OoH= s- -~,
3 2 2
Pe
12"2
[{(-2tqH"J2+I}""-1- ~-~~H
2 "y'J

~
41
a'We ~.uv
4
\\ ' I
I
Obtain solution by using the difference of
; l two Type 3 mechanisms

p • P, - p2
Pe·P1e 1 -P2e2
A
5 "-.I ·,~:/
End panels twrst freely
I~ W Same equat1ons as for Type 3 but replace
K", by K'z

Same equations as for Type 3 but replace


K', by K'3

H"'
7
£....l"-.,.__'~5'--,·__4~~;r~bf V Same equot1ons as for Type 5 but with
;3. 45"

B -"--.1 (; I~
End pone!s twist and
bend freely

e • B/2
Positive pLastic h;nge
Negative plastic hinge
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Figure 16.25 Table of true basic plastic mechanisms (from [16.67])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1324 Plastic Buckling Experiments

(a)

(b)

Figure 16.26 Examples of (a) a true mechanism and (b) a quasi-mechanism in the collapse of thin-
walled structures (from [ 16.67])

reader may refer to this reference for the details, or to [16.70] for some of them). In Figure
16.25, B is the width of the beam or column, His its depth, P is the axial load, ~ is the
central lateral deflection and {3 is the inclination of the plastic hinge, when this hinge is not
perpendicular to the direction of the axial thrust P. The principal effect of the change in the
orientation of the hinges is to increase the bending moment required to form the hinge. The
effect was studied by Murray in [16.73] and found to depend on {3 2 • The relevant factors
that appear in Figure 16.25 are K 1 = I + sec 2 {3, K 2 = sec 2 {3 and K 3 = 2 sec 2 {3. Finally,
the positive plastic hinges indicated in the figure are the ones that rise up out of the plane,
and the negative hinges are those that deflect down below the plane of the plate.
The load-deflection relationships for the first two basic plastic mechanisms of Figure 16.25
were derived by consideration of the equilibrium of one-half of the mechanism after for-
mation of the plastic hinge (or hinges), and then solution of the resulting quadratic equation
for P. The third basic mechanism was analyzed by consideration of the equilibrium of an
elemental column of width dB and integration of the resulting expression for dP over dB.
The load-deflection relationships for the fourth to seventh basic mechanisms were obtained
in a similar manner. The analysis of the eighth mechanism, the so-called "flip-disk," is also
similar, except for two approximations, one assuming the shape of the disk to be a pair of
parabolas, and the second assuming integration by Simpson's rule.
The flip-disk plastic mechanism is an example of curved plastic hinges. Indeed, plastic
hinges of a true mechanism do not have to be straight lines, provided that some elastic
twisting and bending of the plate elements is also allowed. The end panels of the flip-disk
mechanism "twist and bend elastically to accommodate the changes in geometry as the
mechanism deforms. At the edge of this mechanism it is also necessary to introduce a fully
plastic region of very small dimensions which can be referred to as singularities, because
they are regions of infinite strain."
The experimental results obtained for channel columns by Khoo at Monash University
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

[16.74] were compared with analyses by combinations of basic mechanisms and local yield
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1325

zones, as outlined. Figure 16.27a shows the comparison of the curve based on the CWl
plastic mechanism of Figure 16.23 with the results of two tests CA 1 and CA2 (from [16. 74]).
The specimens were loaded axially through their centroid and tested in a deflection-controlled
machine. The local plastic mechanism theory gives a lower limit to the collapse capacity
and underestimates the load-carrying capacity of the channel columns. This is probably due
to the actual flip disk not being located at midheight of the column, its position having been
determined by the point at which yielding commenced, from which the mechanism developed
and cannot move away. Furthermore, because of the development of large strains in the yield
zones in the flanges, strain-hardening (which was neglected) should have been allowed for
as an increased yield stress (possibly 20-30 percent higher) in the fully plastic region. How-
ever, though underestimating, the local plastic mechanism theory presents the observed col-
lapse behavior correctly, whereas the upper curve of conventional rigid-plastic theory (simple
plastic hinge curve), which ignores the local mechanisms, grossly overestimates the load-
carrying capacity of the column.
In Figure 16.27b the experimental results for two nominally identical deeper channel
columns, loaded axially through their centroid, are compared with the local plastic collapse
curve based on the CFl plastic mechanism of Figure 16.23. Fairly good agreement is ob-
served, whereas the upper curve for a global conventional plastic hinge (formed at midheight
of the channel) largely overestimates the carrying capacity of the column.
Axially loaded box columns represent another example of plastic mechanisms developed
thin-walled steel structures. Two types of plastic mechanisms have been observed in box
columns: (a) the flip-disk mechanism and (b) the roof mechanism (see Figure 16.28). In a
series of tests on columns of square box section at Monash University [ 16. 72] it was found
that the type of local plastic collapse mechanism was determined by the (bIt) ratio of the
side of the column. When (bit) > 80 the columns consistently developed a flip-disk mech-
anism and when (bit)< 60, a roof mechanism (a shape of a roof on two opposite sides and
a similar shaped inwards depression on the other sides). Between these two limits either
plastic mechanism could occur. Since the local plastic mechanisms that developed were
known to be a result of the earlier elastic buckles, an elastic analysis for a square plate with
an initial central deflection was carried out. The calculations for variations of (bIt) and initial
deflection showed that for low values of (bit) yielding nearly always initiated at the center

•o,--,~----------------~ lOr----------------------,

0 g 0 g

0 8 08

07 HetQht of chonnel 07
column - 400 mm
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

b1 • 102mm
b1 .. 18 32mm He1Qhf of channel
\0 mm column .. 400mm.
269 MPa. b • 102mm
d • 38 33rnm
0' I"" 1 Omm
ay ..269 MPa.
03

0' 0'
L> CA I
.., CA2 01
o CA 7
0'
• CAB

~L,---L--~---L--~,--~----:

MAXIMUM OVERALL LATERAL DEFLECTION, 6


MAXIMUM OVERALL LATERAL DEfLECTION, 6
(mm)
(mm)
(a) (b)

Figure 16.27 Comparison of rigid-plastic collapse curves, based on simple (global) plastic hinges and
on local mechanisms, with Monash University experimental results for channel columns
(from [16.67]): (a) curves for CWl plastic mechanisms compared with tests on two
channel columns, (b) curves for CFl plastic mechanisms compared with tests on two
other channel columns
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1326 Plastic Buckling Experiments

(d)

Fi~'llrc 16.28 The two plastic collap'" mechanism; ob\crved in axwlly loaded tlun-walled box columns:
(a) the flip-disk mechani>m. (b) the mof mechanism. (c) comer tonk' occurring on load·
ing beyond maximum load for Oip·dl'k mecbam;rn' (from [ 16.72[): (d) ,·jew> of two
large square tubes with differeD! (b/1) ratios after testing in axial compression. the one
on the left is thinnco· nnd has develorcd two Jlip di>k mechanisms, while that on the right
is stockier and has developed a roof mechanism (counesy of Professor N.W. Murrny)

of the plate, while for high value~ of (b/t) it commenced at the edge. h appeared, therefore,
that the location of the point of lir.>t yield detennined the pm.ition of the local plastic mech-
anism, which did not change in location or form as it c.Jevelopcd.
For the square-section column these resu lts have the following implications [ 16.72): "The
roof mechanism is much more ductile than the nip-disk. i.e. n column with a roof mechani sm
does not unload so rapidly after the maximum load has been attained. Thi; is becau;e it is
a quasi -mechanism. meaning that some areas of the plate have to yield 10 allow funher
deformation to take plnce whereas the fl ip-disk is almost a true mechani sm. (A smal l nmount
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

of plate twisting is all that is required.)"


ln the experiments it was also noted that the flip-disks on two adjacent sides of a box-
column did not have a common end point (sec Figure 16.28a). and the horiLontal axes about
which they rotated were separated by a shot1 distance vet1ically. On loading beyond the
maxim um load. these short corner segments of plate between adjacent disk~ folded to permi t
further axial shonening at the comers (see Figure 16.28c). These comer "kinks" were studied
in (16.75).
On the other hand, the roof mechanisms in adjacent sides were always found to be at the
same horizontal level (see Figure 16.28b). The roof mechanisms behaved in this manner
since here the triangular corner areas were fully yielded. accommodating axial shortening of
the comers.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1327

Figure 16.28d shows views of two large square tubes with different (bl t) ratios after testing
in compression. That on the left has a high (b/ t) ratio and has developed two flip-disk
mechanisms, while that on the right is stockier and has developed a roof mechanism.
A number of box columns were tested by Khoo [16.74] and Rawlings and Shapland
[16.75]. Again fairly good agreement was found between the experimental results and theory
based on local plastic collapse mechanisms.
Based on the study of the experimental collapse behavior of many box columns, Murray
[16.72] also advanced proposals for improved design configurations for columns.
One may conclude that the approach of the Monash University group, which involves
close correlation between the study of experimental behavior and relatively simple analytical
models, seems to have considerable potential for future developments. One should recall,
however, that this approach is one of many, as indicated for example in Calladine's study
of the strength of thin plates in compression [16.76] or in the evaluation of similar experi-
ments on local buckling of box columns carried out by T. Usami and Fukomoto at Nagoya
University, Japan [6.12] and [6.107] or Ronda! and Maquoi at the University of Liege,
Belgium [16.77].

16.2.6 Plastic Buckling Tests on Cylindrical Shells under Axial


Compression
Before we discuss specific plastic buckling experiments on cylindrical shells, it may be worth
looking again at the general buckling phenomena in cylindrical shells, as considered in
Section 2.2 of Chapter 2, but here with emphasis on plastic buckling. There are two different
phenomena that constitute buckling: collapse at the maximum point in a load-deflection curve
(the limit load) and bifurcation buckling, both of which were clearly shown in Bushnell's
presentation ([16.3]; see Figure 16.29). The relatively thick, axially compressed cylinder in
Figure 16.29a (the photographs there are reproduced from [16.78]) deforms approximately
axisymmetrically along path OA until the limit or collapse load is reached at A. The perfect
cylindrical shell will fail at this limit load, following either the path ABC, along which it
continues to deform axisymmetrically, or some other path ABD, along which it first deforms
axisymmetrically from A to B, and then nonaxisymmetrically from B to D. At point A limit
point buckling occurs, and at point B the other type, bifurcation buckling, appears. The
equilibrium path OABC, which corresponds to the axisymmetrical mode of deformation, is
called the fundamental path, while the postbifurcation equilibrium path is called the second-
ary path. Buckling of either type occurs here at loads for which some or all the material has
yielded, as can be inferred from the photographs in Figure 16.29a. This example is somewhat
unusual in that the bifurcation point occurs after the limit point has been reached. In this
particular, uncommon case, therefore, bifurcation buckling is less important than the axisym-
metric collapse.
Usually the situation occurring is that shown in Figure 16.29b, in which the bifurcation
point B appears before the limit load A. The fundamental path OAC again corresponds to
axisymmetrical deformation and the secondary one BD to nonaxisymmetrical deformation.
The initial failure of the shell would then generally be characterized by rapidly growing
nonaxisymmetrical deformations. Here, therefore, the bifurcation point is of greater engi-
neering significance than the collapse (limit) load of the perfect structure.
One may recall that in real structures, which contain unavoidable imperfections, a precise
bifurcation buckling phenomenon does not exist, as can be seen in Figure 16.29b. Here tne
actual shell follows a fundamental path OEF, with a snap-through failure at point E at the
limit (or collapse) load .\,. The analytical bifurcation buckling model is, however, still valid
as a convenient approximation to the actual failure load and buckling mode.
To return now to specific plastic buckling tests on shells, it should be noted that as in the
case of plates, for shells many of the experiments discussed in Chapter 9 extended into the
plastic range and therefore many of the points raised there apply here too.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1328 Plastic Buckling Experiments

(Q) 1!1.1) $Mfl.I'TtN1111

lOAD,;,

c
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

TOTALDISPlACl\lt'<l
(b) toRR£S FOND II..C TO IDAD

Fi~:ure 16.29 Load-end shonening and load-dcOec1ion cur'c~ for pla.\tic buckling of cylindrical shells
under axial compression (from [16.3)): (a) load-end shomning curve with lirnll point A.
bifurcation point IJ and postbifurcation equilibrium path 80 (the photographs aro repro-
duced from 1'16.78]). (b) lc):td-deflection curves. a, would be obtained by a general non-
linear :tnalysis. for perfect and imperfect shell<. <howing bifurcation point /J, and limit
points A and £

As pointed out in the historical introduction to Chapter 9. the eatly buckling experiments
on cylindrical shells subjected to axial compression at the beginning of the century. ~uch as
those of Lilly [9.39)-[9.41 ]. most of those of Batling and Webb [9.49), or Robcn~on (9.51)
and (9.521 and others were actually plastic buckling tests because of their thick-walled spec-
imens (with low Rlt ratios). In the decades that fol lowed. aircraft tubes became thin ner and
the yield stress of their material higher. directing the allcnl ion to the intriguing problem of
clastic buckli ng. Only in the late thi rties did the investigators return to the study of plastic
buck ling.
O<good's NACA-sponsorcd "crinkling strength·· test~ carried out at the U.S. National
Bureau of Standards [ 16.79( were a good example of these plastic buckling experiments.
The tem included 2014-T4 aluminum alloy tubes and chrome-molybdenum Meeltubcs. but
mainly the former provided the plastic buckling test data that were later used by Gerard
[16.80]to substantiate hi s plasticity-reduction factors. Osgood's compressive {and crinkling)
tests were made in a hydraulic testi ng machine. and a detai led discussion was presented of
the ways to beware of and minimize load eccentrici ties ttnd rotat ion of the pluten clue to
clearances. Typical crinkli ng test speci mens after failure arc shown in Figure 16.30. exhib-
iting cases of axisymmetric and asymmetric buckling.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1329

figure 16.30 Osgood\ plastic compresshe buckling (crinkling strength) testS on 2017-T4 aluminum-
allo) and chrome-mol) bdenuon steel round au:crnft tubing-typical cnniJing test speci-
rnert.\ (from [16.79)). Note that three of them buckled axisymmetricall) and three asym-
metrically

Another example of test data used by Gerard for the same purpose is those obtained on
606 1-T6 aluminum alloy drawn circu lar tubes at the Aluminum Company of America (AL-
COA), again under NACA sponsorship. in the early fonies 116.81]. Gerard, however. pointed
out that more weight should be placed on Osgood's 2017-T4 test data since they included
also compression stress-strain curves, which the comparable Moore and Holt 6061 -T6 data
did not have . He also suggested that the drawn seamless tubes of (Rir) < 50 used in the
inelastic compression tests. and later in the corresponding torsion tests, contai ned only minute
geometric imperfections in comparison with those prevalent in the much thinner specimens
(often constructed from bent flat sheet) employed in the parallel elastic tests. This fact, he
emphasized, resulted in an apparent adequacy of small-deflection theory for buckling pre-
diction. provided a correct plasticity-reduction factor existed. This was an imponant obser-
vation at a time when imperfection ~en\itivity was not yet fully understood.
Plastic buckling te>ls on aluminum alloy and chrome-molybdenum steel seamless tubing
were also carried out for torsional loading in the same peri od under NACA sponsorship (see
for example 19.261], [ 16.8 1] and (1 6.82]). In most ca~cs the tubes failed as a result of
inelastic buckling in the two-lobe mode. The result~ for torsiona l buckling again verified
Gerard's plasticity-reduction factor;. the agreement being close for the 2017-T4 test data.
for which shear stress-strain data were abo given.
In the sixties. seventies and eighties plastic buckling of axially compres~cd cylindrical
shells was extensively studied both analytically and experimentally (see for example [16.38]
and [ 16.83]-[ 16.891). The well-known studies of Lee [ 16.83 ( and Battermann ( 16.38 ( in the
sixties focused on the evaluation of deformation and incremental theories for prediction of
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

buckling strengt h and postbuckling behavior. and their conclusions differ somewhat. as was
often the case in the discussion of the pla.\lic buckling paradox (see Subsection 16. 1.3). Here,
however. their experiments, with their emphasis on the buckling modes, are of prime interest.
Lee's specimens were made of drawn 3030-0 alumi num alloy tubing. a relatively soft
material, chosen because iL' properties magnify the dill'ercnccs between the buckling stresses
of cylinders of different geometries. The 4-in. outside diameter tubes were found by tests to
be generally free of residual stresses. The 10 specimens had different wall thicknesses, re-
sulting in (RII) ~ 9.4-29.2 and 46.1 for one shell. After machining, each speci men was
placed in an aligned lathe and iLs initial imperfection~ were measured wi th a dial gage, but
onl) the amplitude~ were recorded. a\ was customary at the time. Eight tensile and four
compressive test~ were carried out to obtain the material properties. and their results were
averaged to yield the representative ~tress-strai n curve. The specimens were inserted in an-
nu lar recesses in the head and base hlocks of the test fixture and tested at low strain rates
in a 200,000 lb (- 90 ton) universa l testing machine. Overall deformations were measured
hy three dial gages mounted appropriately on the test fixture. wltile a number of SR-4 strai n
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1330 Plastic Buckling Experiments

gages and two photoelastic plastic strips were attached to each specimen to determine the
longitudinal strains and their distribution.
Pure axisymmetric buckling modes appeared only in two of the three thickest shells with
(RI t) = 9.4, four ring waves in the shorter specimen with (L/ R) = 4.2 and five in that with
(LI R) = 6.3; whereas the third one of this group, the longest shell with (L/ R) = 10.5,
buckled in a combination of local and column modes. In the thinner shells, with (Rit) =
19.4 or 29.2, almost axisymmetric wave' were observed, 5-19 waves, depending on the
length of the shell, (L/R) = 4.1-10.2. Only for the thinnest specimen, with (Rit) = 46.1,
were diamond-shaped waves observed, indicating that at this (RI t) ratio a transition to a
nonsymmetric mode had occurred.
Lee concluded from his study that deformation theory provided a fairly accurate prediction
of the buckling strength but that it failed to give a correct description of postbuckling be-
havior. The incremental theory, on the other hand, was found to overestimate the buckling
strength considerably. Correlation between experimental and theoretical buckling modes was,
however, found to be difficult. The effects of initial imperfections were found to be signif-
icant, but not completely predictable from the measured amplitudes.
Batterman's specimens were all machined from 2024-T4 extruded aluminum alloy tubing
(of 3 in. outside diameter and 0.25 in. wall thickness). This alloy was chosen for its good
machining properties and because it work-hardened appreciably over the entire strain range
considered. Longitudinal tensile, compressive and bending specimens were cut from different
tubes and tested to provide average stress-strain data. After extensive preliminary testing to
determine the best type of specimen, a uniform tube type, to be tested fiat ended between
smooth bearing blocks, was chosen. The specimens were machined so that the ends were
parallel to within ± 0.005 in. ( = ± 0.013 mm). To minimize friction and permit horizontal
expansion of the ends of the tubes, three types of lubricants were tried in the preliminary
tests. The bearing blocks were sprayed in some tests with either a molybdenum disulfide or
a fluorocarbon dry lubricant, whereas in others Teflon sheets were inserted between the
bearing blocks and the specimens. However, no consistent difference between the three types
of lubricant could be observed.
The final test series consisted of 16 specimens with (Rit) = 9.7-89.3 and (LIR = 0.37-
2.92. Again all the thick shells, with (Rit) < 20, buckled in an axisymmetric pattern. The
thin specimens, with (Rit) > 56, buckled in a well-defined inward diamond-shaped pattern.
For intermediate shells with 44 < (RI t) < 56 a transition mode appeared, an axisymmetric
pattern near the ends of the specimen, with a gentle diamond-shaped pattern in the central
region. Near the lower end of the transition region (in one specimen with (R It) = 44.2) the
transition mode initiated as an axisymmetric pattern with some irregularities.
Batterman's predicted plastic buckling loads correlated well with the experimental ones
for (RI t) < 60. He concluded that predictions with 12 deformation theory and J 2 incremental
theory were very close, with the former always predicting lower buckling loads for a perfect
shell than the latter. It was also found that initial small geometric imperfections have only
a small effect on the plastic buckling load in the range of (Rit) < 60, and this was confirmed
by the comparatively small experimental scatter in that range.
The plastic buckling patterns of axially compressed cylindrical shells were the focus of a
careful, well-documented experimental study carried out by Horton and his students at Stan-
ford University in the mid-sixties [16.84]. The purpose of this investigation was to determine
the manner in which buckles generate and develop in thick-walled shells subject to axial
compression and the influence exerted by geometric and mechanical parameters on this
process.
The test specimens employed were uniform-thickness, right-circular cylindrical shells of
either aluminum, with R = 1.165 in. (=25.6 mm), brass, with R = 0.75 in. (=19.1 mm),
copper, with R = 0.437 in. ( = 11.1 mm), or stainless steel, with R = 0.375 in. ( =9.5 mm),
all with accurately squared ends. A wide range of (R It) and (L/ R) ratios was used in the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1331

parametric experimental study. Most of the tubes were tested in as-recehed condit ion. but
some were fabricated by turning from common tube stock to obtain a wider range of (R It)
ratios. The shells were tested in a 60.000 Jb (=27 ton) capaci ty universal test ing machine.
wi th ground steel plates interposed between its platens and the test specimen. In those tests
for wh ich load-displacement histories were recorded. a load transducer was placed between
the lower ground stee l plate and lower machi ne platen. Special care was taken to ensure that
the motion of the platen was vertical. In all the tests the loadi ng was pure axial compression.
applied at different rates of loadi ng.
Honon and his students cla~sified the buckling of axially compressed cylindrical shells
into five groups:
"1 . Those in which lite initial huckle is plastic and axi~ymmetric and remain> so with
subsequent loadi ng.
2. Those in which the ini tial buckle is plastic and axisymmetric but develops into a
nonsymmetric pallern of elliptic, rectangular. triangular or square cross section wi th
subsequent loading. In thi ~ clas~. there are essential ly two subclasses: (a) those in which
the nonsymmetry occurs immediately after the formati on of the first ring. and (b) those
in which a number of rings are fonned prior to the development of the non~ymmetric
pallem.
3. Those in which the initial buckle is plastic and is alway~ nonsyrnrnetric in character.
4. Those in which the initial huck le is elastoplastic in character, nonsymmctric in form
and which become fully plast ic with increasing load.
5. Those in which the init ial buckle is elastic in churactcr. nonsymmetric in form and
which become fully pl(lstic with increasing load."
Their experimental investigation (16.84], however. dealt only with shells of the first three
group~.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
The progressive developmcm of ring buckles in the tc~t of a steel shell. with (R/t) - 18.8.
belonging to the first class. is '>hown in Figure 16.31. Figure 16.31a shows the unbuckled
tube. The first ring buckle forms at the top of the specimen. In Figure 16.3 1b the fir~t ring
has llauened and the second ring buckle is beginning to fonn. When the condition of Figure
16.3 1c is reached, the third ring has developed. The nex l ~tage ca n be seen in Figure 16.3 1d

( b) (c)

(e) (I)

fi~ure 16.3 1 Stanford University pla-uc buckling test< on "eel cyhndrical shclh-devclopment of
ring buckles (from 116.841. courtesy of Profe;sor S.C. Bailey): (a) the unbuckled tube.
(b) complete 0:111ening of the firs t ring :ond initimion of the second ring. (c) third ring
developed. (d) ring buckles sta11 at the other end of the shell, (c) ring well developed at
both ends :ond central l'i ng buckle being formed. (f ) the ultimate developed ring paucrn

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1332 Plastic Buckling Experiments

when ring buckles stan to form at the lower end of the specimen. These develop and at the
next stage, shown in Figure 16.3 1e, the central ring buck le has begun to grow quickly. The
ultimate developed ring pattern is shown in Figure 16.3lf.
This axisymmetric ring pattern, which always occurs for very thick shells, was well known
when the Stanford experiments were initiated. As a matter of fact, it had already been
reported in I 928 by Geckeler [ I 6.90) and appeared in text hooks on elastic stability of shells.
Horton and his co-workers, however, set out to explore the nonsym metric patterns that they
maintained should occur as the (Rir) ratio increased in the transition from the plastic axi-
symmetric mode to the diamond-shaped elastic buckling mode.
Figure 16.32 shows the successive stages in the formation of one type of such an inter-
mediate pattern. Here there is definitely a pronounced tendency for the shell to buckle in a
two- lobe pauern, but the two lobes have not fully developed. Instead the circular cross-
section has become elliptical, this deformaticm bei ng twisted in successive layers. Hence.
one may note in the end view shown in Figure 16.32c that the ellipse in the fi rst layer is
orthogonal to the ellipse in the second layer and parallel to that in the third.
Another series of photographs (Figure 16.33) shows the development of an actual two-
lobed type of failure. another type of intermediate pallern. As the load increases, a ring
buckle initiates at the upper end of the shell (see Figure 16.33a). This ring buckle develops,
but then the specimen begins to depart from ci rcu larity immed iately below the ring and a
fold appears. The material flows into this plastic hinge line (the fold) and the cross-section
of the cylinder alters considerably. The two creases develop and llauen, and then inward
tr iangular indentations appear. in a plane normal to the two previous creases, and progres-
sively increase in size with load (sec Figure 16.33c). The process of folding, indenting and
development of creases continues until the shell has completely defonned (sec Figure
16.33d). This ultimate pauern consists of two-lobe layers in which the lobe direct ion of each
layer is at a right angle to that of the preceding one. Here the fi rst stage of plastic deformation
was the formation of an axisymmetric ring buckle that changed its shape under further
loading until it became a nonsymmetric two-sided figure.
In some cases the process is slightly different. as shown for a test of a copper tube in
Figure 16.34. A number of ring buckles are produced and remain circu lar. but. as the length
of the specimen decreases, instead of further ring buckles two-lobed failure pauerns appear
(see Figure 16.34b). Usually when this symmetric-nonsymmetric pauern occurs the two-lobe
pattern is not very pronounced but the sections distort from circles to ellipses (see Figure
16.34d).
For certain geometries and mechanical properties a three-lobe fai lure pauern appears (sec
Figure 16.35). This figure shows the progressive stages in the development of such a pattern
for a brass tube of (Rir) = 26.8. One can note the promi nent regularity in the folds that
appear. Three-lobe-type failures can occur after the fonnation of a single ring buckle or after

(b) (c)

Figure 16.32 Stanford University plastic buckling tests on cylindrical shclls- dc••cloprnem of nonsym-
metric pauem of rotated elliptic cross-sectional form (from f 16.84]): (a) first stage in the
development of the buckles, which appear at both ends to the same degree. (b) the section
beginning to twist and change shape from circular U) elliptical, (c) end view showing the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

elliptical cross-sections with layer-by-layer rotations


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plas tic Buckling Experiments 1333

(0) (b)

Figure 16.33 Stanford University plastic buckling tests on cylindrical shells-development of a non-
symmetric two-lobe panem of mtated rectangular cross-sectional form (from [ I 6.84]) :
(a) initial development of dng buckling. during which the specimen shows a tendency
to produce a number of ring buckles, (b) inward motion of shell wall has staned, (c)
wall motion occurring at a third point, this motion being in a direction parallel to the
initial one but nonnal 10 the preceding one. (d) end view of lhis buckled cylinder

a number of ri ng buckles have developed (see Figure 16.35b). The triang ular cross-sections.
rotated layer by layer, are clearly shown in Figure 16.35e.
A similar process. but with a four-sided nonsymmet ric pattern instead of the triangular

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
one, was also observed for a particular geometry, again with a layer-by-layer rotation of the
pattem , the angle of rotatio n bei ng now 7T/4 radians. Plastic buck ling with other multi -lobe

tal (b)

tel (d)

Figure t6.34 Stanfo rd University plastic buck ling tests on cylindrical shells-development of a
symmetric-nonsymmelli c pallem formation in a copper shell (from [16.84], courtesy of
Professor S.C. Bailey): (a) copper tube in which an end ring buckle has formed and is
beginni ng to develop into a two-ti ng pauern. (b) the change from axisymmetric ring
deformation to nonsymJnelli c two· lobe failure c an be clearly seen here. (c ) inward mo·
tion of the shell wall developed in a direction nc)rmal 1<-.> the previous one, (d) end view
Copyright Wiley showing b()lh ring andellipt ic sect ions
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1334 Plastic Buckling Experiments

{a) { b) (c)

{d ) (e)

Figure 16.35 Stanford University plastic buckling tcsts- dcvclopmcm of a nonsymme!Ji c pauem of
triangular fo rm in a brass tube (from [ 16.84]. counesy of Professor S.C. Bailey): (a) first
ring buckle fonns at the lower end. (b) second ring buckle has developed completely.
the inward motion of the shell wall is very apparent. there being clear evidence of a
tendency for lhe walls to diston in a manner resuhing in a triangular local CJ"Oss-section.
(c) one triangular layer has fully formed. and the second one is well developed and has
rotated, (d) fonh fold has developed and a fi fth is staning. (e) view showing triangular
cross-sections in the folds rotated layer by layer

patterns (pentagons. hexagons, etc.) can also occur. Exam ination of the nonsymmetric failure
patterns shows that, contrary to the ax isymmetric buckling patterns. which involve large
extensions and contractions, they essentially represent inextensional defo rmatio ns. wi th the
plastic deformatio n localized in the fold lines. The no nsyrnmetric plastic buckl ing panerns
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

actually bear a remarkable resemblance to the Yoshimura panerns (see Figure 9. I 5), into
which the usual di amond-shaped elastic buckl ing develops in tl1e deep postbuckling region .
T he S tanford researchers extensively studied the influence of geometric parameters o n the
character of the nonsymmetric plastic buckli ng pauern. For example, in Figure J6.36 they
examined the variatio n in (circumferential) buckle number, in 2024-T4 aluminum alloy
shells. as a function of (R/1) for constant R, (LI R), material and rate of loadi ng. They noted
that there were almost definite bo unds to the value of (Rit) for a prescri bed number of
buckles n. when the other variables were kept constant. They also analyzed in a similar
manner the effect of the other parameters.
These investigations of Ho rto n and his co-workers were presented in detail si nce they
represent a very careful and broad experimental program on plastic buckl ing of shells that
provided valuable physical insig ht into the nature of nonsymmetric plastic buckl ing modes.
Such extensive parametric experi mental studies are scarce and should be performed more
often to provide the basis for reliable computational tools.
Another comprehensive experimental study on nonsynunetric plastic buckling modes was
carried o ut a decade later at the Department of Engineering of the University of Cambridge
by Johnson and his co-workers [ 16.86]. Motivated by the potential of thi n-walled tubes as
energy-absorbing devices. they considered the nonsymmetrical collapse modes of the cylin-
drical shells with the aim of developing rigid plastic mechanisms for design calcu latio ns of
tubes. somewhat similar to the local plastic mechanisms developed at about the same time
for plates by Murray and his co-workers in Melbourne (see Subsection 16.2.5).
The Cambridge researchers generated two-lobe and three-lobe collapse mechanisms that
simul ated the plastic buckling 1;1ilure of cylindrical shells. The defo rmation in these mech-
anisms consisted of bending at fixed fold lines, while the middle surface of the shell remained
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1335

• L/R=3. 4~, 2024 T~4 Aluminum


R""l. 165m., Loading Rate 1200 lb/min

~ PROBAIH..E ARF..AS OF CHANGEOVER

U1
UJ
_J
::.::
u
::J
en
LL
0

lr 6
UJ
CD
:::;:
~ 4

20 40 60 80 100 120 140 R


l
RADIUS TO THICKNESS

Figure 16.36 Stanford University plastic buckling tests on 2024-T4 aluminum alloy shells-variation
in buckle number as a function of (R It) for a constant R, (LI R), material and rate of
loading (from [16.84])

unextended. They used paper models to demonstrate the mechanisms and showed how the
different modes depended on the geometric parameters of the shell.
They also proposed a progressive collapse mechanism involving the concept of traveling
hinges, which was based on a series of tests on rigid PVC (polyvinylchloride) tubes. Figure
16.37 presents the stress-strain record of a typical test, performed at a strain rate of 0.0028
sec_,, with seven representative deformation stages marked on it. Figure 16.38 then shows
the development of the deformation on seven rigid PVC specimens in axial compression
tests that were interrupted at various stages corresponding to these seven marked ones. In
the figure the photos of the deformed PVC specimens are followed by an equivalent series

10000
!tl---- tn1lJo\ peak stress

~ c Mean
post-buckling
-~ ~

J
5000
_j_'''"'

0
Compressive strarn
percent

Figure 16.37 Cambridge University plastic buckling tests on rigid PVC tubes-compressive stress
versus strain record for a typical thin PVC cylindrical shell (with Rlt = 6.75 and
LIR = 7.4) compressed between flat parallel platens moving at a rate of 0.5 in./min
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
(=0.0028 sec-' strain rate) (from [16.86])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1336 Plastic Buckling Experiments

1r 121 IJl t<l 1s: <6l ( I)

(ol

(b)

(2) (3) (4) (51

Figure 16.38 Cambridge Univcrsi1y plaslic bucking tests on rigid PVC tubes (from [16.86]): (a), (b),
(c) three views of a series of seven specimens from axial compression testS that were
inleiTUpted at various stages of the deformation process. corresponding lO those marked
on 1he stress-strain record shown in Figure 16.37. (d) an equivalent series of paper
models. folded to illustrate the inextensible collapse mode, following the passage of a
traveling hinge. (e) the same paper tubes, undcformed with the hinges marked on them.
1hus showing the tmveli ng hinge thai moves down the she ll. (f) diagrammal ic plan views
showing the progress of 1he traveling hinge AIJCD

of paper models folded to ill us trate the inex tensible collapse following the passage of a
travel ing h inge. and then the same paper tubes arc shown. undeformed, but with the h inges
marked on them. At the bo uom of the figure a diagram demonstrates the relevant trave ling
hinge. Johnson a nd his co-workers examined the character o f the traveli ng h inges in detail
a nd presented an analysis, based on a si mplified mode. for predi ct ion of the mean postbuck-
ling s tress. There was fair agreement between predicted and experime ntal values.
Very thick a nd very s hon PVC tubes, with (Rit) < 4.17 and ( L/R) < 0 .65, were observed
to fail by axisymmetr ic ring collapse. Symmetric ring buckling involves considerable exten -
sion and contrac tion of the s hell middle surface, and therefore on ly for thick-walled tubes
can the work required for this extensional deformation be less than that needed for bendi ng
the walls at the plastic hi nges. It was a lso no ted that the limiting (Rir) values for axisym-
metric ring buckling were generally higher for metal shells than those for PVC ones, probably
due to the more pronounced s train-ha rdening in metals that increased the resistance to high
bending stra ins at the hi nges. For example. they observed axisymmetric ring collapse in
alumi num shells for (Rit) = 15.6. three times thinner-walled than the PVC tubes that ex-
hibited ring buckles.
Copyright Wiley
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1337

Some of the theoretical resuhs of the early eighties (for example Tvergaard [ 16.89)) cor-
related fairly wel l with the experiment ~ of the sixties and seventies. which were therefore
ca lled upon to substantiate the computation~.
The frequent usc of stiffened cylindrical ~hells as components in offshore stntcturcs mo-
tivated extensive small- and larger-scale model buckling tc't~ in the eighties. Among them
were the buckling experiments o n I /20 th scale ring- and ~!ringer-stiffened axially com-
pressed cylindrical shell~ carried out at University College. London (sec [I 3.8 [. [I 3.65) and
[I 3.66). discussed in Subsections 13.2.2 and I 3.3.2 of Chapter I 3. There the emphasis was
o n the clastic buckli ng. but actual ly many of the test~ re lated to plastic buck li ng . Here the
local plastic collapse mechanism of the ring-st iffened steel cylinders is of prime interest and
warrants further allcntion.
These ring-stiffened shells were made of mild steel or high-yield steel (sometimes a com-
bination of both. with only the outer bays of high-yield 'tee]). welded in a ~pecial jig.
de~cribcd in Chapter 13 (see Figure 13.20). The geometry of the relevant ~hells was (R/t)
,. 150. (L/ R) = 1.0 and ring width-to-thickness ratio 8 or 16. The collapse mechanism. in
the advanced postcollapsc range. took the form of either a sharp axisymmetric outward bulge
(see Figure 16.39a) or of a sequence of inward dimples ro und the cylinder (see Figure
16.39b). The nonsymmetric collapse mode was clearly observed to develop in the post-
collapse region and was probably the rc~ult of a supcrimpo~ition of a sinusoidal bifurcation
mode onto the axisymmetric defonnation of the shell that characterized the behavior at
collapse and in the early stages of unloadi ng (like the ;,ituatton depicted in Figure 16.29a).
It was o bserved in the tests that the;,e axially compressed ring-stiffened steel cylindrical
shell s. typical of off-shore construction. invariably buckle and fai l in the vic inity of the end
supports o r one of the intermediate ri ng-stiffeners. The co llapse follows outwa rd displace-
ments. essentially uniform around the ~hell. caused by local bending actions induced by the

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
re!>lraint to Poisson expansion at the end\ or intermediate rings. It may be noted that here
collapse initiates in the typical a.~isynunetric outward bulge also in shells with (R/t) : 150,
a mode usually ob;erved in significantly thicker shells.
Walker and hi s co-workers pointed out that local imperfections in the regio n of the end
supports and the rings me therefore of rrimc importance in pia;, tic buckling. As in the elastic
case. the imperfections of the same shape as the collapse mechanism produce the most
adverse effects on the plastic collapse strength of the ~hell.
One may mention here a similar remark by Bushnell and Meller in their 1984 survey
[ 16.91]: "Fairly thick metallic cylinder;, (R/t < 100) arc not very sensitive to mitial random

(bl
Figure 16.39 Univcr,ity College. London, pla,l ic buckling 1e'" on steel ring-stiffened cyli ndrical ;hells
subjected 10 axial compression (councsy of Profe"or A.C. Walker): (a) three-bay cylinder
showing collapse in the fonn of a 'hall' axisymmetric outward bulge. (b) fi\c-bay cylinder
Copyright Wiley e'dlibiting collapse with a .cqucncc of inwanl dimple' round Jhe shell
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1338 Plastic Buckling Experiments

imperfections if they buckle at stresses above the material proportional limit. The axisym-
metric bulge which develops near an end ... represents a predictable 'imperfection' that
grows with load and is much more significant than any unknown imperfections due to fab-
rication or handling errors."
Bushnell and Meller also point out in this connection that "inclusion in the analysis of
radial restraint at the boundaries," first included by Murphy and Lee [16.92] in plastic buck-
ling load predictions, "essentially eliminates for these rather thick shells the discrepancy
between test and theory." They concluded that consideration of the effect of radial end
restraint would practically eliminate the plastic buckling paradox for axially compressed thick
shells, a somewhat optimistic prognosis that has not yet materialized.
Before we close this subsection, we should point out that inelastic effects may appear also
in shells of fairly large (RI t) ratios that buckle elastically if the material stress-strain curve
exhibits early nonlinearity (though its 0.1 percent yield point is above the buckling stress).
Such, for example, was the case for some stringer-stiffened cylindrical shells, made of a
relatively soft alloy steel, that were tested under axial compression at the Technion in the
early seventies (see [9.222], [9.228] and [13.1]). The buckling loads of these shells were
calculated by Mayers and his associates with a method they developed at Stanford University
[16.93]-[16.95], and found to be noticeably atiected by the nonlinear material behavior.
These material nonlinearity effects clarified some of the scatter noted in the experiments. In
general, such inelastic effects can be of similar magnitude to those of initial imperfections.

16.2.7 Plastic Buckling of Cylindrical Shells Subjected to External


Pressure
Many of the early buckling tests on cylindrical shells under external pressure, mentioned in
the historical introduction and other sections of Chapter 9, dealt with rather thick shells,
extended well into the plastic range, and could therefore be considered as plastic buckling
experiments as well (see for example [9.17]-[9.19], [9.28], [9.29] and [9.38]). This is also
evident for some of the stiffened cylindrical shells under external pressure discussed in
Subsection 13.4.1 of Chapter 13 (see for example [9.63], [13.61], [13.72], [13.104] and
[13.110]).
In particular, one should recall the extensive buckling tests on unstiffened and stiffened
cylindrical shells subjected to external pressure, carried out at the U.S. Navy David Taylor
Model Basin in the fifties and sixties. Some of these tests again extended into the plastic
range, while some were specifically designed as plastic buckling tests (see for example
[13.55], [13.57], [16.96]-[16.97]).
Boichot and Reynolds [ 16.97] posed important questions in the analysis of their results,
which justifies further consideration of their 1965 report. The investigation consisted of
external pressure tests of 69 small 2-in. diameter machined aluminum models, and its purpose
was to study inelastic buckling of near-perfect ring-stiffened cylindrical shells made of strain-
hardening materials. The specimens were machined from 7075-T6 aluminum alloy bar stock
to a configuration with six external frames (see Figure 16.40), covering a geometry range of
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(Rit) = 12.0-51.8, unsupported shell length ratio (L/R) = 0.80-4.47, middle ring spacing,
between four middle frames, (L1/ R) = 0.12-0.70, ring area ratio (A1 / tL1) = 0.2-0.8 and the
(Batdorf) shell geometry parameter Z = \/1 - v 1(F/Rt) = 0.58-3.6. (At DTMB a slightly
different shell geometry parameter 8 = \i'3VZ was used.)
The dominant form of collapse for such shells under hydrostatic pressure was presumed
to be axisymmetric inelastic buckling of the shell between stiffeners. The specific objective
of the experiments was therefore to provide the necessary test data, "through tests of small
machined models with near-perfect circularity having systematic variations in shell thickness,
frame spacing, and frame size-the parameters on which shell collapse strength was expected
to be critically dependent," in order to verify an axisymmetric plastic buckling theory, de-
veloped by Lunchick at DTMB, that took strain-hardening into account (see [16.98]). "It
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1339

END PLUG TYPICAL ON IFI MODELS


(WITH FILLET)

Figure 16.40 U.S. Navy David Taylor Model Basin inelastic buckling tests on 7075-T6 aluminum
alloy, ring-stiffened cylindrical shells under hydrostatic pressure-dimensions of speci-
mens (from [16.97]). Note the fillets on the F-models, and the decreasing frame spacings
near the ends

was believed that many small models mass produced at low unit cost and tested with no
instrumentation would bring a greater return than would a few expensive, elaborately instru-
mented models. In this way a wide parametric range could be studied at moderate cost The
models were designed to allow some overlap into the range where nonsymmetric buckling
(circumferential Iobin g) of the shell between stiffeners takes place." The transition to this
mode was expected as the frame spacing was increased or as the shell thickness was reduced.
"As the tests proceeded, however, it became evident that with the range of parameters
selected, failures could not be confined to the symmetric mode of shell collapse because of
the intervention of a third mode of failure. More often than not, the cylinders were found
to fail by inelastic general instability ... a mode which had not been believed critical because

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
of the relatively short lengths of the cylinders in question. That this had not been foreseen
was due largely to the fact that no method was then available for estimating inelastic general
instability strength."
Since then such methods of calculation have been developed, initially by the researchers
at DTMB [16.99] and [16.100], probably motivated by the test results, and then later as part
of the modern shell codes. The Boichot and Reynolds test results were therefore also used
to evaluate those DTMB methods. It was realized, however, "that because of the parameters
selected for study, those critically affecting general instability strength have received incom-
plete coverage. The effects, for example, of frame shape and cylinder length have not been
adequately explored."
Even a very wide parametric range may require additions to cover unexpected buckling
modes. But one should remember that the finding of new modes is one of the purposes of
a buckling test
The test specimens were indeed near-perfect, with a shell thickness accuracy of ± 0.5-2
percent An effort was made to minimize the detrimental end effects by decreasing the length
of the two end frame spacings (see Figure 16.40). Furthermore, to assess the influence of
stress concentrations on the collapse pressure, 22 of the specimens were duplicated geom-
etries except for the addition of fillets at the ring-shell and end frame-shell intersections, as
shown in the figure.
Each shell was tested to collapse under hydrostatic pressure in a 5-in. diameter tank,
employing oil as the pressure medium. The shells were filled with oil and vented to the
atmosphere to absorb the energy released at collapse. In the tests the pressure was applied
in increments, each being held for at least a minute. The increments decreased as the collapse
was approached, the last increment being usually Jess than 2 percent of the collapse pressure.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1340 Plastic Buckling Experiments

Some specimens collapsed as the pressure was applied, and others failed while the pressure
was held constant. At collapse, extensive damage occurred, and in many cases the shell and
frames were tom apart, the extent of this tearing being significantly less in specimens with
fillets.
Most of the shells collapsed in the general instability mode; the remainder, all except one,
failed in the axisymmetric shell mode. The one exception was a specimen that apparently
collapsed in the nonsymmetric shell mode. Hence, but for the occurrence of general insta-
bility, the specimen design achieved its aim since some overlap from the axisymmetric shell
mode into the nonsymmetric shell mode had been intended. A breakdown of the specimens
in terms of their geometric parameters and their modes of collapse is presented in Figure
16.41. Such a collapse mode distribution chart lucidly summarizes the relation between
geometry and collapse mode. For example, three of the thinnest shells with weak rings, on
the extreme right of the chart, collapsed by elastic general instability, with the rings deform-
ing only in their planes of curvature. In specimens that failed by inelastic general instability,
on the other hand, the rings also twisted and folded out of their planes.
Without going into the details of the evaluation of the inelastic analyses in the report
(which today would have to include more recent methods), it can be concluded that in general
the predicted buckling pressures were somewhat higher, mostly within 15 percent.
One surprising result was the influence of the fillets. Though it was expected that the
benefits of the reduction of stress concentrations, due to the fillets, would affect only the
interframe shell collapse mode and not the inelastic general instability collapse, the converse
was found to be true. The investigators could not point to a specific mechanism that would
explain why the presence of fillets increased the inelastic general instability strength. But
they made an interesting practical observation: that actual welded stiffened shells would have
fillets, and these indeed showed good agreement between prediction and test. They also
cautioned the designers that since the DTMB shells were relatively free of the weakening

•1 Calculations with•n 10 pe1cenl of test data.


Inelastic Jenera! instability Z Calculallons an~ test data drlfer by 10·15 percent.
J Calculations and lest data differ by mo" than 15 percent.
Both designations wrthrn a square rndicate a duplicated model.
Shell la~lure
the F signifying the one with fillet radii.

General rnstabilrty on one model


Shell far lure on duplrcate

Elastic eeoeral inslabrlit)·

Figure 16.41 U.S. Navy David Taylor Model Basin inelastic buckling test on 7075-T6 aluminum alloy,
ring-stiffened cylindrical shells under hydrostatic pressure-collapse mode distribution
(from [16.97])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1341

effects of imperfections and residual stresses, they might represent too perfect models, and
therefore further substantiation with less perfect shells having realistic imperfections was
essential.
Another example of plastic buckling experiments on thick-walled cylindrical shells under
external pressure is the tests carried out by Montague at the University of Manchester in the
late sixties [16.101].
The investigation included 12 mild-steel, relatively thick-walled and short cylindrical shells
of about 5.4 in. diameter. The specimens were designed to collapse inelastically and hence
their (Rit) = 22-55, except for one thin-walled shell with (Rit) = 94, which indeed failed,
as expected, by elastic instability. Two modes of plastic collapse were expected: one "char-
acterized by the appearance of circumferential lobes after yielding of the shell material, the
other by an hour-glass or waisted configuration, also following yielding of the material."
The experimental study aimed at finding relations between shell geometry parameters and
the mode of collapse to guide the choice of the appropriate failure analysis.
The specimens were machined from seamless tubes to within 0.020 in. of their final
thickness, and then left to rest for three weeks before the final finishing. This allowed any
large distortions, caused by the initial heavy cut, to dissipate before the final fine cut. The
finished specimens were also stress-relieved and cooled slowly. The shells had length ratios
(LI R) = 0.563-3.005; most of them were fairly short.
The tests were divided into two series, one of nine shells loaded by external radial pressure
only and the other of three specimens subjected to hydrostatic pressure. The test setup (see
Figure 16.42) consisted of a rigid thick-walled pressure vessel or body (1), into which the
specimen (4), together with an internal traversing mechanical probe for measurement of the
radial deflections (5), was fastened. The pressure vessel was fixed to a frame (9) and was
closed by a cover (3). For radial pressure testing, the axial load applied should be negligible.
Therefore, in the case of radial loading, axial tension (or compression) from an external
source is usually applied to eliminate the axial stress in the shell (see for example [16.103]).
Here, however, this was achieved by attaching the test shell at one end to a ram (2) that was
free to move in the axial direction. Thus, no axial forces were produced by any large de-
formation of the specimen. During the tests the cylinder and ram were in the vertical position
(as shown in Figure 16.42), whereas for draining the test rig was put in a horizontal position.
Hence, taking into account the weight of the ram and the known friction generated by the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 16.42 University of Manchester plastic buckling experi-


ments on thick-walled cylindrical shells subjected
to external pressure-test setup, schematic and in-
complete, in the vertical testing position, without
secondary systems (from [16.101], with omissions)

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1342 Plastic Buckling Experiments

seals around the ram at a specific pressure difference allowed the small residual axial stress
to be assessed. For hydrostatic pressure the ram was pushed to its extreme position and an
additional end plate was fixed to the exposed end of the cylinder tested.
The internal traversing mechanical probe (5) indicated the radial deflection on the external
dial gage (7) with a mechanical magnification of 5.33 (achieved with a level system 6),
which permitted reliable measurement of deflections as small as 0.0002 in. ( ~0.2-0. 7 percent
of the wall thickness), a simple yet accurate device. The traversing system of the probe,
consisting of an elevation spindle (8), an elevation wheel and other elements, is only alluded
to in the figure. Surface strains were measured on both the inside and outside of the specimen
with an array of electrical resistance foil gages (not shown in Figure 16.42).
The author pointed out that the end conditions of the cylindrical shells were not completely
defined. Due to initial deviations from circularity, the ends of the shells had to be push-fitted
into the circular annuli in the end plates, and therefore their rotational restraint was actually
indeterminate. Thus, the end conditions during clastic deformations before collapse had to
be considered somewhere between simple supports and clamping; but when collapse oc-
curred, the tested specimens revealed that all rotational freedom had disappeared (by jam-
ming), indicating that clamped boundary conditions were approached.
Datum readings were taken by the deflection probe at zero load. These datum readings
were not relied upon to provide the initial shape of the cylinder, since the probe did not
revolve exactly on the central axis of the test cylinder but were only used as references for
assessment of the change in deflections. As explained in Chapter 10, such datum readings
could today be employed for determination of the initial imperfection shape by referring
them to a best-fit cylinder.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

For classification of the test results, Montague first employed the thinness ratio A,

~r;;,
A=.~~{E (16.10)

proposed in 1934 by Windenburg and Trilling [9.35J, which includes the basic material
properties (yield stress and elastic modulus), and where as before Lis the unsupported length
of the shell, R its radius and t its wall thickness. Low values of A represent thick cylindrical
shells, for which the failure load can be estimated by simple calculations, based on circum-
ferential membrane yielding, and which collapse by "yield waisting." High values of A, on
the other hand, represent very thin cylindrical shells that fail by elastic instability. In the
intermediate range of A the failure mode was "yield lobing," the appearance of circumfer-
ential lobes after yielding of the shell material, as expected. Indeed, out of the 12 shells
tested (both with radial pressure and hydrostatic pressure), 7 (with 0.4 < A < 0.9) failed by
yield lobing, 4 (with 0.2 < A < 0.4) by yield waisting, and I (with a high A 1.5) by elastic
instability.
Another geometric parameter, the slenderness f = (L!V2Rt), also influenced the collapse
behavior, but for the Manchester University shells the grouping according to slenderness was
very similar to that by the thinness ratio A.
The formation of lobes in a typical sheill failing by yield lobing is shown in Figure 16.43.
The mechanism of lobe formation was studied on this specimen by resting the deflection
probe at the point of maximum deflection (a likely location of the first lobe) at the centerline,
while the pressure was increased. At 830 psi the deflection started to increase, first quite
slowly with a gentle decrease in pressure, and then with an audible bang the deflection grew
instantaneously to a large value, with a pressure drop to 400 psi. The process of lobe for-
mation was similar in all the shells that failed by yield lobing. It should be emphasized that
the collapse was initiated by yielding arrd not by instability. Only after yielding of the
material resulted in large deformations was an unstable condition reached. (For thinner shells,
one recalls, the process differs: first elastic instability appears, and then, as the lobes grow.
yielding may occur at the points of high stresses.)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1343

!obPs becoming

lobe ! abe
I
lobe
deeper

two thre-e 1our

'------------------
time

Figure 16.43 University of Manchester plastic buckling experiments on thick-walled cylindrical shells
subjected to external pressure-formation of lobes in specimen 1.2 (with Rlt = 28.07
and LIR = 2.24) that failed by yield lobing (from [16.101])

After formation of the first lobe, a new pressure buildup took place until at 760 psi a
similar sequence resulted in a second lobe at a different location on the circumference. As
seen in Figure 16.43, the sequence was repeated twice more to reach a four-lobe collapse
pattern.
In another similar specimen, no. 1.4, slightly thinner and shorter (with Rlt = 55.15 and
Ll R = 1.853), careful measurements were taken of the postcollapsc behavior (after formation
of the first lobe). A movement of the lobe (or lobes) already formed to a different circum-
ferential location was observed when a further lobe formed. On formation of lobe 2, lobe I
moved to the left, and with the development of lobe 3, both lobes 1 and 2 moved to the
right. The shell material thus appeared to be in a state of free plastic flow in the circumfer-
ential direction. Finally five lobes formed in the specimen.
In order to arrive at a more general evaluation of the collapse pressure and modes,
Montague also presented tables of some earlier experimental results of cylinders with low
thinness ratios A (in the range A = 0-1.5), which he found to be rather scarce in the literature
(see [10.3], [16.92], [16.96], [16.102] and [16.103]).
His main general conclusions can be summarized as follows:
1. For thin shells, i.e. A > 1.5, collapse initiates by elastic instability and elastic instability
analysis compares well with test results, even if initial imperfections arc not included,
provided they are small.
2. For thick shells, i.e. A < 0.5, collapse is by yield waisting and predictions based on
circumferential yield, with recognition of initial imperfections, compare reasonably well
with experiments.
3. In the intermediate range, 0.5 < A < 1.5, collapse is usually in the yield lobing mode.
Initial imperfections are of importance and can be decisive in causing collapse. Though
predictions from both circumferential yield calculations or elastic instability theory will
probably be too high for the collapse (pressure) in the intermediate range (yield lobing
mode), the number of lobes in this mode is estimated rather well by an elastic instability
analysis.
Before continuing, we should note that the definition of thin or thick shells is rather
arbitrary. For example, here Montague's shells have been referred to as "thick-walled" (since
they fail primarily by plastic buckling), whereas he called them "thin-walled" in the title of
his paper. Usually the definition will be guided by the (R/t) ratio, and when this is small
enough to result in plastic buckling or plastic collapse, the shell will be considered thick-
walled. However, in the design of pressure vessels or containment capsules, thick-walled
tubes are understood to be really thick, with (Rit) = 2-3 (see for example [16.104]).
The collapse behavior of such a thick-walled cylinder (really thick) differs somewhat from
that discussed above (see [16.104]).
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1344 Plastic Buckling ExperimE!nts

Under external pressure, the inside diameter of a thick-walled tube will experience the highest
stress and the tube wall will first yield at this location. Under the increasing pressure, the yielded
region of the tube wall will continue to expand toward the outside diameter until the entire wall
thickness of the tube is plastically deformed.
Once the cross section of a thick-walled tube is fully yielded, increasing pressure will, at first,
collapse the tube uniformly inward. However, small nonuniformities in the tube (initial ovalization
of the cross section, wall-thickness variations, etc ... .) will cause the tube to begin to ovalize or
flatten. In general, the thicker the tube wall (relative to its diameter) and the higher the rate of
strain hardening of the material, the greater will be the resistance of the tube to such ovalization
... the collapse of such (thick) tubes is not sensitive to small imperfections.
As the pressure is increased, the thick tube will further ovalize, till in the final collapse
complete flattening of the circular cross-section occurs, with opposite sides of the tube com-
ing into contact.
At the Battelle Pacific Northwest Laboratory, Richland, Washington, and at the Civil En-
gineering Laboratory of the U.S. Naval Construction Battalion Center, Port Hueneme, Cal-
ifornia, a series of hydrostatic collapse tests on thick-walled outer-containment capsules (with
Rlt = 2.5) were carried out in the early eighties. Correlation of experimental results with
analytical predictions was disappointing and emphasized the need for experimental verifi-
cation of similar designs. Some possible methods for better predictions were proposed, but
they will also require further experiments.
Before closing this subsection, it may be of interest to examine briefly a very recent
experimental study on plastic buckling of ring-stiffened cylindrical shells subjected to ex-
ternal pressure [ 16.1 05]. The experiments were carried out at the Department of Mechanical
and Manufacturing Engineering of the University of Portsmouth in the U.K. by a team that
had performed many external pressure tests on cylindrical and conical shells in the last
decades.
The simple test setup (Figure 16.44) was essentially similar to that employed in earlier
Portsmouth test programs (see for example Figures. 13.79 and 13.64 and [13.61], [13.104]
and [13.125]), or in much earlier U.S. Navy DTMB or Technion tests (see for example
Figures. 10.7, 13.46 and 13.59). The main difference was the heavier construction of the test
tank and the end plates (end cap and tank top) needed for the stiffer, thick-walled specimens
involved in the plastic buckling experiments. As before, however, the test shells were sub-
jected to external water pressure applied by a hand-driven hydraulic pump, and the top
opening facilitated attachment of strain gages on the inner surfaces of the specimens, as well
as visual inspections.
Six models were tested: three short and thick-walled cylindrical shells, with two fairly
heavy rings each, for inelastic shell instability (between the rings) experiments; and three
shells with nine relatively weak rings each, designed to fail by inelastic general instability.
All the specimens were machined from a solid billet of ENlA steel. First the internal surface
-o- RING
BLEED
SCREW
BOLTS

TANK
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 16.44 Portsmouth University experiments on plastic buckling of ring-stiffened circular cylinders
under uniform external pressure-test setup (from [ 16.105])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1345

of each shell was machined, and then it was placed on an internal mandrel for final precise
machining of the external surface, the customary procedure for integrally machined stiffened
shells. The radius-to-thickness ratios of the three specimens for shell instability tests (P 1, P2 ,
P,) were (Rit) = 40.5-42.9, and their effective length ratios (L/R) = 0.616-0.861; whereas
the three specimens for general instability(?.", P 5 , Pr,) had (Rit) = 45.2 and (L/R) = 3.81.
The initial out-of-circularity was measured for all specimens. For the first group it was 0.0025
mm (·-0.2 percent of shell thickness), while for the second group it was 0.0130 mm, 0.0069
mm and 0.0079 mm ( ~ 1.2, 0.61 and 0.70 percent of shell thickness, respectively). Surpris-
ingly, only the maximum initial out-of-circularity of the shells was reported, not their shape,
and apparently no assessment of a theoretical knock-down factor, based on an analysis of
the initial imperfection shape, as discussed in Chapter 10, was made.
On the other hand, extensive plots of the variation with pressure of circumferential strain
around the circumference, at midbay, were presented. It could be observed that the circum-
ferential imperfections, although very small, grew with increase in pressure until failure
occuned.
In spite of many years of efforts to develop reliable prediction methods for plastic buck-
ling, the design method proposed in 1995 is sti Jl a semi -empirical one. Also, that the clas-
sification factor, the thinness ratio A, proposed in the thirties by Windenburg and Trilling
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

and defined by Eq. (16.10), is still valuable today. For general instability of ring-stiffened
shells, a modified fonn of this thinness ratio, an equivalent shell thinness ratio A1, was
proposed by Ross in 1990 (see [13.61]):

I = 4 (LjD,f ra: (16.11)


A (t 11DJ f {£
where L" = length of ring-stiffened cylindrical shell between adjacent bulkheads
D 1 = diameter of the centroid of a typical ring-shell combination
t 1 = equivalent wall thickness of ring-shell combination

16.2.8 Plastic Buckling of Cylindrical Shells Subjected to Bending or


Torsion
As pointed out in Subsection 9. 7.2 of Chapter 9 and Subsection 13.4.2 of Chapter 13, the
bending instability of thicker shells, with say (R It) < 50, is often dominated by the Brazier-
type limit load instability, in which increasing ovalization causes a reduction of the bending
rigidity of the shell. The induced nonlinear ovalization involves large deformations and there-
fore soon introduces inelastic effects. Brazier introduced this collapse mechanism in 1927
for elastic deformations (see Subsection 2.l.l3b of Chapter 2 and [2. 70]), it was extended
to the plastic range by Ades in the late fifties [ 16.1 06] and has since been further developed
by Gellin [16.107], Reddy [16.108], Fabian [16.164], Bushnell [16.165] and Kyriakides and
his co-workers [2.71], [9.253]-[9.255].
Bifurcation buckling may, however, also occur in the elasto-plastic region. It is then char-
acterized by the formation of either wave-type multiple buckles or localization into a single
dominant buckle, depending on the wall thickness.
In experiments on long tubes with constant bending (see for example [ 16.109]), both
ovalization and wave buckles are observed. Ovalization appears from the onset of loading.
As the tube commences to yield, its rate of growth decreases but ovalization continues all
the way to failure (note that the state of stress in the tube is biaxial, due to the concunent
longitudinal and circumferential bending of the tube). "As loading continues, waves are
observed in the compression region. Eventually, one of these waves becomes a dominant
local buckle and others disappear. The actual failure appears to be associated with the local
buckle, which is similar to what might be expected in the bifurcation buckling mode. How-
ever, ovalization does play an important role" in reducing the curvature in the compression
region (thus increasing the effective Rlt ratio).
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1346 Plastic Buckling Experim•~nts

Sherman [ 16.1 09] grouped the observed inelastic failures and postbuckling behavior of
thick shells in three categories, as shown in the load-deflection curves (moment-rotation)
curves of Figure 16.45. As a classification parameter he employed a slenderness a,
a= (EiaJ · (t/2R) (16.12)
which is related to an earlier nondimensional local buckling parameter A used by Schilling
[16.110] for both axial compression and bending, where
A= (aJE)(Rit) = 112a (16.13)
Sherman's grouping according to the slenderness parameter a was:
"(a) For a> 22, a long plastic plateau occurs in the moment-rotation curve (Figure 16.45).
The pipe gradually ovalizes and eventually local wave buckles form, after which the
load slowly decays. One of the waves may become a single dominant local buckle.
(b) For 10 < a < 22, the single buckle gradually forms and the load decays with little
or no plastic plateau region.
(c) For thin pipes with a< 10, multiple buckles form suddenly with very little ovalization
and the bending moment drops quickly to a more stable level."
It was observed that the local buckles were primarily inward indentations.
It may be of interest to compare Sherman's groups with a limit on A proposed two decades
earlier by Shilling [ 16.11 0] in order to ensure the achievement of the full plastic moment in
a bent tube. Schilling proposed A < 0.06, which means a > 8.33, i.e. practically in group
(b) or (a), a very close correlation. Note that if Schilling's limit had been A < 0.05, the two
classifications would have been identical.
Different pure bending experiments in the plastic range, with shells made of different
materials, produced essentially similar collapse modes: significant ovalization with concur-
rent small ripples or buckles. These were observed, for example, by Sherman on large steel
pipes with (RI t) = 9.2-55.4 (see [16.109] and [ 16.111 ]); by Reddy on small (25 mm nominal
diameter) steel and aluminum alloy tubes with (Rit) = 17.4-38.9 (see [16.108]); and by
Kyriakides and his co-workers on similar (25-38-mm diameter) aluminum alloy and stainless
steel tubes with (Rit) = 10.6-30.3 (see [2.71], [9.253]-[9.255]). In many cases the small
ripples developed into a local buckle, but nearly always they initiated the collapse. Various
explanations have been attempted for the appearance of these small ripples, which usually
ride on the dominant ovalization of the tube.
For example, since in Reddy's tests [16.108] the ripples were already observed at a bend-
ing moment half the maximum value, he suspected the influence of imperfections to be their
origin, and he also suggested that the ripples indicated the presence of a deformation mode

THICK SHELLS (a> 22):


M/M p GRADUAL OV~~~'ZATION &
ST~ARD~
1.0 --- ----- - - - - -
MODERATELY THICK
(10<a< 221:
GRADUAL SINGLE
1 THIN SHELLS LOCAL BUCKLE

--
\ Ia< 101:
\ SUDDEN MUL T!PLE
~BUCKLES

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 16.45 Types of inelastic behavior in thick cylindrical shells-moment rotation curves (from
[16.109] with some modifications): M is the applied bending moment, Ml' the fully plastic
Copyright Wiley
moment,
Provided by IHS Markit under license with WILEY 8 the bending rotation andLicensee=McDermott
B, the rotation Inc - at which yielding
India/8215328006, User=G, occurs
Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1347

related to the axisymmetric bucking modes that had been observed in plastic buckling of
shells under axial compression.
Kyriakides and his co-workers, who carried out extensive studies of inelastic bending
instabilities of shells, divided their shells into three broad categories according to their ap-
proximate (R/t) values (see [2.71]):
1. Relatively thinner shells (Rit > 20) for which the usual mode of instability was short
wavelength rippling, which appeared already at bending strains significantly below those
corresponding to the limit load resulting from uniform ovalization, which eventually
caused collapse. These ripples appeared in small pockets, randomly distributed along
the length, and grew in a localized manner. The University of Texas researchers proposed
that the rippling qualitatively corresponded to the axisymmetric plastic buckling mode
of axially compressed cylindrical shells, as had been postulated by Reddy.
2. Moderately thick shells (13 < R/ t < 20) for which the prevalent mode of instability
was, again, short wavelength rippling. In these shells, however, a well-defined limit
moment was recorded. "In addition some non-uniform (localized) growth of ovalization
was observed just prior to failure .... Pockets of short wavelength ripples were usually
observed in the trough of the localized regions. Following the limit load, the shells
buckled catastrophically by the appearance of one sharp, local kink initiated from the
ripples."
The buckling behavior of the moderately thick shells was considered by Kyriakides and
his students from two aspects. One was based on the fact that "the curvature at rippling was
closer to the limit load induced by uniform ovalization," and therefore an influence of long
wavelength imperfections became possible resulting in a lower collapse moment. The other
was that "the localized ovalization in tum triggered the short wavelength ripples, which
made the shell axially more compliant," reducing the overall bending rigidity and therefore
the limit load. Both explanations indicated that the rippling was the origin of instability and
that it reduced the stiffness of the shell and hence its limit moment.
3. Thick shells (Rit < 13), whose collapse was governed by limit load instability caused
by uniform ovalization. Close to the limit moment the influence of long-wavelength
imperfection caused the ovalization to grow non-uniformly. After the limit load, the
ovalization localized significantly, till failure occurred in the region of localization. Here
the localization was a direct consequence of the limit load instability and affected the
behavior primarily in the postbuckling regions.
Sherman's thickness grouping, discussed earlier in this subsection, which was obtained
from a large number of inelastic bending tests on large steel pipes, is very similar to that
proposed by Kyriakides and Ju, as can be seen in Table 16.1. This similarity verifies the
earlier observation of similar collapse behavior, depending primarily on the wall thickness
ratio (R!t) of the shell.
In summary, one may reiterate that only for really thick shells, with (R It) < 13, is there
complete dominance of the Brazier-type ovalization and limit load instability. For thinner

Table 16.1 Thickness grouping of circular cylindrical shells subjected to bending


Kyriakides and Ju, 1992 Sherman, 1986
[2.71] [16.109]
Thickness grouping Group (Rit) Group (Rit) Slenderness
a*
Thin shells (1) (Rit) > 20 (c) (Rit) > 24 a< 10
Moderately thick shells (2) 13 < (Rit) < 20 (b) 11<(R/t)<24 10 <a< 22
Thick shells (3) (Rit) < 13 (a) (Rit) < 11 a> 22
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright*From
Wiley Eq. (16.12), a= (EiaJ(t/2R)
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1348 Plastic Buckling Experiments

shells. both moderately thick and thi n ones. ~hort-wavelength rippling initiate~ plastic buck-
ling in cylindrical ~hells subjected to pure bending. The ripples are a shell-type in<tability.
very similar to that occurring in plastic buckling under axial loading (~ee for example
[ 16.89)), except that here the ax ial stress is not uniform and the ripples interact with the
inherent oval ization.
As mentioned before. plastic buckling of ~hells became .a main research topic in the late
thinies. One of the earliest series of bending te~t~. that specifically aimed a1 plaslic buckling
of cylindrical shell~. was carried out by Osgood at the U.S. Bureau of Standards in the late
thinies 11 6.79], as pan of the NACA-sponsored research program, referred to in Subsection
16.2.6.
The bending teM setup used by Osgood (shown in Figure 16.46) is \till of interest today.
over 60 years later. and warrants a detailed discussion. The test specimen A in the ligurc
( 1-2-in. diameter tubes) was supponed at i1s ends ami was loaded symmetrically a1 the th ird
points unt il the bendi ng moment in the middle portion reached a maxi mum. The supports
were designed to ensure that fail ure occurred in a part of the specimen that was unaiTccted
by local constraints or concentrated loads. To promote this condition. the loacls were applied
through stiff clamps B. which lined the tube ~nugly at the third points. One of these clamps
was always tightened to a sl iding li t only, to prevent torsiona l stresses to be introduced by
any possible rotation between the clamps about the axis of the specimen. The rigid clamps
held the tube circular at the loaded sections :md thus prevented their flallening. The middle
of the specimen. on the other hand. subjected to a constant bending moment. of the same
magnitude as that at the loaded sections, "was free to deform at will and took a characterist ic
ftauened shape at fai lure. All failures occurred at or ncar the mi dd le. Prel iminary tests showed
that a distance of live diameters between load ing points was Mtflicient to permit the middle
section to assume iLs natural shape." All but three specimen\ were. however. tested with a
larger distance between the loading poinLs.
''The loads were applied through knife edge; on the clamps by means of hangers C. which
extended down to an equalizer. The equalizer (not shown in the figure) bore on knife edges
on the lower ends of the hangers and was itself loaded through a knife edge at the center
by the movable head" of a testing machine.
The specimens were usually 19 diameters long (L/ R ;;;; 38). "the ends were plugged, and
they were supported on knife edges D. with the thi nnest part of the tube up (in compres;ion).
The supporting kni fe edges D were spaced accurately by means of spacer bars (not shown)
and rested on hard steel plates £which bore on hard steel balls. The balls were free to roll
on other hard steel plates F. thus practically eliminating axial stresses in the specimen." The
whole test assembly was carried on two structural steel channels and placed at the top of
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 16.46 Osgood's 1938 bending strength tc'" on round au-craft tubing-tc't setup for determi-
nation or bending >trcngth (from (16.791>
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1349

the testing machine. To prevent the assembly from rolling off. stops G were provided. The
setup was simple, yet quite accurate.
Twenty chromium-molybdenum steel tubes (with Rlt = 7.8- 50.0), and 20 d uralumin tubes
(with Rlt = 7.9-49.2), were tested. Some typical fai led specimens are shown in Fig ure
16.47.
The widespread development of the offsho re oil industry in the seventies motivated ex-
tensive study of plastic buckli ng of pipes subjected to bending, and many experimental
investigatio ns of their bending strength were sponsored (sec for example [ 16.1 08] , [ 16. 111 ],
[1 6. 113]- 116.116]).
The tests carried o ut by Reddy at University College. London f 16.108] in the late seventies
are a good example of the experimental investigatio ns of cylindrical shells subject to bending
in that decade. In his widely quoted paper, Reddy fi rst emphasized that, as proposed by A.C.
Palmer. for the study of bending collapse. the curvature was a much more satisfacto ry and
convenient measure of buckling processes than the bending moment; or, alternat ively, the
extreme fiber compressive strain was a bcucr criterion than the corresponding stress. Such a
presemation of experimental values of critical extreme fiber compressive strain (or simply
critical outer strains) versus (R/1) from Reddy's tests and from two o ther contemporary
investigatio ns [ 16.117] and [ 16. 118] is shown in Figure 16.48. TI1e correspondi ng results
from axial compression tests [ 16.38]. also presented, fit qui te well into the scatter band. and
the empirical lines have the same s lope as the classical clastic formu la.
The test setup is shown in Fig ure 16.49. It can be seen that the test specimen (4) is

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
supported at its extreme ends by rectang ular, vertical strips of spring steel (5) (see Figure
16.49b), whose bottom edges are bolted at (7) to the extreme ends of the end blocks of the
specimen ( 10) and whose top edges are attached at (6) to the o uter frame (2). The inner (I)
and outer frames (2) are fabricated from channel sections, and the outer frame (2) rests on
a solid base (3). The frame arrangement is symmetrical with respect to the vertical plane
passing through the longitudinal axis of the tube and that passing through its center cross
section. Additional pairs of spri ng steel strips (8) arc bolted to the specimen end connectio ns
at the poi nts where the collars (9) enter the end blocks ( 10). T he far ends of these pairs of
stri ps (8) arc anachcd to the inner frame (I) at (II). A pure bending mo ment (wi thout shear)
is applied to the specimen by four-poim loading, which is obtai ned by a downward movement
of the inner frame (I). Due to their flexibility. the spring steel strips offer only negligible
rotational restraint at the ends of the specimen. These small restrai ning moments were,
however, accounted for in the calculatio ns of the applied bendi ng moments. and some of
these calcu lations were verified experimentally.
T he load is applied to the inner frame ( I), by a screw jack (beyond the top of the figure)
through a load cell (13). via a centering ball (14). The screw jack is rigidly anached to the
top of a very stiff framework ( 15). A system of cou nterweights (16), ( 17). (18) and (19)
ensures that no loads act on the specimen due to the dead weight of the inner frame.

- ....

Figure 16.47 Osgood's 1938 bending strength tests on round aircraft tubing-typical fai led specimens
(from I I 6. 79])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1350 Plastic Buckling Experiments

"~[
'~
0.04 'f REDOY-steel !(REF. 16.108)
• REDDY -aluminium
0.03 'l Wilhoit 8 Merwin (REF. 16.118)

:~
o Sottermon (REF. 16.38)
E. up
t::. BouwKamp (REF. 16.117)
o.oz

. ...
Class!col elastic
t:~ ~ 0 6t/ R
.., 'l

0.01 •
0

.• ""95
1:::1. 'Q

• 'l

plastic; low hordeflinq


(perfect shell)

t:,~ .4t/R
/.
plastic; hiQh hardening
(perfect shell)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

O.OO!L__ _ __L__ _l__ _t__---l---'L----"-:-:'::-=-


10 20 30 40 50 60 80 100
( R It l
Figure 16.48 Experimental critical extreme fiber strains versus (R It) for plastic buckling of small, thick
tubes subjected to bending (Reddy and two other investigators) or to axial compression
(Batterman) (from [16.108] with some additions from [16.107])

The downward displacement was measured by the dial gage (20), while gage (21) mea-
sured the displacement of a connecting bar between the end blocks of the specimen. Metal-
foil strain gages were attached to the specimens at the mid- and quarter points, at the two
extremes of the vertical diameter at those points.
The test procedure involved the application of load increments and measurement of the
resulting strains, changes in diameter and dial gage readings.
In the tests buckling always occurred suddenly and with a bang, probably because of the
flexibility of the loading rig and the flexibility of the specimen itself, it being relatively long,
with (LI R) ~ 24. Ovality or changes in vertical and horizontal diameters were measured by
means of a micrometer at positions adjacent to the strain gages. One strain gage was mon-
itored continuously with load to obtain readings as close as possible to the collapse point.
Nineteen tubes were tested in this program.
Visual detection of the appearance of wave-like ripples, before collapse, was possible
because of the highly polished surface of the tubes. Some typical longitudinal profiles of
ripples, measured along the compression side of the tubes, are presented in Figure 16.50.
The extensive investigations of plastic buckling of relatively thick-walled cylindrical shells
subject to bending, primarily motivated by the oil industry, continued into the eighties and
nineties (see for example [2.71], [2.72], [9.253]-[9.255], [16.119] and [16.120]).
The University of Texas small laboratory-type bending test facility, discussed in detail in
Subsection 9.7.2 of Chapter 9 (see Figures 9.76-9.81 and [2.71]), is a very good example
of a modern bending test setup developed primarily for plastic buckling studies. It may
therefore be useful to reiterate that discussion here, also in connection with the ripples which
were considered at length in the present section. It may be recalled that there are now actually
two bending test setups at the University of Texas Engineering Mechanics Research Labo-
ratory, the second and more recent one (see Figure 16.69) having been developed for com-
Copyright Wiley
bined
Provided by IHS loading.
Markit under license withThe
WILEYmost up-to-date detailed information onIncthis
Licensee=McDermott remotelyUser=G,
- India/8215328006, controlled
Boopathi bending-
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1351

(a}

(b)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 16.49 Reddy's experiments on plastic buckling due to pure bending. of moderately thick small
stainless ste.e l and aluminum alloy tubes (with R/1- 17-39)-test setup ( from I 16. 108]):
(a} fron t view. (b) side view

pressure test device can be found in [ 16 .11 2 ]. Similar devices have recently been built by
Kyriakides' former students, Corona at the University of Notre Dame [16.1 66] and Yc h at
the National Tsing Hua Uni versity o f Taiwan .
Larger-scale experime nts o n the bending strength o f relative ly thic k pipes have also been
continued. Many o f these tests dealt not only with plastic buckling due to bend ing but also
with that due to combined loadi ng. in particular bendi ng and external or internal pressure,
whic h is considered later in Section 16.3.
Among the researc h centers with extensive bendi ng strength testing on cyl indrical shells
in recent years. o ne may mention the Departme nt of Civi l Engineering of the Un iversity o f
Wisconsi n-Milwaukee (see fo r example [16. 109]. [ 16. I I I] , [ 16. I I 9]-[16. I 2 1]); the Institute
TNO for Building Materials and Build ing Struc tures, Delft, The Netherlands (see fo r example
[16.1 22 ]. f 16.123] and [ 16.129]); and the Departme nt o r Civil Engineering at the University
of Albe rta, Canada (sec for example [1 3.11 4], [ 16 .1 24] and [ 16.125]).
The compre he nsive University of Wisconsin-M ilwaukee test p rogram on elastic flexural
and combined loading buckling of cylindrical shells, whi ch commenced in the mid-seventies
Copyright Wiley
and continued into the eighties and beyond, included
Provided by IHS Markit under license with WILEY
over 100 tests or fairly large circular
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1352 Plastic Buckling Experiments

UftB

Figure 16.50 Reddy's experiments on plastic buckling, due to pure bending, of moderately thick small
stainless steel and aluminum alloy tubes (with Rlt = 17-39)-typicallongitudinal pro-
files of ripples (from [ 16.108])

steel tubes. The tests were divided into two groups according to their type of support and
loading: one of simply supported beams with third-point loads, and one of cantilevers with
end loads, as well as fixed-end beams (with third-point loads) in the early phases of the

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
program. The behavior at the resulting constant moment regions was compared with that at
regions with moment gradients (shears) and restraint against ovalization at the ends. The
overall objective of the more recent phases of the program (see [16.109]) was determination
of the complete moment-rotation characteristics of pipes with an appropriate range of (Rit)
or slenderness a = (EI a)(t/2R) "that would fail by inelastic local buckling, either before
or after a full plastic moment had been achieved."
Sherman also compared the University of Wisconsin-Milwaukee (UWM) results with those
of other similar experimental programs on shells with constant moment regions and empha-
sized the different manufacturing methods employed for the specimens. Some of them were
hot-formed seamless pipes (HFS), some electric resistance welded tubes (ERW) and some
fabricated pipes. He pointed out that since the tubes were made by different methods, "the
tests were conducted on members with different degrees of imperfection, residual stresses
and stress-strain characteristics," and that these differences significantly influenced the bend-
ing strength. For plastic buckling one has to consider not only the correlation between the
manufacturing method and the resulting imperfections, discussed in Chapter 10, but also the
influence of the fabrication processes on the stress-strain characteristics of the material.
A typical setup for a recent UWM experimental program of flexural tests on large diameter
pipes is shown in Figure 16.51.
The specimens were tested as simply supported beams with two symmetric concentrated
loads to produce a region of constant moment, as shown in Figure 16.51 a. Total specimen
lengths were determined by the 100 kip capacity of the loading rig and ranged from 30-40
ft (9.15-12.2 m). However, the important length was the constant moment region, where all
the failures eventually took place. Since the loads were applied through plates that held the
shell section round (see Figure 16.15c), the constant moment regions LP were kept at least
four diameters in length to permit unrestrained ovalization.
Based on the experience of previous testing of large steel tubular members, it was rec-
ognized that the loading system must accommodate deflections of two diameters or more in
order to cover the full range of plastic postbuckling behavior. With deflections of this mag-
nitude, chord shortening is substantial and may cause secondary axial tension and misalign-
Copyright Wiley
ment of the load actuators.
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1353

LS

'p ,.

(c)
~-~- ..
lO!
-
(b)

, ...~
f m••-· ~=~•~•-
~\!3 I"'~
· I .f

~~ -- ~
I I

S<>IH c,..,, Rods


(d) 1
Lp
1

Figure 16.51 University of Wisconsin-Milwaukee plastic buckling test on tubes subjected to bending-
typical test setup (courtesy of Professor D.R. $hetman): (a) schematic of test setup. (b)
overall view, (c) loading system. (d) instrumentation

To overcome such problems. "gravity load simu lators." as shown in Figure 16.5lc, were
used as key clements in the loading system. Similar dev ices were developed at Lehigh
University in the sixties for exerting vertical loads on frames that were allowed to sway
lateral ly (see Figure 6.57 and [6.12 1] of Chapter 6). Basically. such a gravity load simulator
is a mechanism of tension bars and pivots that keeps the loading rod vertical even if the
poi nt of application on the specimen moves horizontally. In the UWM tests two such sim-
ulators were used to apply the loads at the ends of the pipes. They were designed for 6 in.
of lateral motion. and much of this capacity was used in some of the tests.
In order to keep the test section at an easily observed elevation, the specimens were
attached to the test Hoor at the two center points and were loaded by displaci ng the ends
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

upward (sec Figure 16.5Ja and b).


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1354 Plastic Buckling Experim1~nts

Another noteworthy feature of the loading system was that the concentrated loads and
reactions were transmitted through split plate assemblies (see Figure 16.5lc). These 1.5-in.
( ~ 38-mm) thick plates encircled the shells like a stitlener and were hinged to an external
frame to which the load rods were attached. Consequently, the plates could assume the shape
of the specimen while the load remained vertical. The split plates were not welded to the
shell but were shimmed for a tight initial fit. Although the shims tended to loosen under
high loads, the specimen was held round at the load points and no local buckling ever
occurred near the concentrated loads.
The instrumentation (see Figure 16.51d) consisted of load measurements by strain gages
on the four load rods; strain gages for strain measurements on top and bottom of the shells
at two points in the constant moment region, as well as some additional rosette gages around
the circumference for determination of variation of strain through the depth of the specimens;
and an array of various deflection measllfement devices, some mechanical and some elec-
tronic, with some redundancies, as indicated in the figure. The redundancies were considered
necessary to provide cross-checks against malfunctions or misreadings, in particular in the
region of large plastic distortions.
The University of Alberta bending strength tests on large longitudinally stiffened steel
cylinders were discussed in Subsection 13.4.2 of Chapter 13. The test rig used there (see
Figures 13.54 and 13.55 and [13.113]) was practically the same as that employed for buck-
ling experiments on unstiffened steel shells subjected to bending [13.114]. Details of the test
setup, which is also well adapted to plastic buckling experiments, can be found in Subsection
13.4.2. These unstiffened shells were fairly thin-walled (Rit = 149-222) and therefore their
buckling was practically elastic, local buckling with diamond-shaped buckles. The local
buckling commenced with a single depression centered on the extreme compression fiber,
very close (150 mm) to the circumferential weld, that joined the test section of the shell
with the thicker-walled end sections. Buckling spread about the top fiber, and in its final
form three distinct lobes appeared, covering less than half the circumference. At the sudden
failure most of the deformation was already inelastic, though the flattening of the cross-
section (the ovaling) did not exceed 0.6 percent of the initial radius and therefore had only
a negligible effect on the local radius of curvature or moment of inertia.
In their earlier studies [ 16.124 ], Kulak and his co-workers had developed a buckling
strength parameter y
y = (E/ a) 0 5 (t I Rf' (16.14)
which differed from Sherman's slenderness a, of Eq. (16.12), but is related to it
y '= y/2; (f/R) (16.15)
In a recent review paper [16.125], Kulak compared the plots of the results of 42 bending
strength tests of tubular members [16.111!], [16.113], [16.116], [16.120] and [16.124] versus
a and versus y, and found some slight reduction in scatter when y was used (see Figure
16.52). Note that in both presentations the stocky cylindrical shells appear in the right-hand
regions of the plots while the slender ones occupy their left-hand regions.
The University of Alberta team has al~m carried out experimental investigations on large-
diameter fabricated steel cylinders subjected to transverse loads (see [16.126]) as well as on
the local buckling of full-sized line pipes due to inelastic deformations caused by imposed
geotechnical settlements [16.127].
Their transverse load tests were essentially large-sized complements to earlier small-scale
tests on short cantilever steel shells subjected to transverse shear loads, carried out at Liv-
erpool University in the mid-eighties and discussed in Section 9.12 of Chapter 9 (see Figure
9.151 and [9.408]), or to a similar series of model tests carried out at the University of
Tokyo in the same period (see [9.409]). The small-scale 150-mm diameter specimens of the
Liverpool experiments had (R/t) ratios of 125-190 and their (L/R) ratios were 0.83-1.37.
The Tokyo specimens were also in the same range of geometries.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1355

1.6

1.4

1.2
Mu
1.0
My
0.8

0.6

0.4

0.2

0
0 10 20 30
(a) a= (E IO'y)I(D it)

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
1.8

1.4 •
1.2

~ 1.0

My 0.8

0.6

0.4

0.2

0.2 0.4 0.6 0.8 1.0


(b) 0
r=(~ rl (t(l
y

Figure 16.52 The inelastic bending moment capacity of tubular members from five different test pro-
grams, plotted versus two different geometry/material parameters (from [16.1251): (a)
plotted versus Sherman's slenderness a, (b) plotted versus Kulak's buckling strength
parameter 'I

The University of Alberta program, on the other hand, aimed at realistic, full-scale shells.
The two test specimens had a nominal diameter of 1270 mm and an (Rit) ratio of 185. One
shell, Sl, was tested as a beam with fixed end supports and had (L/R) = 1.9, while the
other, S2, was tested as a cantilever with (L/ R) = 1.2. In fact, S2 was made of one of the
spans of S 1 that had recovered completely upon the removal of the load. The specimens
were fabricated by a local steel fabricator using normal shop procedures in order to simulate
realistic shells. The initial geometric imperfections resulting from fabrication, welding and
preparation of the specimen for testing were carefully measured (at 600 locations) and re-
corded.
The test setup and the buckled shape of specimen Sl are shown in Figure 16.53 and those
of specimen S2 in Figure 16.54. The transverse load was transferred to shell S 1 through two
19-mm thick diaphragms designed to transmit a 2000 kN vertical concentrated load into the
thin wall of the test shell by means of shear flow through the weld. The ends of the test
cylinder were also welded to 19-mm thick diaphragms, which in turn were bolted to the
supporting columns. Note that each set of end-supporting columns consisted of five 1-section
columns, though only three of them were fixed to the floor, and two hung freely in order to
provide out of plane stiffness to the end diaphragms.
The residual stresses that resulted from forcing the specimen into circular shape and by
welding were measured by 120 strain gages that had been installed prior to attaching the
shell to the test frame and welding the diaphragm plates. After the initial strain readings had
been taken, after the welding of the shell to the end- and load-diaphragm plates was com-
pleted, and after the process of making the cross-section of the shell as circular as possible,
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1356 Plastic B uckling Experiments

(a)

r- .--
r- ?
• IJ/
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

e
~ ~

(b)

2000

ISOO

~
~
J 6
(c) Ver1i~l Di.splac~ mtnt (mm )

Figure 16.53 University of Alberta tests on large fabricated steel cylinders subjected to transverse shear
loads-test setup and buckled shape for specimen Sl (from 116.126]): (a) test setup. (b)
buckled shape, (c) load-displacement curve

by insertio n of light timber spacers, was fi nished, a set of strain readi ngs was taken. T he
residual stresses measured were found to be mostly small (generally less than 0.3 CT,.) and
rather erratic. Extensive instrumentation was installed at the fixed ends to detect any move-
ment or rotatio n of the ends.
The shell was loaded gradually. in increments of 100 kN (about 5 percent of estimated
buckling load), till nonlinear displacements began to appear (at about three-quarters of the
buckling load), and thence the loading increments were halved. At a load level of 1700 kN
the shell buckled with a lo ud noise. wi th two diagonal waves, shown as nos. I and 2 in
Figure 16.53b. At buckling the load dropped to about 88 percent of the ultimate load. As
loading continued, the load increased again till at 92 percent or the fi rst buckli ng load a
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1357

llumm,Y Oiaphrasm

1273

(Nul~: All dimtttsivns art in mm)


(a)

(b)

(C)

Figure 16.54 University of Albcna ICSIS on large fabricated cylinders subjected to transverse shear
loads-test setup as a cantilever and buckled shape for specimen S2 (from [ 16.126]):
(a) test setup as a cantilever, (b) buckled shape. (c) buckled shape of the finite-elemem
model (magnilled lO times)

second buckli ng occurred with similar waves on the other side of the same span of the
cy linder. The load dropped to 76 percent of the original ultimate load and then recovered to
81 percent. Thence the load-displacemem curve stayed almost horizontal. When the vertical
displacement at midspan approached 5 mm ( - 1.45 t), a single depression buckle appeared,
shown as no. 3 in Figure 16.53b, together with three small buckles at the extreme compres-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

sion fiber of the central test section, shown as no. 4 in the figure. The postbuckling behavior
of speci men S,. discussed above. is summarized in the load -d isplacement curve shown in
Figure 16.53c, where the d isplacement calculated wi th a fi nite-element analysis is also pre-
sented. This analysis (see [ 16.128]) predicted the ultimate load withi n I percent of the
measured one. though the number of buckl ing waves d iffered (two buckles observed, while
three were predicted).
Specimen S 2 was tested as a canti lever beam (see Figure 16.54a) in a similar test setup
and with simi lar instrumentation. One may note the rol lers placed above the loading dia-
phragm tO allow for the horizontal displacement due to edge rotation, as well as the dummy
diaplu·agm and bracing clements added for additional stiffness and stabili ty. The observed
buckled shape and the predicted o ne arc shown in Figure 16.54b and c.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1358 Plastic Buckling Experimen ts

The other recent University of Alberta test program on local buckli ng of line pipes due
to geotechnica l seulcments actually sllldied full-scale steel pipes, 0.386 m and 0.508 m in
diameter (with Rlr ~ 62-64 and LIR 3.33-5.22), subject to combi ned bending and axial
compression. Since in previous full-scale tests fai lure had always been observed to be highly
localized. an eccentric column configuration with internal pressure was employed here in-
stead of the usual four-point bending. Details of the tests and their results can be found in
[ 16.127].
The experimental studies perfo1111ed at TNO for Bui lding Materials and Building Struc-
tures. Delft (sec [ 16. 123)), focused on the combined loading of ex ternal pressu re, bending
and axial force, and therefore relate also to the discussion in Section 16.3.
The TNO Delft test facility for combined loading is shown in Fig ure I 6.55. It consisted
of a 350-mm diameter pressure vesocl. from which the test specimens protruded at each end.
The loading equipment that applied bending and axial forces was therefore si tuated outside
the pressure vessel. a configuration that permitted easy access to this equipment. The typical

(a)

(M )

( P)

outhnf' of IN lt st toc•hly
(b)

Figure 16..55 TNO for Building Materials nnd Building Structures, Delrt, C<)mbined external prcswre.
bending and axial fo rce tests on moderately thick-walled pipcs - schcmmic diagram of
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
tc~t facility (from [ 16. 123[, counesy of Mr. P. de Wimer)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1359

in,trumemation of each specimen included six specially designed calipers. at different lon-
gitudi nal locations, for measurement of change of diameter during loading, eight strain gages
for local determination of changes of curvature. two denection transducers and four specially
designed displacement transducers for measurement of the cha nge of the average curvature
over a certain pipe length (about 300 mm). The concurrent measurement of local and average
curvature was found to help in the interpretation of the test results.
The pipes tested were moderately thick-walled. with (Rit) = 12-16, and made by difl'ercnt
method~ of fabrication. welded pipes. seamless commercial pipes and specially machined
test ~pecimens. The TNO Delft test results. as well a~ related results of a 1977 Shell De-
velopment Company test program, were companed to predictions by a TNO derived theory
that u~ed a simple rigid-plastic model. Satisfactory agreement was found between theory and
tcM (sec 116.123]). It was. however, shown that the loading sequence signi ficamly innuenced
the deformat ion capacity of the pipes.
Since the test setup did not completely simulate the <tctual pipe-lay ing situation, a new
larger. 1200-mm diameter. test faci lity was planned at TNO, but it was not bu ilt. probably
due to shortage of funding.
Tests on the torsional strength of thick-walled tubing were carried out in the twenties and
thinies (see for example ( 16.1301 and ( 16.131]) concurrently with the early experiments on
ela.\tic buckling of thin-walled cylindrical sheUs. discus~cd in Subsection 9.7.3 of Chapter
9. As mentioned there, in the late thinics an extensive series of NACA-sponsorcd torsion
tc't~ was carried out at the U.S. National Bureau of Standards (see [9.261]). The main
purpose of this systematic test program. which inc luded 63 small chromium-molybdenum
steel tubes, of (Rit) a 6-37 and (LI R) "' 15-160. and 102 aluminum-alloy ones. of
(R/1) • 4- 50 and (L/ R) 20- 120. was to study the relationship between torsiona l strength
and the (Rit) ratio. as well as the fai lure modes. Three different types of failure were con-
sidered to be of particular signific:mce in the design of circu lar tubes subject to torsion:
1. Two-lobe buckling of the wall of the tube
2. Helical defom1ation of the axis of the tube
3. Plastic yielding of the material
The first two types of failure arc caused by elastic instability, but in most of the tubes tested
considerable yielding apparcnc ly preceded the instabili ty. which should therefore be consid-
ered an clasto-plastic instability.
Figure 16.56 shows four typical 2017-T aluminum-alloy tubes after completion of the
tor>ion test. The top specimen B, (with Rlt = 39.3 and failure stress T,. < T,) failed with a
loud snap by two-lobe buckling. clastic buckling followed by plastic deformation. The second
tube J was thicker (with Rlt = 14.0 and T. > T,): after yielding. the torque increased ~lowly
with increasing twist until failure occurred by gradual two-lobe buckling. typical plastic
buckling. The third, longer tubeS, (with Rlt = 9.8 and T,. > T.) failed, after yieldi ng, by
helical deformation of the axis of the tube (the second type of failure).The last and th ickest

J
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

F igure 16.56 U.S. National Bureau of Standards torsion tc<ts of tubes in the thinie;- four typical
20 17-T alu minum-alloy tubes aOcr completion of torsion tests (from !9.26 11): 11, failed
by elastic two-lobe buckling, J, yielded und then J'uilure occurred by gmdual two-lobe
buckling. S, yielded und then failed by helical deformation of the axis of the tube. T,
failed atypically by a sudden fracture
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1360 Plastic Buckling Experiments

specimen in lhc figure T, (wi th Rlt = 4.6 and 1', > 1') fai led by a sudden fraclure (an
atypical mode of failure 1ha1 occurred in on ly 3 of the 102 alumi num-alloy lubes lested).
Based on their careful parametric ex.perimcntal study. the Bureau of Standards researchers
also developed an empirical formula and design chart.~. which were used for decades. As
mentioned in Subsection 16.2.6. parallel NACA-sponsorcd SIUdies were carried out in the
la1e thirties at the Aluminum Company of America, ALCOA (see [ 16.8 1] and 116.82]).
Very few experimental investigations of plastic torsional buckling of ci rcular cylindrical
shells have been reported in the literature in recent decades. One ex.ample is the tests carried
out at the Dcpanment of Mechanical Engineering. Middle East Technical University. Gazi·
antep. Turkey. in the mid-eighties [16.1321. Three sets of copper shells (nine specimens with
Rlt = 13.4-26.0) and one set of brass shel ls (four specimens with Rl t 21.8-22.8) were
tested in a 4000 Nm capaci1y Trebel Torsion Tesler. which also provided applied torque
versus angle of twist plots. The mechanical properties of the specimen materials were ob-
tained from tension tests. The experimental critical shear stresses were compared with those
predicted by an approx.imatc theory. and reasonable agreement (within 4 percent) was ob·
tained.

16.2.9 Plastic B uckling Tests on Conical, Spherical and Tor/spherical


Shells
AI the outset of this subsection we should reiterate that for these shells, as for the cylindrical
ones, many of the test seiUpS and techniques employed for elastic buckling were actually
also used for elasl.o-plastic buckl ing. The reader may therefore wish to consult concurrently
the relev;tnl sections in Chapter 9 and 13.
For conical shells, one of 1he earlier studies of plastic buckling was that carried out by
Rantsey at the University of British Columbia. Vancouver. Canada, in the mid-seventies
[16. 133]. In the experimental pan of the study, small 6061-T6 aluminum and type 416
stainless >teel steep. truncated conical shells were loaded in ax.ial compression. The shells
were 75 mm and 70 nun high. respectively, and 1heir average radius-to-wa ll 1hickness ratios
were about 12 for the lO alumi num alloy shells tested and about 10 for the 7 stainless stee l
shells tested: both groups could therefore just be classified as thick shells.
The specimens were cut on a lathe from solid circular bars. The edges at the small end
were rounded to permit free rotation and thereby approach a boundary condilion of zero
bending moment. The large ends 1erminated in heavy flanges (see Figure 16.57). The 6061-
T6 aluminum cones were not heat-treated in any way after machining, si nce in this material

0
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 16.57 University of Brilish Columbia plnslic buckling ICsts on steep. truncated conical shells
;ubjected to axial compression- cross-sections of buckled specimens, a stainless steel
one on the left nnd an aluminum one on the right of the figure. the scale being in em
(from [16.33))
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1361

heat treatment cannot remove residual stresses or material anisotropy. On the other hand, the
type 416 stainless steel cones were extensively heat-treated after machining, this material
having been selected because of its good response to heat treatment. Stress-strain curves for
the two materials are shown in Fig. 16.58. The aluminum demonstrates a sharply defined
knee and a constant tangent modulus past the yield point, while for the heat-treated stainless
steel the curve bends over much more gradually and has a continuously decreasing tangent
modulus, which is, however, much larger than that of the aluminum.
The shells were tested in a displacement-controlled testing machine. The loading plate at

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
the small end of the conical shells was beveled to ensure a bearing surface normal to the
generators of the cone, while at the large end the loading plate was fiat. No lubrication was
used. In the tests the load increased steadily to a maximum and then decreased slowly as
the buckling deformation progressed. A representative load-axial deflection curve, for one
of the stainless steel cones (see Figure 16.59), shows the typical limit load behavior of
axisymmetric plastic buckling (see Figure 16.29a). Testing was continued until the load
decreased to about 90 percent of its maximum value.
Buckling was always confined to a single axisymmetric bulge, localized near the small
end (one should remember that in a uniform conical shell the axial stress is a maximum at
the small end). The bulge was practically identical for all specimens of the same geometry,
even when the material was different (two aluminum specimens were of the same dimensions
as the stainless steel ones). This can also be seen in Figure 16.57, showing cross-sections of
two buckled specimens, one of aluminum and one of stainless steel.
"There appeared to be sufficient friction between the small end of the cone and the beveled
loading plate to prevent any slippage. This was checked by coating the loading plate with
blue marking ink. A well-defined circle of contact was left after the test, with no appearance
of slippage." Direct measurements of diameters at the small end also confirmed that there
was no change in the middle-surface diameter, though the flattening of the rounded small
end by local plastic flow made precise determination of the middle surface rather difficult.
The spread of the plastic zone from the small end was observed in eight specimens by
means of photoelastic coatings and a reflection polariscope, using both high-sensitivity and
low-sensitivity coatings (supplied by Photolastic Inc.), with up to 5 percent and 50 percent
elongations, respectively. The coatings were cast in sheets approximately 1.3 mm thick, cut
into strips covering a quarter circumference, and after complete curing were bonded to the
conical shells. The different-sensitivity coatings were located so that their edges butted to-
gether along a meridian of the cone, permitting simultaneous viewing of both sets of fringes.
Black-and-white photographs of the fringes, formed by a green light from a mercury vapor
lamp, were taken at different values of the applied load. The positions of the plastic boundary
determined from the fringe order were found to agree well with membrane analysis.

60C
0
0..
:::;;
400

Al

0 4 6 8 10
3
Stra1n x 10

Figure 16.58 University of British Columbia plastic buckling tests on steep, truncated conical shells
subjected to axial compression-stress-strain curves for 6061-T6 aluminum (AL) and
for heat-treated type 416 stainless steel (SS) (from [16.133]). Note the very significant
difference between the two materials
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1362 Plastic Buckling ExperimE!nts

250,-----------------------------------,

200

Testing
stopped
.,;
0
~ 100

50

0 0.5 1.1) 1.5 2 0 2.5 3.0

A<ial deflection, mm

Figure 16.59 University of British Columbia plastic buckling tests on steep, truncated conical shells
subjected to axial compression--load axial deflection curve for a stainless steel specimen
(average (Rit)"" 10, cone angle a"" ll.3o and height= 70 mm, from [16.133])

In the case of the aluminum specimens, for which the strain increased very rapidly once
the yield point had been reached (see Figure 16.58), the position of the plastic boundary
could also be observed directly by the marked change in fringe spacing across it. The final
position of the plastic boundary in the loaded conical shell was checked by the remaining
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

fringes after unloading. No residual strains and hence fringes remained beyond that boundary.
In the aluminum shells the plastic zone extended into the cone about twice the distance of
the bulge.
In the stainless steel conical shells the strain distribution during loading was similar to
that in the aluminum ones, except that the plastic boundary moved more rapidly and reached
the large edge of the cone quite early. Therefore, after loading, residual strains appeared over
the entire shelL
Once the shape and extent of the bulge had been determined for the stainless steel conical
shells, strain gages were mounted on one steel specimen to observe the growth of bending
strains. High-elongation strain gages, capable of measuring up to 10 percent strains, were
employed. They were placed on both the inside and outside surfaces at points 5, 9 and 24
mm down from the small end. The plot of load versus meridional strain at the gage location
9 mm down from the small end, where maximum bulging occured, is shown in Figure 16.60.
The strains on both the inside and outside surfaces were compressive and continued to grow
even while the load decreased. The curves ended when the downward motion of the load
was stopped, and their difference represents the bending strain. Hence it can be seen that

~lns1de
surface

50

0.02 0.04 006 0 08 0,10


Strain

Figure 16.60 University of British Columbia plastic buckling tests on steep, truncated conical shells
subjected to axial compression-load-meridional strain curves for a stainless steel spec-
imen at point of maximum bulging (from [16.133])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1363

there was bendi ng already at half the ultimate load and that from there the bending strain
increased rapidly with load.
In Ramsey's paper a perturbat ion analysis was also developed, which provided a reason-
ably good prediction of the buckling deformatio n, compared to the measured one, but did
not predict an ultimate load as observed in the tests. The experimental pa1t, however. excels
in the detailed discuss io n of the specific experimental techniques of plastic buckling, and
has therefore been quoted and discussed here at length.
The experiments on plastic collapse of ring-stiffened conical shells under external pressure
carried out by Ross and his co-workers at the University of Portsmouth, England (13.126],
[ 16.134] and [ I 6. I 35] represent an example of modern plastic buckling tests of shells.
While the earlier tests, discussed in Chapter 13 [ 13. 126], focused o n general instability.
the two more recent series of experiments [ 16.134] and [16.135) dealt with plastic shell
instabili ty of conical shells (see Figure 16.61). The specimens were therefore thin-walled
cones with two intermediate ring-stiffeners and ftanges at either end, having four equally
spaced bolt holes for attachment to the test tank and the end closure plate. Two series, of
three conical shells each, were machined from solid stock EN I A mild steel. This steel was
chosen because of its low strength and the incl usion of lead in its matrix, which aided
mach ining; while machi ni ng from solid block was considered to yield more accurate ge-
ometries.
The internal faces of the cones were machined fi rst, and then they were placed o n a conical
mandrel for final culling of their external surfaces. The dimensions of the specimens were
carefully measured wi th a coordinate measuring machine and a proximity probe, as pointed
o ut in Chapter 13 as being the practice in Portsmouth University tests. Out-of-roundness
(initial imperfection) measuremenL~ were made in the middle of the primary test section, the
section between the two intermediate stiffeni ng rings. Typical ci rcu larity plo ts of two shells
(A and C of [16. 134]) are shown in Fig ure 16.62. The maximum o ut-of-roundness for these
three conical shells (A, B and C) was ± 0. I-0.2 percent of the wall thickness r, rather low
values. indicat ing very precise machining of the speci mens. The shape of the initial circu-
larity plots (see Figure 16.62) differed significantly from the sine shape assumed in most
theoretical studies, thus tempting the researchers to classify them as random, though this was

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
pro bably not justified (see Chapter 10). In the second series of three conical shells (cones

Figure 16.61 Universily of Ponsmouth 1994 plas1ic buck ling expe1imen1s on ling-stiffened steel con-
ical shells-Jwo parallel modes of failure of a 1ypical 19. 1° conical shell (of the first
series): axisymmet1ic collapse in lhe first (bonorn) bay and plas1ic shell inslability in lhe
middle bay (counesy of Professor C.T.F. Ross)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1364 Plastic Buckling Experiments

(a) (b)

Figure 16.62 University of Portsmouth 1994 plastic buckling experiments on ring-stiffened conical
shells-circularity plots of two 19.1° conical shells, of the first series (from [16.134]):
(a) circularity plot for cone A, (b) circularity plot for cone C

7, 8 and 9 in [16.135]) the magnitude of the measured out-of-roundness was slightly larger,
about 0.3 percent of the wall thickness, but still low, thereby verifying the accuracy of the
machining.
The specimens of two series were small cones with an average middle radius Rav = 35
mm. The first series (shells A, Band C of [16.134]) had an overall height H = 123 mm, a
cone angle of 19.1° and (R,jt) = 32-35; whereas the second series (of [16.135]) had H =
212 mm, a cone angle of 10° and (RaJ t) = 43-44. The shells could therefore be classified
as being at the lower end of the range of thin shells. The difference between the three
specimens in each series was the length of the unsupported shell in the central bay, a primary
parameter determining the shell instability.
The shells were tested in the same relatively simple test setup used in the earlier Ports-
mouth University general instability experiments (see Figure 13.65 and [13.126]). Ten linear
strain gages were bonded to the inner surface of each conical shell in a circumferential
direction in its central bay for detection of the shell instability modes, and four or five rosettes
or 90° gages were placed along one of its meridians to measure the meridional strain distri-
bution.
In the tests of the first series (conical shells A, B and C of [16.134 ]), yielding was observed
before shell buckling occurred. As a matter of fact, the first pressure drop appeared at 89-
96 percent of the ultimate pressure, and only after further pressure increases did inelastic
shell buckling, at the ultimate pressure, appear in the middle bay, with four to five lobes.
After the removal of the specimens from the test tank it was found that axisymmetric plastic
failure had occurred between the large end of the cone and the first intermediate stiffening
ring in a manner that formed an artificial stiffener in the lower bay (see Figure 16.61). The
axisymmetric plastic stiffening enabled the shells to carry additional pressure till they even-
tually failed by lobar shell buckling. In the second series of tests (conical shells 7, 8 and 9
of [16.135]) this clear axisymmetric yielding and plastic stiffening of the lower bay appeared
in only one of the specimens. The phenomenon is, however, of considerable importance for
experimenter and designer alike.
As for cylindrical shells, for other types of shells, such as conical, spherical and tori-
spherical shells, many of the experiments discussed in Chapters 9 and 13 extended into the
plastic range and should be consulted in conjunction with the tests considered here.
For spherical shells, extensive experimental studies of plastic buckling were initiated in
the early sixties, in the wake of the abundance of their elastic buckling tests in the fifties
and beyond, discussed in Section 9.10 of Chapter 9. The plastic buckling test programs
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

carried out at the U.S. Navy David Taylor Model Basin (DTMB) at the time, in conjunction
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1365

with their clastic instability tests [9.1 50], [9.5 13], [9.324], [ 16.136] and [ 16. 137], were
among the earliest systematic inelastic investigations of spherical shells. A representative
example is Krenzkc's 1962 study [9.150], discussed in Subsection 9.10.3. Since the DTMB
researchers were interested in the design of deep submergence vehicles, Krenzke primarily
considered thicker shells, and out of his 26 aluminum alloy hemispherical shell s, with
(RI 1) = Il -l 05, only 5 could buckle elastically. while the remainder were in the plastic
range.
1ltc tests were designed "to investigate the strength of accurately fabricated deep spherical
shells wi th ideal boundaries and to provide a foundation for future experimental as well as
theoretical treatment," in particular for inelastic failures . The specimens were small , 0.80-
0.86 in. (;;;r20-22 mm) mean radius. precisely mach ined hem ispherical shells, bounded by
ring-stiffened cylindrical shells to provide ideal boundaries (see Figure 16.63). The stiffened
cylindrical portions of all models except three were designed to provide boundary conditions
of membrane deflections and no rotation of the edge of the hemispherical ends. These ideal
boundaries were intended to ensure that the stress patterns in the hem isphere were identical
to those in a complete sphere. It was realized that the resu lts obtained on these very accu-
rately machined models with ideal boundary conditions were of little direct value to the
designer and could only serve as upper bound references. but as such they indeed provided
the desired foundation for future efforts. They also provided the data for a practical empiri cal
formula, based on a plasticity reduction factor and the observed collapse strength of the
clastic models.
Simi lar studies on larger hemispheres made of high-strength steel were conducted at
DTMB in the followi ng years. For example. Dadley's investigation [9.324] dealt with five
hemispheres of 3.2-4.0 in. (;:81-101 mm) mean radius and one of 2.0 in. ("'51 mrn) mean
radius. all precisely machined from high-strength, special-treatment steel (STS). As before,
the models were supported by ring-stiffened cylinders, designed to maintain lower stresses
in the cyli ndrica l portion than in the spherical one and thus to permit the hemispheres to
si mulate complete spherical shells. Of the six hemispherical shells, four were relatively thick-
walled, witlt (RI 1) = 27- 60, and therefore in the plastic range; whereas two shells, one with
(RI I ) = 145 and in particu lar one with (RII) = 304. buckled elastically.
The study was later extended to larger, fabricated HY-80 steel shells by Kieman and
Nishida (sec [16.136]). Forty models. 16 of 15 in. (;;,3 18 mm) mean radius and 24 of 33

\

~~
~ ~
c

CJ
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 16.63 U.S. Navy David Taylor Model Basin 1962 inelastic buckling tests on machined deep
spherical shells under cxtemal hydrostatic pressure-606l -T6 aluminum models after
collapse (from [9. 150)). Note that the collapse mode is usually away from the boundary
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1366 Plastic Buckling Experiments

in. (20838 mm) mean radius, were tested. Thirty-six were hemispheres, fabricated as in
practice, by pressing and welding six 60° "orange peel" segments and a 60° spherical cap,
and four were spun ellipsoidal shells. The models were divided into three series, one with
(Rit) =
160, one with (Rit) 20 96 and one with (Rit) 20 60. The main purpose of the tests
was an assessment of the effects of initial imperfections, residual stresses and penetrations
on the elastic behavior of the shells and collapse strength. Most of the shells were sufficiently
thin-walled to buckle elastically, but in some of these imperfect hemispheres local inelastic
strains occurred, changing the collapse mechanism to inelastic buckling.
The investigators realized that the effect of initial imperfections, particularly the out-of-
roundness, was the main contributor to their significant experimental scatter. They therefore
introduced very detailed measurements of initial imperfections, about 1000 radius measure-
ments on each model, and presented them as contour maps for each specimen (see for
example Figure 16.64). In addition, about 125 thickness measurements were taken on each
model by ultrasonic techniques. A special sphericity measuring instrument, with an accuracy
of 0.001 in. (about 0.4-1.0 percent of the wall thickness), was used for the imperfection
measurements, which were processed by the DTMB Applied Mathematics Laboratory com-
puter. It was found that the collapse strength of spherical shells with initial imperfections
depended primarily on the local geometry over a critical arc length. Based on this obser-
vation, an imperfection analysis was developed (see [16.138]) that permitted calculation of
membrane stresses from the reduced initial imperfection data, which were usually within 10
percent of the stresses determined from experimental data. In all, a rather advanced procedure
at the time, well on the way to the modern methods discussed in Chapter 10.
Costello and Nishida extended the external pressure buckling strength investigations of
HY-80 steel hemispherical shells to similar shells with thicker walls (see [16.137]). They
tested eight 33-in. (20838-mm) radius specimens, divided into two groups, one with (Rit) =

I
~.·1.03
R 11 ~·1.12
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

l
h1 •L03 h, •1,02
A1111 ·1.11 "1/111 •1.20

Figure 16.64 U.S. Navy David Taylor Model Basin 1966-7 buckling tests on fabricated HY-80 steel
spherical shells-plot of typical deviations from sphericity, here for specimen 81, which
failed by plastic buckling (from [16.137]). The surface enclosed by the solid circle rep-
resents a hemisphere unfolded into a fiat surface whose radial scale remains constant.
Contours are plotted in mils, minus contours indicate inward deviations
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Plastic Buckling Experiments 1367

33 and one with (Rit) = 41, both well in the plastic range. The models were fabricated from
seven pressed and welded segments, as in the earlier test series, and half of them were stress-
relieved prior to testing. An internally ring-stiffened cylinder was welded to each model (see
Figure 16.65). Since, because of fabrication tolerances, it was not practical to design the
cylinders to ensure membrane boundaries precisely at the junctures of hemisphere and cyl-
inder, the cylinder walls were arbitrarily thickened by about 10 percent over those that would
provide membrane conditions.
Again, comprehensive initial imperfection measurements were taken, about 1000 radius
measurements on each model from a fixed point within the shell to the inside surface. The
results of these measurements were again presented in the form of contour maps (see for
example Figure 16.64) "that represent the inside view of a hemisphere unfolded into a fiat
surface whose radial scale remains constant. This scale can be determined by dividing one
half of the circumference of the sphere by the diameter of the contour map." Mismatch
between the welded segments in terms of the deviation from sphericity is also given in the
plots like Figure 16.64, but the midsurface shell contours have not been corrected for shell
thickness variations.
Each specimen was instrumented with approximately 70 foil strain gages, their locations
being selected on the basis of fiat spot calculations, mismatch data and thickness measure-
ments.
The models were tested in oil in the DTMB 6-ft. testing tank (see Figure 16.65). Note
the sturdy hydrostatic testing tank.
The measured local geometry in the area of failure was used to calculate the buckling and
yield pressures. By utilizing the imperfection analysis of [ 16.138] and [ 16. 136] and extrap-
olating previous test results of [16.136], it was possible to predict the collapse pressures
within 10 percent of the experimentally observed ones.
We may point out again the outstanding quality, innovation, precision and detailed plan-
ning of the experiments in shell stability carried out at the U.S. Navy David Taylor Model
Basin (called now Naval Ship Research and Development Center) in the fifties, sixties and
seventies, which make them a source of inspiration even today.
Before we leave the topic of spherical shells, we should remember that, as in the case of
elastic instability, concentrated loads may cause plastic buckling in thick-walled spherical
shells. An example of such an experimental study of the inelastic response and instability
of spherical shells loaded by a concentrated inward load applied through a rigid boss is that
carried out by Leckie and Penny at Cambridge University in the late sixties [ 16.139]. Their
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

HOOT HYOROSTATIC
TESTING TAHK

Figure 16.65 U.S. Navy David Taylor Model Basin 1966-7 plastic buckling tests on fabricated HY-80
steel spherical shells-test setup in DTMB 6ft testing tank (from [16.137])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1368 Plastic Buckling Experiments

efforts to manufacture accurate shells are of particular interest. At first, shells were spun
from soft aluminum sheets, but their geometry was very variable and it was difficult to
choose the appropriate value of the yield stress. It was then decided to manufacture the shells
from the solid. Dural, chosen as the material, had a well-defined yield stress. A special
method of fabrication was persistently developed till it had become a simple process that,
however, still yielded accurate shells. The process consisted of rotating the workpiece (which
eventually became the hemispherical shell) about one axis while an offset cutter in a milling
machine was rotated about another axis at a vastly different speed. Because of the speed
difference, a surface of revolution was generated, whose cross-section at any level was a
circle. Eventually a sphere with its center at the intersection of the two axes of rotation was
obtained.
Five shells, of radius R = 3 in. ("'E76 mm) and wall thickness t = 0.05 in. ("'El.27 mm),
were made. The shells had four different boss sizes, two specimens having the same boss
size for checking of repeatability, which was found to be very good.
The shells were loaded through a screw-driven cross-head. A load cell measured the
applied load. The movement of the cross-head was also measured and thus experimental
load-displacement curves could be plotted, which showed the character of the buckling and
postbuckling responses. Comparisons of ultimate loads with predictions of rigid-plastic the-
ory showed them to exceed the experimental values, as expected.
Buckling of practical torispherical shells is also mostly elastic-plastic or plastic buckling,
as pointed out in Subsections 9.11.2 and 9.11.3 of Chapter 9.
For example, in the many experimental investigations on torispherical heads under external
pressure carried out at the University of Liverpool in the eighties (see Subsection 9.11.2 and
[9.356], [9.363]-[9.366], [9.370] and [9.371]), the failure of most specimens was by elastic-
plastic or plastic buckling. Different failure modes were observed: axisymmetric yielding of
the knuckle, edge buckling, usually appearing as a single asymmetric buckle or as four lobes,
or an axisymmetric buckle at the crown of the torispherical head. The initial imperfections
apparently influenced the type of plastic collapse mode that eventually occurred in the test.
The buckling of tori spherical shells under internal pressure is, however, the more important
phenomenon. It was therefore extensively studied at a number of research centers in the last
decades (see for example [2.59]-[2.61], [9.163]-[9.166], [9.372], [9.383]-[9.393] and
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

[9.379]-[9.381]) and again, as stressed in Subsection 9.11.3, most of the shells tested failed
by elastic-plastic buckling or plastic buckling.
Many of these plastic buckling tests were discussed in detail in Chapter 9 and need not
be recounted here. However, the behavior during and after plastic buckling of internally
pressurized torispherical shells warrants further consideration. The discussion will partly
follow the representation by Roche and Autrusson in their 1986 paper on the Saclay exper-
iments [9.390].
The reader may remember that internally pressurized torispherical shells can fail either by
axisymmetric collapse or by buckling, usually inelastic, which involves circumferential waves
in the knuckle region. In the description of their test procedure, and in particular of the
actual test results, Roche and Autrusson emphasized that the buckling pressure obtained
depended notably on its definition. They proposed three definitions, each of which suited
certain geometries of the torispherical shells tested.

Definition A: "For thin heads, the examination of the axial deflection recording versus
pressure shows a sudden drop in pressure as buckling occurs. This drop in pressure can
be explained as follows. When the buckling waves appear, they create bending stresses
that cause local plastic instability, and hence the formation of a plastic buckle. This buckle
increases the volume, and, since the delivery of the pump is low, it cannot maintain the
pressure, so that the pressure drops sharply.
It should be noted that this occurrence is very rapid, and that the place where the crease
(plastic buckle) is formed is not predictable ... This buckle is not the simple amplification
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Combined Loading Tests in Plastic Buckling 1369

of an initial geometric imperfection. When the buckle was clearly visible, the pressure
preceding the sudden drop was taken as the buckling pressure."
Definition B: "No sudden drop in pressure occurred during buckling of thick heads. Nor
did a buckle suddenly appear, but rather the progressive formation of buckling waves."
For the thicker shells, with knuckle radius ratios (rlt) = 20-30, "the buckling pressure
was taken as the value of the pressure at which it was possible to detect dimples or creases
with the naked eye under intense light."
Definition C: The investigators found that Definition B was barely satisfactory and could
lead to an overestimation of the buckling pressure. They therefore introduced a rotary
probe to measure the initiation of buckling in the knuckle area. The records of this probe
showed that at a certain pressure a low-amplitude sinusoidal waviness occurred fairly
suddenly, with a large number (20-40) of circumferential waves, which rapidly trans-
formed, as the pressure increased, into a much smaller number ( ~6) of visible waves.
The initiation of the low-amplitude waves was taken as the third definition of the buck-
ling pressure, which, however, led to much lower values than the second definition.

The just perceptible low-amplitude waviness or ripples that preceded visible plastic buck-
ling waves in the Saclay tests were similar to such ripples noticed in the earlier Nottingham
tests (see Subsection 9 .11.3 and [9.388] and [9.389]) and somewhat similar to the precollapse
ripples observed by Reddy and Kyrikides in their bending experiments on cylindrical shells
(see Subsection 16.2.8 or for example [16.108] and [2.71]).
The postbuckling behavior of the torispherical shells under internal pressure in the Saclay
tests was stable, similar to that observed in the Nottingham tests. Buckling did not cause
fracture or total collapse of the torispherical head but was manifested by the formation of
an initial buckle, which was followed by further buckles as the pressure continued to rise.
Pressures could be raised far beyond the buckling pressure (more than twice the buckling
pressure in about half of the Saclay tests). No significant time delays between the attainment
of the critical pressure and formation of buckles, as observed in the Nottingham tests, were
noticed in the Saclay experiments.
Because of the stable postbuckling behavior of the Saclay torispherical shells under in-
ternal pressure, the influence of initial imperfections in their buckling pressure was found to
be relatively small. Three of the shells had significant geometric imperfections in the form
of creases or folds in the knuckle area. In two of them the height of these creases was 80
percent and 200 percent of the wall thickness, respectively, but the folds did not grow and
buckling occurred at a different location, with no or small reduction in the buckling pressure.
The remaining shell, however, exhibited a large initial crease with a height four times the
wall thickness. In this case the initial fold grew during the test and precipitated buckling at
a lower pressure, 30 percent lower. In view of the very large initial imperfection, this re-
duction was no surprise.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

In parallel with the experiments on torispherical heads, similar investigations were carried
out on ellipsoidal shells, with similar results (see for example [9.162], [9.386], [9.387] and
[9.385]).

16.3 Combined Loading Tests in Plastic Buckling


Combined loading appears in plastic buckling as it does in elastic instability. Therefore,
many of the salient features, problems and test techniques discussed in the relevant sections
of Chapters 6, 8, 9, 13 and 15 also apply in the inelastic case. But in the plastic region the
interaction may be broader and different since it involves the nonlinear material behavior.
For example, the etiect of a tension load added to a cylindrical shell subject to external
pressure, which would be stabilizing in elastic buckling, may be destabilizing in the plastic
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1370 Plastic Buckling Experiments

range due to softening or reduction in the stiffness of the material beyond the yield stress.
as will be shown in detail in Subsection 16.3.2.

16.3.1 Biaxial Loading in Columns and Plates


Very few experimental studies on the inelastic behavior of columns subjected to biax ial
loading have been reported in the literature. An example is the tests on biaxially loaded
rectangular tubular steel columns carried out by Ra11aq and McVinnie in the mid-eighties
[16.140). They studied the inelastic behavior of square and rectangular tubular columns in
response to a constant small or medium axial load combined with a gradually increasing
biaxial end moment.
Special gimbals were developed to permit free flex ural rotation of the column ends about
any horizontal axis. The speci al gimbals may be of interest to experi menters and there fo re
warrant detailed discussion here. thoug h they do no t apply specifi cally to inelast ic buckl ing.
but ntther to the entire clastic-plastic range.
A pair of these highly precise Meel gimbals was developed to provide the simply ~upported
end conditions assumed in the analysis (derived in 116.141 ]). to whose predictions the test
resuiLs were compared. Figure 16.66 shows one of these end fixtures. "The inner party of
the gimbal. with eight holes. rotates on a shaft supported by self-aligning spherical roller
bearings housed in walls of the o uter part ... The ends of the test column arc each welded to
a 1.9-cm {l>:0.75-in.) baseplate. which is then "bolted (with eight bolts) to the inner part of
the end fixture. This provide~ nearly friction lc~s rotatio n at the end of the specimen. about
the centerline of the sha ft that correspo nds to the theoretical y-axis. The o uter part of the

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(a)

(b)

Figure 16.66 Biaxially loaded lubutat columns-an end fix1urc simutali ng a biaxial hinge (from
[ 16.140] . cour1esy of Professor Z. RaZLaq): (a) view of lhe lop end gimbals from below.
(b) closeup of iMide of end fix1ure
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Combined Loading Tests in Plastic Buckling 1371

gimbal rotates on two stub shafts that are firmly attached to its built-up walls and supported
by bearings in pillow-blocks." Nearly frictionless rotation of the entire gimbal is therefore
achieved, about the centerline of the stub shafts that corresponds to the theoretical x-axis.
The combination of the two rotations (the gimbal action) permits rotation of the ends of the
test column about any horizontal axis. The inner parts of the gimbals at both ends of the
column are therefore considered rigid parts of the specimen and are added to the effective
column length (in the manner discussed in Subsection 6.1.5 of Chapter 6). When a load cell
was added between the top end gimbal and the test column, it too had to be added to the
effective column length.
In the vertical test setup "the top end gimbal is attached to a movable cross-beam. Plates
with machined spherical recesses are attached to the cross-beam .... Ball bearings are placed
in the recesses with a total of eight balls used at each end. These ends fit into a pair of
guides and ride on smooth vertical inner surfaces to give free vertical movement. Horizontal
adjustment of the movable cross-beam is provided by the threaded rods and used for align-
ment of the specimens."
Razzaq and Me Vinnie tested three columns with square cross-sections and two with rec-
tangular ones. In the tests the axial load was applied by two hydraulic jacks acting on the
cross-beam and was kept constant. The biaxial moment was applied by means of a lever
arm attached to the inner part of the end fixture. A moment-producing load W applied a pair
of moment components at the lower end of the column A, M~ about the x-axis and M'A about
the y-axis, whose ratio depended on the ratio of the load eccentricities e' and e-'. In these
tests the angle of attachment of the lever arm was such that the moment ratio (M~ I M~) =
(e '/ e') = 0.5 and was kept constant.
The comparison of the experimental and theoretical peak moments and moment-deflection
curves showed very good agreement (within 2 percent).
Plate elements subjected to in-plane compressive biaxial loading frequently occur in ship
structures. Their buckling behavior under combined loading has therefore been of concern
to naval architects for decades, and one may expect to find examples of experimental inves-
tigations in the relevant literature. However, as mentioned in Subsection 8.6.2 of Chapter 8,
the earliest pertinent experimental investigations, though preceded by extensive theoretical
and numerical studies, were apparently those of Becker and his co-workers, sponsored by
the U.S. Ship Structures Committee [8.154]-[8.156] and reported only in the seventies.
Becker's studies, and the later Cambridge University ones [8.157], were primarily aimed
at elastic buckling of plates subjected to biaxial and polyaxial loading (biaxial membrane
loads combined with lateral pressure), but for the smaller (bit) ratios failure was the result
of inelastic buckling. Some of the salient experimental techniques employed in the Ship
Structures Committee studies, as well as those used by the Cambridge University team, were
discussed in Subsection 8.6.1 of Chapter 8. Here only some points specifically related to the
plastic range will be mentioned.
The results of Becker and his co-workers [8.154]-[8.156] showed a definite dependence
of the load interaction behavior on the (bit) ratio. For their thickest plates, with (bit) = 30,
where failure was by biaxial plastic buckling, a large reduction in longitudinal strength was
observed when transverse membrane stress was applied, as can be seen in Figure 16.67,
presenting some of the Ship Structures Committee test data. In the figure, test results for
only two groups of specimens have been reproduced for clarity, one for (bIt) = 30, and one
for (bit) = 90, well in the elastic region. In Figure 16.67a two different types of interaction
behavior are evident: one for plastic buckling, which is close to the linear interaction curve
shown, and one for elastic buckling, which exhibits a convex nonlinear interaction curve
(not drawn in the figure). The test data for two other intermediate groups (not shown here),
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`

with (bit) = 50 and 70, indicate interaction curves that fit the same pattern, with the spec-
imen having (bit) = 50 representing a transition between the two types of curves. Similar
types of interaction curves were observed in the later sse tests [8.155], as well as in the
Cambridge University tests [8.157], in particular for their thickest plates, with (bit)= 35.0-
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1372 Plastic Buckling Experiments

4.0 ------.,----,---

3.0

ky
2.0

INTERACTION
CURVE
1.0 0

0 b/1 ; 30 0

6 bit; 90 0

(a) oL____i_~--~--~~--~
0 0. 25 0. 50 0.75 1.00

I
O~JOxu

1.0 (b.{) = 30

~ 0 "
0
0
(Oy jao)
0.5

(b!() :90
-t:::.--A-- 0
..............

(b) 0 L.___ _......L__ _ ----:c'::-~


0 0.5 1.0

Figure 16.67 U.S. Ship Structures Committee biaxial strength data for two groups of specimens, one
with (bit) = 30, in the plastic range, and one with (bit) = 90, well in the elastic region
(from [8.154], with omissions): (a) plot of the transverse buckling coefficient k, versus
the axial stress ratio, (b) plot of transverse stress ratio versus the axial stress ratio, where
a 0 is the uniaxial yield stress

35.8, that showed plastic buckling behavior. One should note that the biaxial strength data
in Figure l6.67a is presented as the transverse buckling coefficient k" versus the axial stress
ratio (a) a",). Since the transverse stress a, is related to k, by ·
(16.16)
the more logic plotting of the transverse stress ratio a,.! a 0 , instead of k,, versus the axial
stress ratio aJ a 0 (where one assumes a"' = a 0 ), would have shifted the biaxial strength
data points by [1/(blt)Z]. This would have inverted the positions of the interaction curves,
with the (bl t) = 30 data moved to the top and that for (bl t) = 90 to the bottom (see Figure
16.67b-note that the curves drawn in the figure only indicate the trends).
A couple of years later Dier and Dowling presented a numerical procedure to analyze the
elasto-plastic large deflection behavior of plates and applied it to a parametric study of the
strength of plates subjected to biaxial in-plane forces [16.142]. They compared their calcu-
lated interaction curves with the U.S. Ship Structures Committee biaxial strength data, ob-
tained on steel square cross-section tubes [8.154]. The comparison is shown in Figure 16.68
(from [ 16.142]). The plates tested had an aspect ratio of 3 and (bIt) ranged from 30-90.
The test results are classified according to the sequence of loading, which apparently has an
effect. This influence was, however, difficult to determine in the absence of repeated tests
of identical geometries with different sequence of loading.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

The computed results generally gave a lower bound to the test results, and the agreement
Copyright Wiley
was considered satisfactory by the authors, though
Provided by IHS Markit under license with WILEY
there was considerable scatter even for
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Combined Loading Tests in Plastic Buckling 1373

~
8 Expenmenlcl
<>y
~ res.J:ts (Ref 8.15')
a=;; b/t
? 31 9
b/ l • 30 0 53 2
c. 74 4
0 95 7

5 "
" •" .
,.,.,_ ax compOnoz'1l
applied 1as t
rry ccmp::ment
applrozd !est

Figure 16.68 U.S. Ship Structures Committee biaxial strength data (from [8.154]) compared with
interaction curves obtained by numerical elasto-plastic large deflection calculations (from
[16.142]). The plates have an aspect ratio of 3, and a 0 is the uniaxial yield stress.

the same sequence of loading. Dier and Dowling stressed the paucity of tests for plates under
biaxial loading, which was also underlined by C.P. Elinas, W.J. Supple and Walker in their
review of buckling of offshore structures [6.14]. The authors wish therefore to re-emphasize
the need for further tests, in particular in the inelastic region.

16.3.2 Biaxial Loading in Cylindrical Shells


Whereas only sporadic experimental studies of plastic buckling of columns and plates sub-
jected to combined loading were performed, in shells the phenomenon was extensively in-
vestigated, the motivation coming primarily from problems in the oil industry, in particular
those arising in pipe laying in deep waters (see for example [9.243], [9.255], [16.114],
[16.143] or [16.144]). The extent of the relevant experimental work is exemplified by one
of these studies, performed at the Battelle Columbus Laboratories in the seventies [16.114].
Here over 200 small-scale pipe tests were carried out on specimens subjected to bending
and combined bending and external pressure, as well as a number of full-scale bending and
external pressure tests on long 20-in. diameter pipes.
The combined loading tests on plastic buckling of cylindrical shells fall broadly into two
groups: external pressure loading combined with bending, and external pressure loading
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

combined with axial compression or tension. Most of the work in the first group was carried
out at the Battelle Columbus Laboratories (see [16.114]), at the University of Strathclyde in
Glasgow (see [16.144]), and at the University of Texas (see [9.255] and [9.243]). The in-
vestigations in the second group can be subdivided into those whose axial loading is com-
pression, mainly performed at the Structural Research Laboratory of the University of Cal-
ifornia, Berkeley (see [16.117]), Liverpool University (see [9.116]) and the Chicago Bridge
and Iron Company (CBI) (see [16.145]), and those whose axial loading is tension, carried
out at Caltech [16.143], [16.146] and [16.148].
The University of Texas investigations on plastic buckling of cylindrical shells subject to
combined external pressure and bending [9.255], [9.243] and [16.112] were mentioned in
Subsection 9.7.2 of Chapter 9, in connection with bending instability. The combined bending-
pressure experiments were carried out in the test setup shown in Figure 16.69, which is
essentially a development of the Texas University pure bending facility (discussed in Sub-
section 9.7.2; see Figures 9.76 and 9.77) with a special matching pressure vessel. This 20-
in. (508-mm) diameter pressure vessel, into which the bending rig rolls, has a working
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1374 Plastic Buckling Experiments

(a)

Pressure Gages

Figure 16.69 Umven.uy of Texas combined bending-pressure experiments on plastic buckling of cy-
lindncal ~bells-the combined pressure-bendtng tc't setup (from (9.255)): (a) .-iew of
the tc't facility. (b) schematic diagram of the facility in operation

pressure of 5000 ps i (350 bar) and the test dev ice is operated remotely. as pointed out in
Section 16.2.8.
Seamless. drawn stainless steel 304 tubes of 1.25 in . (3 1.8 mm) diameter, with (Rit) =
12.3 and 17.4 and lengths ranging between 18-24 tube diameters. were tested under pure
pressure. pure bending and their combinations. Under combined loading the collapse pres-
sure, the maximum bending moment capacity and the corresponding curvature were found
tO be reduced due to the interaction of geometric and materials nonlincaritics. Geometrically.
the two loads interacted through tube ovalization, while the material interaction occurred
through inelastic effects. S ince the Iauer are path dependent. both the response of the tubes
10 combined loading as well as their instabilities were found in the tests to depend on the
loading path. A set of typical measured moment-curvature responses for different values of
pressure are presented in Figure 16.70a. and typical collapse panems are ;hown in Figure
16.70b.
The Universi ty of Liverpool 1985 investigation (9.116 ) included combined loading tests
on 35 steel shells ( 17 machined and 18 welded shells of 150 mm diameter with (Rir) ""
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

100). The test results indicated. however. that as expected. for most o f the shells buckli ng
was elastic, while only for some. primari ly the shorter one~ with L/ R "' 0.33 (Z"' I0). the
interaction curve. or part of it. rcprc~entcd plastic buckling. The shells were tested in a
special combined-load facility shown in Figure 16.71. Axial compressive forces were applied
in this test rig by pressurizing the lower pan of the cylinder. while external latcml pressure
was applied independentl y by prcssuriting the annulus around the model and the load cell.
The dimensions of the models (out-of-roundness, out-of-straightness and thi ckness) were
measured prior tO final installation in the test setup. and the radial displacements were mea-
~ured in situ. during the test_ by an LVDT mounted on a rotating central shaft.

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Combined Loading Tests in Plastic Buckling 1375

I2
!1.
•• ss- 30'
110 oin
f:: 12 . 3

08

06

o•

02

{a) 0 02 0 <1 06 08 1.0 1.2


_,I 4 16
.... ....
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

{b)

Figure 16.70 University of Texas combined bending-pressure experiments on plast ic buck ling of cy-
lindrical shells (from [9.255]): (a) typical measured moment-curvature responses fo r
stainless steel tubes with (R it) = 12.3 and different pressures. (b) test specimens col-
lapsed at different pressures (and curvatures)

Figure 16.71 Liverpool University combined axial compression- external pressure experiments on
150-mm diameter steel cylindrical shells, with Rlt "' 100- diagram of combined-load
test facili ty (from [9.116])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1376 Plastic Buckling Experiments

The Berkeley experiments mentioned above [ 16.117] represented a full-scale laboratory


test program for evaluation of the structural behavior of the 48-in. diameter trans-Alaska
pipeline, subjected to compressive axial forces, to lateral loads causing bending and to in-
ternal pressure, a combination actually beyond the bounds of the second group. The tests
were carried out in the 4,000,000 lb capacity Southwark-Emery universal testing machine at
the Richmond Field Station of the University of California, Berkeley. The well-planned test
setup, the extensive instrumentation, the test procedure and results are discussed in detail in
[16.117], which is therefore well worth reading in preparation for similar large-scale tests,
though it was written over 25 years ago.

16.3.3 Caltech Tests on Combined External Pressure and Axial Tension


The more recent smaller-scale Caltech studies on combined external pressure and axial ten-
sion loading are of special interest for a different reason. They deal with a biaxial loading,
one of whose components is structurally stabilizing but materially destabilizing, representing
a phenomenon peculiar to plastic buckling. Hence they warrant a more detailed discussion.
The first Caltech experimental program studied the plastic buckling of short, thick-walled
cylindrical shells under combined loading of external pressure and axial tension, applied in
a simple nonproportional manner. In these tests the prebuckling state was axisymmetric and
buckling was by bifurcation (for the perfect structure) into a nonsymmetric mode containing
several circumferential waves. Suitable shell wall thickness and length were chosen, such
that the shell was already in the plastic range during prebuckling and bifurcation occurred
in the plastic range. In this biaxial loading the external pressure induces compressive hoop
stresses that are structurally destabilizing, because of the possibility of buckling (structural
instability), while the axial tension is structurally stabilizing, through stiffening of the shell
against buckling and through the observed reduction of initial imperfections. Axial tension,
however, is simultaneously materially destabilizing since it moves the material farther into
the plastic region and reduces the stiffness of the structure (material instability). It was
therefore possible to observe buckling of the test shell under constant external pressure with
increasing axial tension when the shell was loaded in the plastic region.
In order to provide useful data it was considered imperative to carry out a large number
of experiments. Hence, a low-cost experimental setup (including computer-controlled data
acquisition, function-controlled loading, displacement sensing of shell wall deformation),
using inexpensive test specimens, was developed.
The tests were carried out by attaching the shell structure to appropriate end plugs that
could be axially loaded in a conventional testing machine. The pressure was applied using
a pressure chamber (a thick-walled cylinder with two round removable end plates) surround-
ing the test specimen (see Figure 16.72). Seals at both ends allowed the axial load to pass
through the pressure chamber end plates, as shown in the figure. The steel tubes extending
from the test chamber were not part of the test specimen, but were end plugs that supported
the test shell. Epoxy was used both to provide a watertight seal between the end plugs and
the specimen and to hold the specimen firmly in place during testing.
Pressurization of the test chamber and the test specimen was accomplished by pumping
a hydraulic fluid into the chamber, while axial loading was provided by an external loading
device in which the test chamber was mounted (essentially an MTS system; see Figure
16.73). Feedback-controlled servo-controllers were used to provide accurate loading.
The data acquisition system was set to record periodic pressure and tension values during
the experiment. In particular, each time circumferential scans were made of the shell wall,
the load data were recorded concurrently.
The test specimens were drawn 6061-T4 aluminum cylinders, cut to the required length
from a standard 12-ft (3.66-m) tube with 1.5-in. (3.81-cm) diameter and 0.028-in. (0.711-
mm) wall thickness. Thus, (Rit) = 26.8. The specified thickness was 0.028 in. (0.71 mm),
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Combined Loading Tests in Plastic Buckling 1377

/ENDF\.UG

D'i '40 '"5


--------~---------

0~0?7?1 ~ ~0
,.
(b)

Pressure
Transducer
Feedback

Pressure Signal

Load SiQnal
(c)

Figure 16.72 Caltech plastic buckling tests on short cylindrical shells subjected to biaxial loading-
test chamber and test shell attachment (from [16.146]): (a) section of test chamber, (b)
bond of shell to end plug, (c) hydraulic configuration of external loading device

but the actual wall thickness varied from tube to tube. It is important to recognize that the
tubes were of regular stock and not specifically manufactured for this particular use and that
they were not subjected to any special processes or treatments to minimize geometric and
material imperfections.
A 1.5-in. (3.8-cm) long test shell segment was used in the experiment with Ll D = 1. The
actual shell was 3.5 in. (8.89 em) in total length, to provide l-in. (2.54-cm) adhesion surface
at both ends of the shelL The shell-to-end plug bond provided a means of transmitting the
axial tensile load from the end plug to the test shell and served as a seal to prevent the
hydraulic fluid from entering the scanning section (see Figure 16.72b). Special care was
needed in surface preparation and bonding to ensure a satisfactory bond. Once this was
achieved, the aluminum test shells were able to substantially yield before the bond material
failed.
After each test shell was cut from the 12-in. (3.66 m) tube, a series of measurements was
made using a micrometer to determine the average wall thickness of each test section. Wall
thickness varied between 0.026 and 0.032 in. (0.66 and 0.81 mm) for all test shells, with
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1378 Plastic Buckling Experiments

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Figure 16.73 Caltcch plastic buckling test• on short cylindrical shells subjected to biaxinl loading-
experimental setup in MTS testing machine (from I 16. 147), courtesy of Dr. J.J. Giezen)

most shells averaging near a 0.030-in. (0.76-mm) wall thickness for the first series of spec-
imens (set A) and 0.028 in. (0.71 mm) for the second series (~et B). The end plugs were
machined prior to testing, to fit the vari ous test shells.
Geometric imperfections were measured with a special displacement sensing device
(OSD). The ampli tudes (n > I) were found to be a maximum of 0.003 in. (0.076 nun).
which was relatively small considering that no special care was taken in the manufacturing
process to minimize such imperfection~. Most initial imperfection scartS showed a" = three-
or four-wave imperfection pattern, which could have been the result of handling, storage and
possibly a worn die. When axial tension without pressure was applied to the test shell, the
initial three- or four-wave imperfection seemed to dimini sh during scans at successively
higher axial loads, whereas imperfections increased when ex ternal pres~ure was applied.
During most tests buckling was easily detected by a sudden decrease in pressure and a
loud snap when the buckling waveform appeared. However. once the specimen was loaded
deep into the plastic range. the buckling process was much more gradual. and it was nec-
essary to develop appropriate instrumentation to detect the onset of buckling. Improved
buckling detection was accomplished through the use of a small circumferentially scanning
probe inside the cylindrical test specimen.
This displacement sensi ng device (DSD) measured the lateral deflection of the shell during
the loading process, as well as the initial geomcuic imperfections. A rotating mechanism
was designed since it was important to detect deflection at several points along the shell
wall for a complete picture of the buckling wave pauern. The probe and scanning device
arc shown in Figure 16.74. It used a noncomacting eddy current displacement measuring
probe (a Bentley Nevada probe: see Figure 16.74c). which was rotated on a shaft driven by
a small electric motor.
The circumferentia l position of the probe was given by a photo-optical device (see Figure
16.74a) that detected the teeth of a po lished gear (as in [9. 194) and [ 1.25J). which was
auached to the drive shaft. The data acquisition system then recorded the voltage of the DSD
Copyright Wiley measure of the probe and wall separation).
(a
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Combined Loading Tests in Plastic Buckling 1379

spring

pcobo~
(a) (c)

Figure 16.74 Caltech plastic buckling tests on short cylindrical shells subjected to biaxial loading-
displacement sensing device (DSD), probe and scanning apparatus (from [16.146]): (a)
schematic diagram of DSD, (b) detail of probe attachment, (c) Bentley Nevada probe

Since the gear used in the DSD had 64 teeth, a full circumferential scan consisted of 64
data points. Buckling of the 1.5-in. (3.81-cm) diameter test shells was expected to produce
four or five waves around the circumference (under this type of loading), which averages to
approximately 14 data points per wave. This was considered a high enough resolution to
observe the buckling process in detail. The DSD was used to record the growth of the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

imperfections with load and to detect buckling by recording sudden growths of displace-
ments.
Prevention of damage to the sensitive scanning probe during buckling turned out to be an
essential aspect of the experiment. Hence, a signal interruption system (SIS) was installed.
The SIS compares the feedback signal from the pressure transducer and the control signal
from the pressure servo-controller. In normal operation these signals are equal or very close.
The difference between them is the error signal that is the crucial input to the SIS. If this
error suddenly exceeds a certain preset level, indicating a momentary rapid drop in pressure
in the chamber associated with buckling, the SIS immediately ramps the load control to zero
and prevents any further deformation. This protects the probe because the deformations
remain moderate. Continuation of loading is possible only after the SIS is manually reset.
In these cases, when buckling is violent there is, however, no record of a postbuckling path.
In the cases where the tension load is very large, the postbuckling path becomes stable and
the SIS does not interrupt the loading. Since all the displacements now remain relatively
small, additional scans can be safely made during the postbifurcation phase.
The uniaxial material properties of the test specimens were measured on uniform axial
strips cut from the original stock tube. A special routine procedure was developed in order
Copyright to
Wileyspeed up this part of the pretest preparations (see [16.147]). Stress-strain data were also
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1380 Plastic Buckling Experiments

recorded during these tests with the aid of the data acquisition system and plotted for im-
mediate observation of the material behavior. A similar test was designed to evaluate cir-
cumferential material behavior from internal pressure loading for comparison with axial
stress-strain data. Although the hoop stress in the shell is compression, it was felt that
properties from a uniaxial tension test would indicate the amount of anisotropy. Figure 16.75
presents two sample stress-strain curves for a 6061 aluminum tube, which indicate the ani-
sotropy. Ramberg-Osgood fits to these curves facilitated their use in the calculations.
To demonstrate the use of the various automated systems and programs typical of modern
precise buckling experiments, a complete experimental test run will be briefly described.
First the wall thickness of the specimen is measured (the material properties having been
determined earlier), and then it is glued to the end plugs and cured. When the bonds are
completely cured, the test specimen and data and end plugs are placed in the test chamber
(see Figure 16.72). The DSD is then inserted in the proper end plug and turned till it is
locked in place, care being taken that the retracted probe does not touch the wall.
The probe signal of the DSD is carried by a coaxial cable that runs down and out of a
side slot on the lower end plug. A slipping coaxial joint just outside this slot allows the
probe to turn continuously without winding the coaxial cable and possibly disturbing the
probe position. The coaxial cable transmits the probe signal undisturbed to the data acqui-
sition system (see [16.147] for more details).
Probe positioning is the next step since the probe is still fully retracted after insertion.
The positioning bolt is turned until the probe to wall distance is within the linear range of
the DSD calibration. A digital voltmeter assists in this positioning. Next a scan is completed,
and on the basis of this scan the probe is repositioned. This procedure is repeated until a
satisfactory position, such that all circumferential points are within the linear range (but
skewed towards the higher part of the calibration curve), has been found. As the load in-
creases, the deformations will tend to be more radially inward than outward and thus the
probe reading will remain within the linear range.
Position data are picked up by the DSD and sent to the data acquisition system. An
external platform supports the pressure sleeve since the end plugs are free to move within

351?100

CIRCUMFERENTIAL
3001!10

AXIAL
25000

20008

15£1BB

10000

5000

."'" STR..'.IN C%)

Figure 16.75 Caltech plastic buckling tests on short cylindrical shells subjected to biaxial loading-
axial and circumferential stress-strain plots for typical 6061 aluminium tube (from
[16.146])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Combined Loading Tests in Plastic Buckling 1381

the pressure chamber and do not provide support. Hydraulic fluid is now introduced near
the bottom of the pressure chamber, with air escaping from a bleed valve.
The initial scan is made with no load applied. To start it, the DSD electronic circuitry is
reset, and the system is now waiting for the correct starting position. After triggering by an
external clock, data acquisition begins. The SIS is now activated and the actual loading of
the specimen can commence. The experiment is now being run by the function generator
control, and the operator only specifies when the scanning system should perform a scan.
Regular intervals, which reduce as buckling is approached, are usually designated. The more
frequent scanning near buckling assists in accurate determination of the buckling load and
also provides more points, leading to better Southwell plots, which are used for smoothing
of the experimental results.
Buckling often occurs quite suddenly, accompanied by a noticeable pop, and is immedi-
ately followed by a load decrease, due to the action of the SIS. With high tension loads the
buckling is more gradual and then the SIS does not shut off the load control. The probe will
then continue to scan, producing the postbuckling shape.
Removal of the buckled specimen is possible by reversing the assembly process. Due to
action of the SIS, the specimens showed little indication of buckling, since large postbuckling
deformations had been prevented.
Two sets of experiments (set A and set B) were carried out with cylindrical specimens
that had average wall thicknesses of 0.030 in. (0.76 mm) for set A and 0.028 in. (0.71 mm)
for set B. Measurements of the shell wall were made with a micrometer prior to the exper-
iments and were averaged over the circumference. Thickness variations of the shell wall
were found to be within 0.002 in. (0.051 mm) on either side of the average of each set. The
first buckling waves appeared at those locations where the shell wall was the thinnest. This
was verified in experiments in which the buckling system was shut down at the first signs
of buckling.
In these experiments the test wall was considered buckled according to the displacement
data when the buckling waveform was present. The displacements were so small that a visual
inspection did not show signs of buckling. Those specimens in which a definite buckling
waveform could be observed were subjected to continued loading to bring out the bifurcation
deformation more clearly (see Figure 16.76). The axial wave in the buckled shell was always
a half-wave, while the number of circumferential waves depended upon loading path, ge-
ometry and end conditions. The probe scanned at the midlength of the shell since the axial
half-wave had the largest displacement at this point.
After the experiments had been completed, data were available in the form of load versus
displacement plots and Southwell plots. The Southwell plots of load-displacement data were
used to obtain the bifurcation loads of the corresponding "perfect" shells.
For both sets of experiments (sets A and B), first the axial tension was held constant, and
then for other specimens the reversed loading sequence was applied, in which the pressure
was held constant and the tensile stress was increased. For the constant pressure loading, the
Southwell method did not perform as well as for the constant tension loading.
The influence of the sequence of loading on the buckling loads bifurcation wave number
is small, as can be seen in Figure 16.77, presenting all the results for set A of the experiments.
For set B the results for the constant pressure loading path are even closer to those of the
constant tension path (see Figure 12a of [16.146] and Figure 3.24b of [16.147]). The ex-
perimental results in Figure 16.77 are presented in the stress space, and most results are for
constant tension tests, with only a few test results for constant pressure loading. The initial
von Mises yield surface, based on a yield stress of 18,000 psi (124.1 MPa) for the 6061-T4
material, is also plotted in the figure to provide an overall reference on how deep into the
plastic region buckling occurred.
The buckling results of set A are plotted in Figure 16.78 smoothed with the Southwell
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

method. They appear to fit well along the curve drawn (by visual fit), showing that the
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1382 Plastic Buckling Experiments

(a)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(b )

Figure 16.76 Callech plastic buckling tests on short cylindrical shells subjected to biaxial loading-
typical buckled test shells (from [16. 147). courtesy of Dr. J.J. Giezen): (a) shell no. 5,
(b) group or specimens

smoothing was effective. The perfect shell loads predicted by the Southwell plots were onl y
a few percent above the measured incipient (maxi mum) buckli ng loads.
The experimental results were compared with predictions by the shell code BOSOR 5
[2.63] and [ 16.149], which allows for both 11 defotmation and J2 incremental theory plastic
models to be used in the calculations. The comparison of numerical (BOSOR 5) and exper-
imental results (of set A) is shown in Figure 16.79. Since the loading here is nonproportional,
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Combined Loading Tests in Plastic Buckling 1383

171 00

* Constant Tension
"""""

*'
0 Constant Pressure
,.... 128. 25

,.... !:_
O' 1,....
~ ~
"
..
(....
~
85 50
o·"' 1.
::.:J
*' "'"""
~ :: ~

s
D'
t ..... c
~ 42. 75

Initial von Mises Yield Surface


[Yield Srress = 18000 psi (124.1 MPa)] ,...
i
HOOf' STRESS <PSl>
'
-197. 0 -157. 6 -118 2 -78. 8 -39 4
Hoop Str-ess <Mpo)

Figure 16.77 Caltech plastic buckling tests on short cylindrical shells subjected to biaxial loading-
experimental results in stress space for set A (from [ 16.146])

the difference between deformation and incremental theory is significant. Deformation theory
(the left curve in Figure 16.79), although path dependent, captures (though with exaggera-

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
tion) the weakening behavior of the material with increasing axial load over the entire range,
as observed in the experiment. Incremental theory, on the other hand, appears to display a
trend towards the stiffening character observed for elastic shells (axial tension on elastic
shells increases resistance to buckling under external pressure).
Note that when one checks (in the buckling calculations with BOSOR 5) the prebuckling
strains corresponding to the 12 incremental results in Figure 16.78, one finds large strains in

171.00

.
\.
22. . .

2.... I '
128 25 ,.... • *
~ ,.... '
~~
.
~
1Ali!li!8 •
~ 85 so E
~
12li!0B

~ lBmlll •
i3
<-- ~ .... t ''=Buckling pressure using Southwell Method

::I~-~---~-------------+-...1-S---- •
42. 75 • =Incipient buckling pressure (rna."<. recorded)

0 2··: . . .
~

EXTE~AL. PRESSURE O'S D

0 1. 56 3. 12 4.68 6. 24 7 80
External Pressure <M~o)

Figure 16.78 Caltech plastic buckling tests on short cylindrical shells subjected to biaxial loading-
Southwell plot smoothing of experimental results for set A (from [16.146] and [16.147])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1384 Plastic Buckling Experiments

2-49Ba
\
-1 [80SOR5l l NCR. THEORY CGl

22BSB

2001119

18000

16999

~ 14191

li 12809
1:!
t;
"j 109011

~ ....
.....
.....
2BIB

..!l .lll .m ! ~
.
I!
EXT:ERNAL PRESSURE <PSI>

Figure 16.79 Caltech plastic buckling tests on short cylindrical shells subjected to biaxial loading-
comparison of numerical (BOSOR 5) and experimental results (from [16.146])

excess of 2 percent at tensile stresses above 10,000 psi. For example, at 10,000 psi tensile
stress the maximum circumferential plastic strain is 1.79 percent and the maximum merid-
ional plastic strain is 2.12 percent. At 12,000 psi tensile stress the maximum circumferential
plastic strain is already 2.73 percent and the maximum meridional plastic strain is 3.66
percent. At higher tensile stresses the strains increase rapidly. However, it is obvious from
Figure 16.75 that such large strains indicate material failure. This means that these buckling
prediction are invalid, since material failure will preempt bifurcation here. The large calcu-
lated strains, which will cause material failure before buckling can take place, occur only
along the incremental theory curve. This curve is therefore cut off in Figure 16.79 at a tensile
stress of 10,000 psi because it would be fictitious beyond that stress level. If it were continued
beyond material failure (which would be fictitious), it would show an opposite trend to that
of the experimental results, a curving to the right of the figure with increasing tension,
whereas the experimental points clearly curve to the left. In other words, incremental theory
does not predict lower buckling pressures with increasing tensile loads, as observed exper-
imentally.
One should also note that the curve f<2!_ BOSOR 5 incremental theory in Figure 16.79
employs the instantaneous shear modulus G [see Eq. (16.2)] in the bifurcation analysis. The
corresponding pure incremental theory curve, calculated with the usual elastic shear modulus
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

G, would appear to the right of that shown and bend farther to the right. It has not been
plotted here, since again material failure would occur earlier, making the pure incremental
theory curve fictitious.
The main aim of this study was to present a series of experiments with well-defined
loading and boundary conditions, as well as measured material properties and initial imper-
fections (recorded and reduced in the manner discussed in Chapter 10), which could be used
to evaluate different plasticity models. From the results shown in Figure 16.79 (and similar
ones for set B of the experiments) it can be concluded that the commonly used plastic models
(J 2 incremental and J 2 deformation theories) do not model correctly the type of nonpropor-
tional biaxial loading studied here. The 12 deformation theory at least shows the correct
trend, whereas J2 the incremental theory leans away from it. Furthermore, for relatively
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Combined Loading Tests in Plastic Buckling 1385

moderate axial stress (> I0,000 psi), large plastic deformations and material failure will occur
prior to buckling, invalidating bucking predictions of the incremental theory for large tensile
loads. The results provide another confirmat ion of the plastic buckling paradox. discussed
in Subsection I6. I .3.
Calcu lations were also carried out with J 2 corner theory [ I 6.33], yielding better correlation
with the experimental results (see [16.1 47)). But though one could choose the corner theory
to reasonably fit the experimental results, o priori prediction was not possible!
Whereas the primary purpose of the fi rst Caltech experimental program was the evaluation
of plastici ty models, the second Caltech experimental program. on the collapse of long. thick-
walled cylindrical shells under combined loading of external pressure and axial tension
[I 6. 148], was motivated by the practical problem of pipe laying in deep waters.
A series of collapse tests was conducted on thick-walled. small-diameter tubes of two
ditlerent materials, subjected to combined external pressure and axial tension. The (Rit)
range chosen was 5- 20. the range of primary interest to the oil industry, and careful mea-
surements of geometrical and material parameters preceded the tests.
The experiments were conducted using a speciall y designed test setup. The test facil ity is
shown in Figures 16.80a and 16.80b. The test chamber consists of a thick cyl inder with
scaling end caps at both ends. The pressure chamber had an inside diameter of 2.5 in. (6.4
em) and an internal length of 42 in. (I 07 em). The setup had a pressure capacity of I 0,000
psi (69 MPa) and an axial load limit of 20,000 Jb (89,000 N). A servo-controlled hydraulic
system was used to apply axial load to the specimen. The chamber was pressurized with
water, using a positive displacement pump that applied lateral pressure on the tube outer
diameter (see Figure J6.80b). A data acquisition system was used to monitor and record
axial tension, pressure and ax ial elongation during the tests.
Thick-walled tubes with ini tial imperfections show a load-deflection response with a load
maximum. Applying axial tension was shown in !his study to maintain the limit- load-type
response. However, the presence of axial tension makes the load-deflection response sofler
due to material inelastic behavior. The limit pressure and tension loads at which the tube

(a) (b)

Figure 16.80 Callech collapse tests on long cylindrical shells subjected to biaxial loading-test setup
--`,`,`````,`,````,,``,``,,`-`-

(from [16.148]. cou11esy of Dr. R. Madhavan): (a) view of test rig in operation, (b)
schematic diagra.m of LC!SL setup under combined tension and external pressure
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1386 Plastic Buckling Experiments

loses its pressure-carrying capacity were assumed to be the collapse strength of the tube
under combined loading.
In the load-controlled experimental setup (see Figure 16.80b) onset of catastrophic failure
at the limit loads created a change of volume occupied by the pressurizing fluid, water. Since
the pressure chamber was relatively stiff and water is practically incompressible in the pres-
sure range of interest, this change in volume appears as a sharp drop in pressure. Hence, the
maximum pressure and tension loads measured prior to the sudden pressure drop, indicating
collapse failure, were reasonable approximations for the collapse strength of the tube under
combined loading.
In the experimental setup (see Figure 16.80a and b) the pressure (test) chamber stands
vertically. A hydraulic cylinder was employed to pull the specimen from the top. A 62-Series
Moog servo-valve system, located at the side of the hydraulic actuator, was used to manip-
ulate the axial loading. A diaphragm-type pump, driven by high-pressure air, was employed
to pressurize the test chamber; and a synthetic fluid-water mixture was used as pressurizing
fluid to minimize any rusting.
Commercially available 304 stainless steel tubing and 6061-0 aluminum tubes of nominal
diameter 1.25 in. (32 mm) and 1.375 in. (35 mm) were used for preparing the specimens.
Each test specimen was prepared by bonding end plugs to an appropriate length of tubing
with a high-strength epoxy. This led to a test section of approximately 20 times the diameter
of the tube between the end plugs. A schematic of a test specimen is shown in Figure 16.81.
In the test setup one end of the specimen is rigidly attached to the bottom end cap, with its
threaded end plug, and the sealing plug of the other end projects through the top end cap
as shown in Figure 16.80b. The tensile loads are applied to this sealing end plug via a load
cell.
The thickness and outer diameter of each specimen were measured before the end plugs
were bonded. This was followed by out-of-roundness measurements on each specimen. An
experimental setup was developed to measure these initial imperfections with sufficient ac-
curacy, using a concept generated by Arbocz and Babcock (see Chapter 10, Section 10.3,
and [9.194], [9.221] and [10.5]). The specimen was mounted between two centers on a lathe
and a high-resolution displacement transducer (LVDT) was used for the imperfection mea-
surements (see Figure 16.82). The spring-loaded probe of the transducer touched the tube
surface as the tube was rotated about an axis between the two centers. The setup could
accommodate specimens of up to 45 in. (114 em) in length, and the resolution of displace-
ment measurement was within 0.00003 in. (0.001 mm). The measured radial displacement
data 8(6) from 140 points around the circumference of the tube, was then fitted with a Fourier
series and reduced in the manner discussed in detail in Chapter 10. These measurements
were repeated at 5-10 axial locations.
No effort was made to obtain a three-dimensional mapping of deviations from perfect
geometry or to model the longitudinal and circumferential imperfections. The simplified two-
dimensional formulation presented in this study models a tube with initial imperfections that
are uniform along the length of the tube. Hence, measuring the profile of the tube at several
axial locations and choosing the largest imperfection allows a reasonable lower bound for
the predicted collapse strength to be obtained, though a rather conservative one.

Figure 16.81 Caltech collapse tests on long cylindrical shells subjected to biaxial loading-collapse
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

test specimen, schematic (from [16.148))


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Combined Loading Tests in Plastic Buckling 1387

Figure 16.82 Cahech collapse 1es1s on long cylindrical shells subjcc1ed 10 biaxial loading-experi-
mental facilily 10 measure inilial ovalil)' (from [ I 6.1481. courtesy of Dr. R. Madhavan)

T he material properties of the specimen were obtained from tensile specimens cut along
the axis of the tubes. The 304 stai nless tubes showed smooth stress-strain behavior. and it
could be fitted with a three-parameter Ramberg-Osgood (R-0) relation for strains up to 3
percent. Figure 16.83 shows the typical stress-strain behavior of 304 stainless steel and its
R-0 fit. Si milar behavior and curves were obtained for the 606 1-0 aluminum tubes.
The tubes tested showed yield anisotropies due to the manufacturing processes. These
appear as different yield behavior in the axial and circumferenti al directions. T hese initial
yield anisotropies are characterized by the parameters and S~ where S, is the ratio of the s.
initial yield stress in the hoop direction to that in the axial direction (=u0 of cr0..J. and S, is
the ratio of 1he in it ial yield stress in the radial direction 10 that in the axial direction (=u.,,.
I cruJ The special case of S,! = S, = S = I represents an initially isotropic material. Stress-
strain curves in both axia l and hoop directions were obtained from specimens cut from the
tubes to measure these ani sotropies (see [ 16. I 50]). which were then introduced into the
calculations accordi ng to a method developed in r16. 15 1).
Si nce the collapse pressure and tension loads are characterized by limit loads that occur
after considerable plastic defotmation of the tube cross-section, the loadi ng path is likel y to
affect the lim!t loads. Th is was investigated by conducti ng collapse tests involving two
different loadi ng paths. The loading path in which the specimen is first subjected to a given
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

30 - ..J "'
"- ~:::;,-;::::::,:::.:::.:.:-:-t

: . ..
500

t ..--.:----:....... ..... .........................{ a


~ r--~ E,~OI ~ental w~e
! •• -1
T :::
I
;
~
~ 00

L
Rambero·Osgood 111,
30 Modified Iii ,
E .. 29000 ksl ( 199955 Mpa)
+ 20.0

20 Cly • 44 S ksi (307 Upa), n • 13.4

,, -
I
10::1

0 ·-i--
,.
Figure 16.83 Cahech collapse tests on long cylindrical shells subjec1ed to biaxial toading-
experimemal and fitted Stress-Strain curves for 304 stainless steel (from f J6. t48]). Note
that for large strains the modified f11 is much closer to 1he experimental curve !han the
R-0 fi1
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1388 Plastic Buckling Experiments

axial tension and then pressurized until collapse is denoted by T ---+ P. The test procedure
was such that further loading was stopped when collapse was detected through a sudden
drop of pressure. This left a localized collapse pattern, which extended over a length of three
to five diameters.
The other loading path consisted of applying a certain external pressure on the specimen
and then increasing the axial tension until collapse. This load path is denoted by P ---+ T.
For tests involving loading path P---+ T, an accumulator was used to keep the pressure steady
while increasing the axial tension. Occunence of collapse was indicated by a sudden drop
in pressure. This was followed by the propagation of the collapse extending over the entire
length of the specimen, since the accumulator partially restored the momentary drop in
pressure.
Some tests using initially ovalized tubes to study the sensitivity of collapse strength to
initial ovality were also conducted. Initial ovality was induced by crushing the specimens
between two parallel plates of a hydraulic press. The induced ovality on test specimens,
(Dmax - Dmin)/2D,, varied in the range of 0.0005 to 0.040.
The axial elongation of the specimens recorded during the collapse tests gave a measure
of the axial strain at collapse, which was used as another check on the predictive capability
of the model employed in the calculations of the study.
Additional details of test procedures, initial imperfection measurements and material prop-
erties are presented in [16.150].
The tests included 64 stainless steel and 6 aluminum tubes with 6 different (RI t) ratios
(Rit = 5-20), and the experimental collapse strengths were compared with predictions by
a two-dimensional elasto-plastic J 2 flow rule (incremental theory) model, taking the initial
imperfections into account. The predicted results were found to be in reasonably good agree-
ment with the experiments (see [16.148] and [16.150] for details).
A parametric study was performed to check the sensitivity of the collapse load to various
parameters. The main conclusions of the combined experimental and analytical study were
the following:
1. Axial tension was found to reduce the collapse pressure. The extent of this reduction,
and hence the nature of the collapse pressure-tension profile, depend strongly on the
stress-strain behavior of the tube material.
2. For tubes of a (Rit) range 5-20 considered in this study, collapse occurs after the tube
exhibits considerable plastic deformation. Results show that the yield stress and the
postyield shape of the stress-strain curve are the most important material parameters that
affect the collapse envelope.
3. The study conducted on tubes of low initial ovalities and two different loading paths
(T---+ P and P---+ T) shows that the loading sequence has very little effect on the tension-
pressure collapse envelope.
4. Initial ovality was shown to be a very important geometric parameter that reduces the
collapse strength at all axial loads.
5. The two-dimensional elasto-plastic model employed predicts the experimental results
well, provided accurate measurements of the tube geometry, initial imperfections, ma-
terial stress-strain behavior and anisotropy in yielding are obtained and used.
It appears that the limit-load-type response here, in contrast to the bifurcation response of
the short shells in the first Caltech experimental program, allowed the J2 incremental theory
to perform satisfactorily. Thus, no plastic buckling paradox arose in this case.
One important point relating to the Ramberg-Osgood approximation of stress-strain curves
ought to be stressed here. It is probably most conveniently demonstrated with an example
presented in Figure 16.84, which shows experimental results for the tension pressure collapse
envelope of a tube of (R It) = 12.3, correlated with numerical predictions. These results are
for a loading path of T---+ P. In spite of the scatter in experimental points, especially due to
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Southwell's Method in the Plastic Range 1389

o.ao T ----

0. 6 0 -----..___

-----------
• 0
"-
0
0.40
'
0

..
"- • Experiments

- Theory, R-0 fit •


0.20 t ··· Theory, Modified fit

No.G3 Rm/lo12.3,5=1,T-P

0.00 -+----+----

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
0.00 0.20 0.40 0.60 0.80 1.00 1.20

T cIT~

Figure 16.84 Caltech collapse tests on long cylindrical shells subjected to biaxial loading-
experimental tension-pressure collapse envelope, of a shell with Rlt = 12.3 and T--> P
loading, compared with predictions based on R-0 fit and modified fit representations
for the stress-strain behavior of Figure 16.83 (from [16.148])

slight variations in material and geometric parameters among specimens, a trend of the slope
of the pressure-tension collapse envelope diminishing at higher axial loads (when the axial
collapse tension is above 0.9 the yield tension) can be discerned. Numerical predictions
based on R-0 (Ramberg-Osgood) approximation of the stress-strain behavior did not show
such a trend in the pressure-tension collapse envelope. From Figure 16.83, it can be seen
that for high strains (especially above 2-3 percent level) a modified fit rather than the R-0
fit gives a better representation for the stress-strain behavior of 304 stainless steels.
In this modified fit (see [16.151] and [16.152]) the R-0 representation is used up to 1.5
percent strain. Above this strain level the stress-strain curve is approximated by a straight
line, which is tangential to the R-0 fit, which ends at a strain of 1.5 percent. Returning to
Figure 16.84, one notes that the collapse predictions based on the modified fit better repro-
duce the trend of the experimental collapse envelope.
This example illustrates the sensitivity of the collapse pressure-tension interaction curve
to the shape of the stress-strain curve of the material and stresses the care needed in its
approximation. The experimenter should recall that the very useful R-0 fit may have limi-
tations for large strains, and when large strains are involved, some modification to the R-0
fit, like that used here, is essential.

16.4 Southwell's Method in the Plastic Range


The Southwell plot, originally proposed in 1932 by Southwell [4.12] for determination of
the theoretical buckling load P£ of a perfect column from experiments on real imperfect
columns, has been widely used and extended to other structures (see for example Section
4.5 of Chapter 4 and Section 8.3 of Chapter 8). The Southwell method and its extensions
are based on small-deflection theory and assume elastic material behavior. Hence, its appli-
cation to plastic buckling appears to be ruled out a priori. Southwell observed that when he
applied the method to the classical strut tests of von Karman [4.4] the results were excellent
for slender struts but failed, as he expected, for medium and thick struts since practically
every observation there related to deflections for which the material had ceased to be elastic.
Tuckerman [4.30], when reviewing extensions of the Southwell plot, stated clearly that
"None of these methods can be expected to give straight line graphs for high loads at which
plastic yielding has become appreciable, and for which the theory is no longer applicable."
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1390 Plastic Buckling Experiments

However, the beauty and simplicity of the Southwell method, and in particular its useful-
ness in smoothing experimental data in parametric studies, by removing most of the imper-
fection effects (geometric imperfections, load eccentricities and minor variations in stress
distribution) related to the specific specimens, as pointed out in 1934 by Bridget et a!. [4.31 ],
has tempted the investigators to apply the method even well into the plastic range. And
surprisingly, sometimes it worked extremely well (see for example [16.153] or [16.154]).
In 1948, Wang [16.15] showed that the Southwell method can be extended into theine-
lastic region, provided that E,, the Engesser-von Karman double modulus, is approximately
constant. Wang applied the Southwell plot to von Karman's 1910 short columns, to Gerard's
1947 Republic Aviation Corporation short columns (detailed in [6.15]), and then applied the
Southwell plot, as modified by Lundquist [4.25], to Horsfall and Sandorff's 1946 short
column (see [16.155]), and showed that the test data lay closely to straight lines and that
the Southwell plot predicted buckling loads close to the double-modulus theory buckling
load P,. It is astonishing that this important 1948 paper was completely forgotten, not having
been mentioned in any of the reviews or scores of papers on applications of the Southwell
plot published in the 40 years that passed till Singer revived it in 1988 [16.156], when he
discovered it by chance (after having himself finished a similar derivation independently and
having applied the method successfully, and with great joy, to all von Karman's short col-
umns).
Even C. Massey's 1964 paper [4.41], in which a modification of the Southwell plot for
lateral instability of I-beams was developed and shown to be approximately applicable also
in the inelastic region, does not mention Wang's work. Nor is it mentioned in Newman's
1973 paper [16.153], which successfully employs the Southwell plot to predict plastic buck-
ling of pressurized tubular columns, though without justification; nor in Sobel's 1983 paper
[16.154], which applies the Southwell method to plastic buckling loads of elbows.
This important theoretical justification for the use of Southwell's method in the plastic
region, which puts the later (and future) applications on firmer ground, is therefore derived
in detail. Singer's derivation [16.156], which differs slightly from Wang's, follows.

16.4. 1 Extension of Southwell's Method to Inelastic Columns


If one reflects and reconsiders the derivation of the Southwell method, one realizes that a
certain class of plastic-buckling problems may also theoretically qualify. In other words, if
buckling occurs well into the plastic region and the measurements are practically all taken
when the dominant part of the structure is plastic, and in addition the stress-strain behavior
is essentially linear strain-hardening, the conditions resemble elastic buckling and the South-
well plot should yield good and consistent results. It will, however, not predict the classical
elastic buckling load of the perfect structure, but its plastic buckling load, based on the linear
strain-hardening material properties.
A simply supported column is reconsidered and its buckling in the plastic regime is an-
alyzed. The shape of the stress-strain curve of the material is assumed to be such that it can
be approximately replaced by two straight lines (see Figure 16.85)-one elastic (up to the
yield stress a) with slope E (Young's modulus), and one plastic (from u, onwards) with
slope E, (the tangent modulus).
For a perfect column, the equation of equilibrium when u > u, is now, according to the
Engesser-Karman double (or reduced) modulus theory (see for example [6.46, pp. 12-13]),

dc.yo (El ,+ Er2=-y


I) P (16.17)
dx-
If one writes
(16.18)
or --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Southwell's Method in the Plastic Range 1391

Et (loading
fibers)

E (unloading
fibers)

Figure 16.85 Idealized uniaxial stress-strain diagram: linear strain hardening.

(16.18A)

Eq. (16.17) becomes


E,ly," + Py = 0 (16.19)
For an imperfect column, with initial geometric imperfections y 0 (x), and if y(x) is the addi-
tional deflection (see Figure 16.86), Eq. ( 16.19) is replaced by
E,Iy,,x + P(y + Yo) = 0 (16.20)
Following the usual derivation of the Southwell plot by assuming

y 0 (x) = ~
n~l
w 0 , sin (nTTx)
L (16.21)
and y(x) = ~
n~l
w, sin (nTTx)
L

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
leads to

y(x) = 2:
X
W ,
0
[ 2p r
np - 1] -j
sin ( nLTTX ) (16.22)
n-I

where

(16.23)

the critical load according to the Engesser-Karman reduced (or double) modulus theory.
As the critical load is approached, when P is a fairly considerable fraction of P,, the
central additional deflection,

(16.24)

As P --+ P,, the imperfection component that represents the buckling mode is the one that
is primarily magnified. Hence, one can write
(16.25)
and Eq. (16.24) can be rearranged as

8 = P, (~)- W0 (16.26)

One may note that the derivation is somewhat similar to that of Southwell for elastic
columns, presented in Chapter 4.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1392 Plastic Buckling Experiments

The inverse slope of the plot of versus 8/ P versus 8, the Southwell plot, therefore yields
the buckling load Pr in the inelastic range of the corresponding perfect column, according
to the Engesser-Karman reduced modulus theory. Note that P, is the bifurcation critical load
in the inelastic range, and that as in the elastic range the Southwell plot yields the Euler
load, which is the elastic bifurcation load, it yields Pr in the plastic range. One should
remember that the reduced modulus theory assumes strain reversal (see for example Sub-
section 2.1.13c of Chapter 2 and Subsection 16.1.2 here).
It is important to reiterate that the Engesser-Karman reduced modulus theory is the gen-
eralization for the inelastic case of the reasoning used in elastic bifurcation theory (see
Subsection 16.1.2 and [16.14]). The Southwell method for plastic buckling should therefore
predict the reduced modulus (or double modulus) critical load P, for the corresponding
perfect column, though imperfect columns usually buckle nearer to the tangent modulus
critical load Pf' as pointed out by Shanley [2.11] and generally accepted today (see for
example [6.46] and [16.15] and [16.157]). One may note that also by using a Shanley model,
plastic buckling according to the Engesser-Karman reduced modulus theory implies a neutral
equilibrium postbuckling path (see [16.44]), as in the case of the elastic Euler column.
In the above analysis linear strain hardening has been stipulated to make the derivation a
rigorous generalization of the Southwell':; analysis. This restriction can, however, be relaxed,
and the derivation also applies to the more general case of a strain-hardening material, whose
stress-strain curve can be approximated by linear portions at the points of buckling.
As mentioned earlier, for experimental verification Wang applied the extended Southwell
method to the short columns of von Karman [4.4], to the short column of Horsfall and
Sandorff [16.155] and to Gerard's tests [16.15]. The application of von Karman's columns
has also been independently reconfirmed by Singer with practically identical results. Figure
16.87 shows the application of the extended Southwell method to von Karman's mild-steel
short columns. Similar excellent fits were shown in [ 16.15] for the remaining two short
columns of von Karman and for the 2024-T aluminum alloy columns of Gerard and the
column of Horsfall and Sandorff.
The details of these applications are presented in Table 1 of [16.15] (reproduced also in
[16.156]) and show that the extended inelastic Southwell predictions from the tests are indeed
very close to the buckling stress calculated with the Engesser-Karman double (or reduced)
modulus theory for perfect columns, ar. Except in one case, the Southwell predictions are
within ± 5 percent of ar, and the predicted asouthw for the corresponding perfect column is
always above or equal to the experimental buckling stress of the imperfect column, +0
percent to +9 percent above.
In [16.156] the extension of the Southwell method to plastic buckling was broadened to
other classes of columns, such as columns on elastic foundations or columns with different
combinations of end conditions and with varying flexibility.
The main conclusion of this discussion is that the Southwell plot for columns is as valid
in the inelastic region, provided the material is strain-hardening, as it is for elastic bifurcation

:r
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

I X

I •
L

I
_j__

Figure 16.86 Deflections of an imperfect column


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Southwell's Method in the Plastic Range 1393

N"
E
....E
"'
ZIO
E
E
·=
"'Q
" 5 t--t~-A-;,ft-t't---,1;--;(t
coolb

Defleetion 8 in mm.

Figure 16.87 Southwell plot of von Karman's test results for short columns (from [16.15])

buckling, to which it has been restricted since conceived by Southwell in 1932. Further
theoretical justifications are needed for reliable extension of the plastic validity to other
structures such as plates or shells. However, in practice such extensions can be tentatively
attempted, as they are on as firm a ground as many of the extensions in the elastic region
that relied only on Southwell's original work.

16.4.2 Southwell Plots in Plastic Buckling of Shells


The extension of the Southwell method to shells was thoroughly discussed in Section 9.6 of
Chapter 9, where it was shown that the method has a wide range of applicability in shell
buckling. For the types of loading and shells whose postbuckling behavior is not far from
neutrally stable, the method has been shown to work very well. Even when the postbuckling
behavior is unstable, it was shown that the Southwell plots can be resorted to with confidence,
provided the unstable postbuckling behavior is relatively mild, as for example in closely
stiffened shells.
Most of that discussion also applies to plastic buckling, though as yet without theoretical
justification, and the Southwell method has been successfully employed in many cases of
plastic buckling of shells (see for example [9.372], [16.153] and [16.154]). The important
use of Southwell plots for smoothing of data in parametric studies has also been effectively
applied to plastic buckling (see for example [16.146]).
One of the early inelastic uses of the Southwell method was that of Newman for his plastic
column buckling tests of internally pressurized tubes, performed at the Westinghouse Bettis
Atomic Power Laboratory at the beginning of the seventies [16.153]. The experiment dealt
with inelastic column buckling of annealed aluminum tubes in the presence of a stabilizing
internal pressure and can therefore be considered either column or shell tests.
The test setup (see Figure 16.88a) consisted of a tubular specimen with hemispherical end
fittings (see Figure 16.88b) located in spherical cups (see Figure 16.88c), through which
axial thrust from a universal testing machine was applied. Pressurization was achieved
through a hose attached to the bottom end fitting. The effective length of the specimens
(influenced by the rigidity of the end fittings and their rolling kinematics in the end cups)
and the restraint of the pressure hose were determined by special elastic buckling tests of
three longer, similar specimens.
The 14.55-in. (-370-mm) long test specimens were fabricated from commercially avail-
able 3003-H14 aluminum tubing, of 0.5-in. (12.7-mm) diameter and 0.028-in. (0.71-mm)
wall thickness and were recrystallization annealed. Stress-strain data were obtained from
tensile specimens cut from the same stock as the buckling specimens.
The internal pressure was generated by a pressure gage tester with a screw-driven hydraulic
ram and measured by a calibrated 5000 psi precision pressure gage. The pressurization hose
was hung from very flexible springs to minimize its restraint.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1394 Plastic Buckling Experiments

Jtn:LO. )( 0.028" WALl. )00l ·~l ol ALUMrNIJM


TU8!N G • 'NN'£ AUO B f i"OR[ W(LOING
0 1$ OIA' SPH(AIC AL
$VRfAC( • 16 fl'MS
W£l0·CAC• FNO~
• ill , , ·-

)1 4 MOfA Wl l OAf!t( J
ALV I.O IHU M IIAH
1/ll~ PIP( TII O
0.!10 O([P • ( ACH lNO

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
- - - !4 .$6 ~ NON tNAl - - - - - 1

(b)
BUCKLING TEST SPECI MEN

(a) (c)

Figure 16.88 Newman's West inghouse plastic column buckling tests (from [16. 153)): (a) buckling test
assembly, (b) test specimen. (c) spherical cup

The lateral deflections of the specimens were measured by four dial gages (sec Figure
16.88a), and the buckli ng loads were obtained from these deflections by application of the
Southwell method. After the preliminary elastic-buck ling tests in dry and greased loading
cups. it was decided to test the inelastic specimens in greased cups.
The Southwell plots were first applied in the usual manner (as detai led in Section 4.5 of
Chapter 4) to the clastic-buckling tesL~. which preceded the main inelastic-buckling tests.
Then they were used to dete1mi ne the plastic buckling loads of the pressurized tubes, by
replacing Young's modulus E wi th the tangent modulus £, in the Southwell formulation, Eq.
(4. 11) of Chapter 4. The expression used for the Southwell plots was therefore
5 = P,(Dip) - w, ( 16.27)
where

p = r.ZEJ
, u ( 16.28)

is the tangent modulus cri tical load, Sa measured central deflection and w, the initial central
deflection at zero load. The interchange of P" with P, implied the valid ity of Shanley's
tangen t modulus theory for the Southwell method, whereas it has been shown in the previous
subsection that the inelastic Southwell plot predicts the Engesser-Karm;1n reduced modulus
critical load (being the inelast ic generalization of the reasoning used in elastic bifurcation
theory). Newman's assumption that his Southwell plots will predict the tangent mod ulus
critical load P, was therefore inaccurate, an inaccuracy that in practice, however, was prob-
ably absorbed in the experimental scatter.
Thi rty-nine specimens were tested in the plastic range, II with no imcrnal pressure and
28 with an internal pressure at buck ling of 800-1700 psi. Four Southwell plots, typical of
these tests, are shown in Figure 16.89. They appear very consistent. Newman pointed out
that since the slope of the Southwell plots was obtained fro m the straight lines fitted to the
data, which usually did not extend up to actual buckling. satisfactory results cou ld often be
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Southwell's Method in the Plastic Range 1395

SPECIMEN 4014 539 SPECIME~. IA14.538


TEST 8 PSI/LB THRUST TEST 1250 PSI
Tc; 146 LB T0 ; 187 LB

1.0 I .50 r

tn
::>
0::
:r: 0
0
z
::>
so
!' 40
i
~
I /'
/

~ ~ .60
z<n
~ ~
;' 30 ~
I ,/
X

g~ 40
/
X
20
r /
1t
/
X
/X
X
10 I1 ,
' ;x

X __ __L _
0 _. _J____
0 X _J____~

0 40 80 120 160 0 20 40 60 80
DEFLECTION- i0- 3 INCH~S DEFLECTION-I0- 3 1NCHES

25 :o ~ /
/
~ ~ 20 /X
o:: ::> X 08 l' ,/
~it /
~ ~
~ u~
! 5
X/
X 06 ,I
" f /;;::'~';,'~::,~"
f- '
~ ~ ~
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

10 /
u: '? I / SPECIMEN 2814.531
i':5 C::: • x TEST P;O
.OSj / Tc"327LB 002 /[;t',' Tc;332LB

0
t
L___J__ _j__-'-~ L___J__ _.!__..l--~

0 20 40 60 80 0 10 20 30 40
DE FLECTION -10- 3 INCHES DEFLECTION-I0-3 INCHES

Figure 16,89 Newman's Westinghouse plastic column buckling tests-Southwell plots for four typical
buckling tests, with different internal pressures, showing the least squares-fitted straight
lines, from which the critical loads were determined (from [16.153])

obtained without attaining buckling, i.e. from "nonbuckling tests." This is in accordance
with the experience on elastic columns (see Section 4.6. in Chapter 4). For other structures,
like plates or shells, it is known that one has usually to approach the buckling load quite
closely, often making nondestructive use of the Southwell method fairly difficult.
A plot of buckling loads versus internal pressure at buckling, exhibited, however, consid-
erable scatter. Newman attributed the scatter to variability of material properties, probably
due to some inadequacy in the annealing process, and to effects of plastic deformation at
the ends of the specimens. A comparison with predictions using deformation or incremental
theory showed a trend of better agreement with the predictions of incremental theory than
with those of deformation theory. The large scatter in the experimental results, however, casts
some doubt on this trend, which is contrary to observations in practically all other experi-
mental studies (as pointed out in Subsection 6.1.3).
A decade later, Sobel [16.154] applied the Southwell method to plastic bending and buck-
ling tests of two 304 stainless steel piping structures, performed earlier by Peters at the
Westinghouse Advanced Reactor Division. Each of the nominally identical structures con-
sisted of a 90° seamless elbow and two straight tangent pipes. One end of the structure was
fixed, and its other end was subjected to an in-plane closing bending moment M (see Figure
16.90). Two types of tests were carried out: cold tests, at room temperature, and hot tests,
at 1100oF (593°C). For this case, the moment M increases monotonically with the rotation
11 until M reaches the nonlinear buckling (or collapse) value at the maximum point of each
M-11 curve (see Figure 16.91). As pointed out by Sobel, the elbow problem is a limit point
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1396 Plastic Buckling Experiments

Extrados

- -_:::?!
I R 1
I

Figure 16.90 Sobel's application of the Southwell method to plastic buckling of elbows-the West-
inghouse piping structure and notation (from [16.154])

1.0

.9

.8

.1
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

.6

....
l .5
::E

.4

.J

.2

.1

{j (dogrHS) Or y(in.)

Figure 16.91 Sobel's application of the Southwell method to plastic buckling of elbows-moment versus
rotation e and versus vertical displacement v (from [ 16.154])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Southwell's Method in the Plastic Range 1397

problem and not a bifurcation buckling problem. This is evident in the moment versus
rotation or vertical displacement v curves for a typical hot test, where Mh is the moment at
buckling (Figure 16.91). Similar curves were obtained when the moment was plotted versus
hoop strains.
In 1955, in their application of the Southwell method to stiffened cylindrical shells sub-
jected to external pressure, Galletly and Reynolds [9.214] suggested using only the nonlinear
portion of the total hoop strain in the Southwell plot in order to obtain straight lines from
the test data. This suggestion has been accepted by many investigators since, with good
results. For example, in 1977 R.F. Jones, Costello and Reynolds [4.49] reiterated that "It
should be noted that the deflection to be used in the Southwell method is the nonlinear
component of the total deflection."
For the elbow, the nonlinear component of the deformation, a"', can be computed from a
load-total deformation curve (M-a) according to
for each M, (16.29)
where a, is the extrapolated linear component of the deformation, obtained from the contin-
uation of the initial linear portion of the M-a curve. a can also represent a strain £. One
should note, however, that the linear and nonlinear components of the total strain £, £ and 1

£"', are not equal to the elastic and plastic components £ and £1' of£. This is so because at
1

an arbitrary point the elastic component £en,e of the effective strain ce~r at that point is linearly
related to the effective stress uctP since cerr.c = Ue~rl E. But since O'e~r does not vary linearly
with M at the high load levels, because of both plasticity and geometric nonlinearity, cctr.e
does not increase linearly with M in the nonlinear region of the M - ccn curve, whereas £ct1 _,
varies linearly with M (by definition). Hence, c,, =!= c)
The nonlinear component of the deformation (or strain) a,, generally consists of a geo-
metric part and a material part. Strictly speaking, only the geometric part should be used in
the Southwell plot, but the test measurements provide only the total nonlinear component
a"'. To estimate the relative error introduced by the use of the total a"' instead of only its
geometric part, predictions corresponding to the two possibilities were made from the results
of a finite-element analysis [16.158] for a cold test of the elbow. The predictions differed
by less than 1 percent, indicating that the measured nonlinear component a"' can be safely
used in the Southwell plots.
Figure 16.92 shows two plots of the dimensionless moment versus the nonlinear compo-
nent of deformation, based on the rotation 8 measured in a hot test of the elbow. The
dimensionless variables M* = M I Mh and a;;, = 8"'/ 8"'." express M and a"' as a fraction of
their respective values at buckling M" and e,,.JJ. The circular points in the figures are the
experimental values, identical in both plots (a) and (b). The curves present a rectangular
hyperbolic fit
(16.30)
which was made to pass through the points indicated by the arrows (thus determining the
constants a 1, a2 , and a 3 ). In Figure 16.92a one of the fitting points was selected to be at the
origin, resulting in a 3 = 0. In this case Eq. (16.30) becomes
(16.31)
This equation is identical in form to the extended Southwell formulation, Eq. (16.26). In
Figure 16.92b a somewhat better fit is obtained by not forcing the hyperbola to go through
the origin, i.e., if the constant term a 3 is nonzero but quite small. In either case, Figure
16.92a and b both show that the experimental points exhibit hyperbolic behavior, especially
as M __, Mh. One should remember (see Subsection 4.5.1 of Chapter 4) that the rectangular
hyperbolic nature of an appropriate load-deformation curve determined the basic applicability
of Southwell's method. Hence, Sobel argued that the method is valid for any structure whose
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

load versus nonlinear component of deformation behavior is monotonic and increasingly


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1398 Plastic Buckling Experiments

.B

.6
~

"
i,,
~
.4

.2
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

.2 .4 .6 .B 1.0

1.0

.B

.6

;); "
i
j;
.4

.2

(b) 0
.2 .4 .6 .B 1.0

Figure 16.92 Sobel's application of the Southwell method to plastic buckling of elbows-hyperbolic
fit through the points indicated by the arrows closely approximates the experimental
moment versus nonlinear component of rotation results (from [16.154]): (a) origin se-
lected as one of the fitting points, (b) a better fit obtained with small but nonzero a,

hyperbolic as the load approaches its buckling value, irrespective of the region it is in, elastic
or plastic. Therefore, the closeness of the hyperbolic fits to the experimental points in Figures
16.92a and b provides justification for the application of the Southwell method to elbow test
results.
Further justification for the use of Southwell plots in elbows loaded by a closing moment
(as in the Westinghouse tests) is provided by a heuristic argument, based on the load-
deformation behavior of elbows as buckling is approached. "Buckling of an elbow may be
described as the phenomenon in which a small increase in load causes a disproportional
large increase in the nonlinear deformation component o,. Thus, in the vicinity of buckling,
a barely perceptible change of moment causes a finite change in deformation (essentially
rectangular hyperbolic behavior). Consequently, the Southwell plot of o"' versus (o,J M) is
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Southwell's Method in the Plastic Range 1399

approximately linear in this vicinity, M-- M", as 8"' is essentially being plotted against itself
divided by an almost constant value M = M". The slope of this nearly linear portion of the
Southwell plot is therefore [8,/(8,,/M)] = M = M".
The same conclusion is obtained if one studies the qualitative analysis performed by Sobel
(see Appendix A of [16.154]). In that analysis the slope m of the Southwell plot of (8,,/M)
versus 8"' (here the ordinate and abscissa were temporarily interchanged for convenience) is
expanded in a Taylor series about the buckling point, which yields

_ d( 8,,/
m - dB
M) = __!___ [
M 1 _ 8,/'
M
(dd8M)
2

2 (B,,
_ 0"''b) J
h h b
Ill' !l
(16.32)
as

This equation relates the slope of the Southwell plot in the vicinity of buckling (M -- M")
to the shape (the curvature) of the corresponding M-8,, curve (for example Figure 16.92a).
Examination of Eq. (16.32) shows that

1. at (16.33)

i.e. the reciprocal of the slope of the Southwell plot at its buckling point is precisely the
buckling load M".

2. I
m = M = const. as (16.34)
b

provided that the curvature at the buckling point is small (as is the case for the elbows
tested). In other words, the upper portion of the Southwell plot, corresponding to the
immediate neighborhood of buckling in the associated moment-deformation (M-8,) curve,
provides a close approximation, Msouthw = (1 /m) = M,, to the buckling load.
3. ( 16.35)
i.e. the Southwell plot prediction is an upper bound."
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

One should, however, note that also the further justification and its qualitative analysis are
limited to the behavior of the specific type of elbow and do not apply to shells in general.
Returning now to the Southwell plot prediction for the hot elbow tests, the nonlinear
deformation component 8,, should be used, rather than the total deformation 8. To reiterate
this point, typical Southwell plots employing both 8,, and 8 are shown in Figure 16.93, based
on the measured rotation e. Similar plots, based on vertical displacement and crown hoop
strains, are presented in [16.154] (Figures 9, 10 and 11 there) and demonstrate the same
advantage of using the nonlinear components only as that shown in Figure 16.93. In the
figure the actual experimental buckling load M;, is represented by the bisecting straight line
M* = (M I M;,) = I with unit slope. As one can see, the plotted points corresponding to 8"'
are closer to this actual buckling load line and define a longer linear region than the points
obtained from the total deformations 8. The slope of the linear region of the experimental
points yields the Southwell method prediction Msouthw·
The values of the Southwell prediction computed from the total measured rotation, vertical
displacement and crown hoop strains were found to be on the average 17 percent higher
than M;,, whereas the predictions computed from their nonlinear components were on the
average 5.5 percent above M". Also the variations between the results obtained with different
deformation variables were much smaller when the nonlinear components were employed.
Sobel also compared the Southwell prediction for "nonbuckling tests," i.e. if the test had
been terminated before buckling took place. For example, it was assumed that the test was
terminated at a load level for which the elbow experienced nonlinear behavior, say at 90
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1400 Plastic Buckling Experimtmts

1.0

.B

.6 Actual Buckling
Load, M" = 1

.4

Hat Test Pesults

.2

0
0 .2 ,4 .6 .8 1.0

8. 8 .nl
or

M• M•

Figure 16.93 Sobel's application of the Southwell method to plastic buckling of elbows~dimen­
sionless Southwell plots based on the measured rotation (from [16.154])

percent of the buckling load, when the four deformation variables were at 48-56 percent of
their values at buckling. The predictions, based on the nonlinear components of the defor-
mation variables, for (M I M 1,) ::; 0.9 were found to be very close to the experimental buckling
load (within less than 3 percent). Similar comparisons for the cold elbow test yielded simi-
lar results. The Southwell predictions based on the nonlinear component of deformation
were on the average about 2.5 percent greater than M;,. If the test had been terminated at
(M I M;,) = 0.92, for which the elbow still exhibited significant nonlinear behavior, the
Southwell predictions would have been about 5 percent below the buckling load. Note that
this negative difference does not contradict the upper bound assertion of Eq. (16.35), since
it is valid only in the immediate vicinity of buckling.
Hence it appeared that for the elbows the Southwell method provided a reasonably accurate
estimate of the buckling load, without the structure actually having to be tested to buckling.
Thus, here the Southwell plot proved to be a nondestructive test method. It should, however,
be recalled that the nondestructive application of the Southwell method is limited to certain
types of structures and buckling behavior and requires considerable care, since usually the
buckling load has to be approached rather closely.
More recently, Galletly and his co-workers at Liverpool University applied the Southwell
method to plastic buckling of internally pressurized machined torispherical shells [9.372].
Their test setup (Figure 9.143a), the techniques for detecting the threshold buckling pressure
and the development of the buckling waves were discussed in Subsection 9 .11.3 of Chapter
9. Here only some aspects of plastic buckling and the applicability of the Southwell plot
will be considered.
One of the purposes of this investigation was a clarification of the plastic buckling paradox
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(discussed in Subsections 16.1.3 and 16.3.3). The bifurcation buckling pressures of the six
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Southwell's Method in the Plastic Range 1401

torispherical domes tested were therefore computed with the BOSOR5 code, using both the
deformation theory option and the flow theory one. As mentioned in Subsection 9.11.3 (or
see [9.372]), two experimental buckling pressures were obtained for each dome: one, the
incipient or threshold pressure, at which low-amplitude waves were detected by probes in
the knuckle region (or alternatively it was obtained from the chart recorder traces), and the
other, significantly higher buckling pressure at which the initial buckle had grown to be
visible to the naked eye. The BOSOR5 predictions were compared with the threshold ex-
perimental pressures.
The predictions with the deformation theory option of BOSOR5 were in fair agreement
with the experimental results for all six domes, withpoxr(threshold)/p,hder = 0.71-0.94. The
flow theory option, on the other hand, did not predict any buckling for four out of the six
domes, though in two of these four buckles visible to the naked eye were observed and in
the other two the measured low-amplitude waves could be felt with one's fingertips. For the
remaining two the flow theory option predicted buckling but at significantly higher pressures,
19 and 35 percent above the experimental threshold pressures, compared with 6 and 16
percent for the deformation predictions. (One may recall that Bushnell used the deformation
theory shear modulus in the BOSOR5 flow theory option; otherwise the predictions would
have been even higher.)
These Liverpool experiments provide therefore a further confirmation of the plastic buck-
ling paradox.
Galletly pointed out [9.372] that preliminary results of computations, using flow theory,
by Dr. Combescure with the French code INCA [16.159], and by Professor Esslinger with
the German code F04B08, but with the inclusion of small imperfections, predicted buckling
for dome 2. Thus, the small imperfections changed the no-buckling prediction of the cor-
responding perfect dome. However, though the most degrading imperfection shape, that of
the plastic buckling mode, was chosen, the predicted buckling pressure was still rather high,
unless the imperfection amplitude was large (of the order of the wall thickness). The inclu-
sion of small initial imperfections, tried many years before in other structures (see for ex-
ample [16.28]) therefore does not solve the plastic buckling paradox here.
For three of the domes, the buckling pressure was also obtained by application of South-
well plots. The procedure for a typical specimen, say dome 2, was as follows. From the
deflection data stored in the Opus PC, three points (numbers 70, 148 and 174) on the dome,
which were located roughly where buckles formed eventually. The total deflection o of these
three points were plotted versus the pressure p. A plot of the nondimensionalized total
deformation (oft,), where t, is the mean thickness in the knuckle region at one of the points,
point 148, is shown in Figure 16.94a, for probe D (a probe located at about the middle of
the knuckle). The nonlinear component of the deflection om is obtained, as displayed in the
figure.
Then the values of (omit,) are plotted versus (o,,Jt,)lp in Figure 16.94b, being the South-
well plot for dome 2. In the figure the triangles represent values for point 174, the squares
those for point 70 and the circles those for point 148. The points were close to a straight
line, as is apparent in Figure 16.94b, and its slope yielded the predicted buckling pressure
p,, = 0.86 N/mm2 . The magnitude of the predicted buckling pressure was found to depend
on the probe location and hence an average of the predictions for C and D was used.
The Southwell predictions for the three domes considered were 18-24 percent above the
experimental threshold buckling pressures, or 8-16 percent above the chart recorder ones.
The experimental values were expected to be lower than the Southwell predictions, which
are estimates for the corresponding perfect shells.
One should note that the pressures used in the construction of the Southwell plots were
close to the predicted buckling pressures. The range of pressures available for the Southwell
plots was quite limited, the lowest pressure ±rom which the construction could commence
being 0.84-0.89 of the predicted buckling pressure. This indicates one of the potential dif-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

ficulties of the method.


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1402 Plastic Buckling Experiments

-6-
1
mD2
°

o.o
4
L ~-
·-;;-_':;;_:------::-':::------ _ ____j
{/JPomt 148

no 0.25 oso o.7" 10

Ia)

Dome 2, ProbeD

0 25

0 20

Poim 70 . . . ._
6nl 0 15-
tm
Pomt 148 ~
Per= BC/AC = 0.86


0.10

0 05

) c
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

0 Ot
0 0.1

IbI

Figure 16.94 Galletly's application of the Southwell method to plastic buckling of internally pressured
torispherical shells (from [9.372]): (a) radial deflections at a typical point versus internal
pressure, (b) Southwell plot for dome 2, probe D

Finally, it may be noted that for the domes considered the Southwell buckling pressures
agreed well with the predictions of deformation theory.
Another successful engineering application of the Southwell plot to plastic buckling was
the investigation of short, thick-walled cylindrical shells under combined loading by Giezen,
Babcock and Singer at Caltech [16.146], discussed in Subsection 16.3.3. The Southwell
method was employed there to smooth the data of the parametric experimental study. As
shown in Figure 16.78, the smoothing was effective for set A of the tests. Figure 16.95
presents a similar smoothing for the remaining tests (set B), again shown to be effective.
It should be reiterated that as yet no rigorous proof has been derived for the applicability
of Southwell plots to plastic bifurcation problems of shells. Though the use of Southwell
plots in inelastic columns has been theoretically justified for strain-hardening (see Subsection
16.4.1 and [16.156] and [16.15]), its extension to plastic buckling of shells is still only based
on experimental experience. This does not preclude its use in future tests, since the actual
engineering applications have been rather successful, but simply indicates that it should be
done with caution.
Before we leave this subsection, a Liverpool University study on the plastic buckling
paradox should be mentioned [ 16.160], though it did not use the Southwell method. It is an
extension of the earlier Caltech study on short cylindrical shells [16.146]. The loading was
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Some General Remarks on Plastic Buckling 1403

171 00
• = Buckling pressure using Southwell Method
I =Incipient buckling pressure (max. recorded)

128.25

0
"-
,....
~

.
~ 85.50
I.
<J> 1112100

!!
;; .... • •
c

>- 42.75 ....
.... •
....
0

0 1. 56 3. 12 4.68 6. 24 7.80
Externol Pressur& (Mpo)

Figure 16.95 Caltech plastic buckling tests on short cylindrical shells subjected to biaxial loading-
Southwell plot smoothing of experimental results for set B (from [ 16.146] and [ 16.147])

similar, external pressure plus axial tension, but the test specimens were machined mild-steel
models instead of the drawn aluminum alloy tubes of the Caltech experiments and the ex-
perimental, setup differed slightly.
Thirty mild-steel, 50-mm diameter cylinders, with (R/ t) = 24-27 and (L/ R) = 2, 3 and
4, machined from a 3-m long 70-mm diameter bar, were tested. Each cylinder had two
relatively heavy integral flanges. The radial displacement measurements were taken with a
Bentley Nevada proximity probe rotating inside the specimens and logged by a computer.
The cylinders were mounted inside a pressure chamber to allow simultaneous exertion of
both external pressure and axial tension. In the majority of the tests the axial force was kept
constant while the radial pressure was increased, i.e. nonproportional loading.
The BOSORS code, with J 2 flow theory and J2 deformation theory options, was employed
for prediction of buckling pressures. The comparisons of experimental results with predic-
tions for two of the groups of shells, the short ones with (L/ D = 1 and the relatively long
ones with (L! D) = 2 are presented in Figure 16.96. A similar comparison for the medium-
length shells with (LID) = 1.5 is given in Figure 10 of [ 16.160].
From the comparisons it is evident that the numerical predictions based on J 2 deformation
theory were in reasonably good agreement with the experimental results, whereas some of
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

the J 2 flow theory predictions did not agree with the test results. But the longer the specimens,
the better the agreement between the two theories. For the higher axial stresses, however, J2
flow theory failed to predict bifurcation, whereas buckling did occur in the tests.
Hence, these Liverpool tests are another confirmation of the plastic buckling paradox.

16.5 Some General Remarks on Plastic Buckling


As was evident throughout this chapter, the plastic buckling paradox is still unexplained, in
spite of considerable experimental and theoretical efforts (see Subsections 16.1.3, 16.3.3,
16.4.2 and the references cited there, as well as [16.161]-[16.163] and [16.167]-[16.169]).
The scarcity of plastic buckling experiments, in particular for shells, as well as certain in-
complete theoretical justifications were also emphasized there. Considerable research efforts,
especially experimental investigations, are therefore needed before reliable predictions for
plastic buckling strength become available for most structures.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1404 Plastic Buckling Experiments

AXIOI
S~ress Ax1al Tl?llSJOfl

~:!" ', 1) lkNIO D

20
{Lower\ I Upper
Yield)""' I Upper
Yield)
Yield)
·'
v'\ ·"
,F
\ "•
\ '"
\ ss. \ ll
\ F \
\ I S•
I
I
I
\
\ • I \ •"
\
11

•=Expl'I"Jmerlt \ \ •= Experiment
0' O.fucmatuon Thoocy \.so : 0: DehrmaHon Theory
F' Flow Theory \ F: Flow Theory
1

\
(a) o" ·~.-----o---------oc-l---'<L---. 'o ·~,~----~----~,~~~~-.
piNimm'l (b) oiNilnm'l

Figure 16.96 Liverpool University plastic buckling tests on mild-steel shells subjected to combined
axial tension and external pressure-comparison of experiment and deformation and
flow theories (from [16.160]): (a) interaction buckling curves for the short shells with
(LID)"" 0.98, (b) interaction curve for the longest shells with (LID)"" 1.96

The experimenter should remember, however, that in actual structures elasto-plastic buck-
ling behavior is governed by a complex interaction of geometric and material nonlinearities
and can therefore be only partly explained by simple theoretical models. Hence, more so-
phisticated plastic models will probably have to be referred to for comparison with experi-
ments.
Finally, (following some comments by Professor D.R. Sherman of the University of
Wisconsin-Milwaukee), one should remember that in inelastic testing, including plastic buck-
ling tests, large displacements are applied and the resulting loads are measured, whereas in
elastic testing usually loads are applied and the resulting displacements are measured. Fur-
thermore (as remarked by Professor S. Kyriakides of the University of Texas at Austin). in
plastic buckling one is interested in both the deformation (strain) and the load (stress) at the
onset of instability. Since in many applications the processes are deformation controlled, they
extend deeply into the plastic region. The strain is therefore an important variable in plastic
buckling tests.

References
16.1 Gerard, G. and Becker, H., Handbook of Structural Stability, Part III: Buckling of Curved Plates
and Shells, NACA TN 3783, 1957.
16.2 Gerard, G., Handbook of Structural Stability, Supplement to Part III-Buckling of Curved
Plates and Shells, NASA TN D-163, 1959.
16.3 Bushnell, D., Plastic Buckling, in: Pressure Vessels and Piping: Design Technology-1982, A
Decade of Progress, S.Y. Zamrik and D. Dietrich, eds., ASME, New York, 1982, 47-117.
16.4 Hutchinson, J.W., Plastic Buckling, Advances in Applied Mechanics, 14, 1974, 67-144.
16.5 Harding, J.E. and Dowling, P.J., Analytical Results for the Behaviour of Ring and Stringer
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Stiffened Shells, in: Buckling of Shells in Offshore Structures, J.E. Harding, P.J. Dowling and
N. Agelidis, eds., Granada, London, 1982, 231-256.
16.6 Tvergaard, V. and Needleman, A., On the Foundations of Plastic Buckling, in: Developments
in Thin-Walled Structures-], J. Rhodes and A.C. Walker, eds., Applied Science, London, 1982,
205-233.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1405

16.7 Hutchinson, J.W., Post-Bifurcation Behavior in the Plastic Range, Journal of the Mechanics
and Physics of Solids, 21, 1973, 163-190.
16.8 Hutchinson, J.W., Imperfection Sensitivity in the Plastic Range, Journal of the Mechanics and
Physics of Solids, 21, 1973, 191-204.
16.9 Needleman, A. and Tvergaard, V., An Analysis of the Imperfection Sensitivity of Square Elastic-
Plastic Plates under Axial Compression, International Journal of Solids and Structures, 12,
1976, 185-201.
16.10 Tvergaard, V., Buckling of Elastic-Plastic Cylindrical Panel under Axial Compression, Inter-
national Journal of Solids and Structures, 13, 1977, 957-970.
16.11 Tvergaard, V. and Needleman, A., On the Localization of Buckling Patterns, Journal of Applied
Mechanics, 47, 1980, 613-619.
16.12 Engesser, F., Uber die Knickfestikcit gerader Stabe, Zeitschrift des Architekten und Ingenieur-
vereins zu Hannover, 35, (4), 1889, 455-462.
16.13 Shanley, F.R., The Column Paradox, Journal of the Aeronautical Sciences, 13, (12), 1946, 678.
16.14 von Karman, T., Discussion of "Inelastic Column Theory," Journal of the Aeronautical Sci-
ences, 14, 5, 1947, 267-268.
16.15 Wang, C.T., Inelastic Column Theories and an Analysis of Experimental Observations, Journal
of the Aeronautical Sciences, 15, (5), 1948, 283-292.
16.16 Lin, T.-H., Inelastic Column Buckling, Journal of the Aeronautical Sciences, 17, (3), 1950,
159-172.
16.17 Cicala, P., Column Buckling in the Elastoplastic Range, Journal of the Aeronautical Sciences,
17, (8), 1950, 508-512.
16.18 Pearson, C.E., Bifurcation Criterion and Plastic Buckling of Plates and Columns, Journal of
the Aeronautical Sciences, 17, (7), 1950, 417-424.
16.19 Hoff, N.J., Plastic Column Behavior, Journal of" the Aeronautical Sciences, 17, (11), 1950, 743-
744.
16.20 Duberg, J.E. and Wilder, T.W., III, Column Behavior in the Plastic Stress Range, Journal of
the Aeronautical Sciences, 17, (6), 1950, 324-327.
16.21 Duberg, J.E. and Wilder, T.W., III, Inelastic Column Behavior, NACA TN 2267, Jan. 1951.
16.22 Hoff, N.J., Buckling and Stability, 41st Wilbur Wright Memorial Lecture, Journal of the Royal
Aeronautical Society, 58, (!), 1954, 3-58.
16.23 Hoff, N.J., Inelastic Buckling of Columns in the Conventional Testing Machine, in: Theory of
Shells, Proceedings of a Symposium to Honor Lloyd Hamilton Donnell, D. Muster, ed., Uni-
versity of Houston, Texas, 1967, 385-402.
16.24 Templin, R.L., Sturm, R.G., Hartman, E.C. and Holt, M., Column Strength of Various Alumi-
num Alloys, Aluminum Research Laboratories Technical Paper 1, Aluminum Company of
America (ALCOA), Pittsburgh, 1938.
16.25 Stowell, E.Z., Critical Shear Stress of an Infinitely Long Plate in the Plastic Region, NACA
TN 1681, August 1948.
16.26 Stowell, E.Z., A Unified Theory of Plastic Buckling of Columns and Plates, NACA Report 898,
1948.
16.27 Stowell, E.Z., Compressive Strength of Flanges, NACA Report 1029, 1951.
16.28 Onat, E.T. and Drucker, D.C., Inelastic Instability and Incremental Theories of Plasticity, Jour-
nal of the Aeronautical Sciences, 20, (3), 1953, 181-186.
16.29 Hutchinson, J.W. and Budiansky, B., Analytical and Numerical Study of the Effects of Initial
Imperfection on the Buckling of a Cruciform Column, in: Buckling of Structures, Proceedings
of IUTAM Symposium, Harvard University, Cambridge, Mass., June 17-21, 1974, B. Budi-
ansky, ed., Springer-Verlag, Berlin, 1976, 98-105.
16.30 Batdorf, S.B., Theories of Plastic Buckling, Journal of the Aeronautical Sciences, 16, (7), 1949,
405-408.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

16.31 Hill, R., The Essential Structure of Constitutive Laws for Metal Composites and Polycrystals,
Journal of the Mechanics and Physics of Solids, 15, 1967, 79-95.
16.32 Hutchinson, J.W., Elastic/Plastic Behavior of Polycrystalline Metals and Composites, Proceed-
ings of the Royal Society of" London, Series A, 319, 1970, 247-272.
16.33 Christoffersen, J. and Hutchinson, J.W., A Class of Phenomenological Comer Theories of Plas-
ticity, Journal of the Mechanics and Physics of Solids, 27, 1979, 465-487.
16.34 Needleman, A. and Tvergaard, V., Aspects of Postbuckling Behavior, in: Mechanics of Solids,
The Rodney Hill 60th Anniversary Volume, H.G. Hopkins and M.J. Sewell, eds., Pergamon
Press, Oxford, 1982, 453-498.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1406 Plastic Buckling Experiments

16.35 Bodner, S.R., A Lower Bound on Bifurcation Buckling of Viscop1astic Structures, Acta Me-
chanica Supplement, 3, 1992, 181-190.
16.36 Rubin, M.B. and Bodner, S.R., An Incremental Elastic-Viscoplastic Theory Indicating a Re-
duced Modulus for Non-Proportional Buckling, International Journal of Solids and Structures,
32, (20), 1995, 2967-2987.
16.37 Ore, E. and Durban, D., Elastoplastic Buckling of Axially Compressed Circular Cylindrical
Shells, International Journal of Mechanical Sciences, 34, (9), 1992, 727-742.
16.38 Batterman, S.C., Plastic Buckling of Axially Compressed Cylindrical Shells, AIAA Journal, 3,
(2), 1965, 316-325.
16.39 Ore, E. and Durban, D., Elastoplastic Buckling of Annular Plates in Pure Shear, ASME Journal
of Applied Mechanics, 56, (3 ), 1989, 644-651.
16.40 Bauer, M., Faltenbildung beim ebenen Torsionsversuch, Ingenieur-Archiv, 57, 1987, 39-50.
16.41 Lange, K. and Bauer, M., Determining Flow Curves of Thin Sheet Metal by the Plane Torsion
Test, SME (Society of Manufacturing Engineers) Technology Review, 2, 1987, 346-352.
16.42 Bauer, M., Ermittlung der Fliesskun,en von Feinblechen im ebenen Torsionsversuch, Institut
fUr Umformtechnik Universitat Stuttgart, Springer-Verlag, Berlin, London, New York, 1989.
16.43 Osgood, W.R., Column Strength of Tubes Elastically Restrained against Rotation at the Ends,
NACA Report 615, 1938.
16.44 Chilver, A.H. and Brivtec, S.J., The Plastic Buckling of Aluminum Columns, in: Aluminium
Structural Engineering, Proceedings of an Institution of Structural Engineers and Aluminium
Federation Symposium, London, June: 1963, Aluminium Federation, London, 1964, 47-58 and
113-130.
16.45 Gerard, G., Secant Modulus Method for Determining Plate Instability above the Proportional
Limit, Journal of the Aeronautical Sciences, 13, (1), 1946, 38-44, 48.
16.46 Bijlaard, P.P., A Theory of Plastic Stability and Its Application to Thin Plates of Structural
Steel, Proceedings Kon. Akad. v. Wetensch. Amsterdam, 41, (7), Sept. 1938, 731-743.
16.47 Bijlaard, P.P., Theory of the Plastic Stability of Thin Plates, Pub!. International Association for
Bridge and Structural Engineering, Zurich, 6, 1940-1941, 45-69.
16.48 Bijlaard, P.P., On the Plastic Stability of Thin Plates and Shells, Proceedings Kon. Akad. v.
Wetensch. Amsterdam, 50, (7), Sept. 1947, 765-775.
16.49 Bijlaard, P.P., Kollbrunner, C.F. and Sttissi, F., Theorie und Versuche tiber das plastische Aus-
beulen von Rechteckplatten unter gleichmiissig verteilten Langsdruck, International Association
for Bridge and Structural Engineering . 3d Congress, Liege, Sept. 1948, Preliminary Publication,
1948, 119-128.
16.50 Bijlaard, P.P., Theory and Tests on the Plastic Stability of Plates and Shells, Journal of the
Aeronautical Sciences, 16, (9), 1949, 529-541.
16.51 Lundquist, E.E., Schuette, E.H., Heimerl, G.J. and Roy, J.A., Column and Plate Compressive
Strengths of Aircraft Structural Materials-24S-T Aluminum-Alloy Sheet, NACA ARR No.
L5F01, June 1945 (Wartime Report).
16.52 Heimerl, G.J. and Roy, J.A., Column and Plate Compressive Strengths of Aircraft Structural
Materials-Extruded 75S-T Aluminum Alloy, NACA ARR No. L5F08a, July 1945 (Wartime
Report).
16.53 Heimerl, G.J. and Roy, J.A., Column and Plate Compressive Strengths of Aircraft Structural
Materials-Extruded 24S-T Aluminum Alloy, NACA ARR No. L5F08b, July 1945 (Wartime
Report).
16.54 Heimerl, G.J. and Fay, D.P., Column and Plate Compressive Strengths of Aircraft Materials-
Extruded R303-T Aluminum Alloy, NACA ARR No. L5H04, October 1945 (Wartime Report).
16.55 Pride, R.A. and Heimerl, G.J., Plastic Buckling of Simply Supported Compressed Plates, NACA
TN 1817, April 1949.
16.56 Schuette, E.H. and McDonald, J.C., Prediction and Reduction to Minimum Properties of Plate
Compressive Curves, Journal of the Aeronautical Sciences, 15, (1), 1948, 23-27.
16.57 Heimerl, G.J. and Roy, J.A., Determination of Desirable Lengths of Z- and Channel-Section
Columns for Local-Instability Tests, NACA RB No. L4H10, 1944 (Wartime Report).
16.58 Ilyushin, A.A., Stability of Plates and Shells Beyond the Proportional Limit, Prikladnaya Ma-
tematika i Mekhanika, 8, (5), 1944; Translation NACA TM 1116, Oct. 1947.
16.59 Handelman, G.H. and Prager, W., Plastic Buckling of a Rectangular Plate under Edge Thrusts,
NACA Report 946, 1949. --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1407

16.60 Inoue. T., Analysis of Plastic Buckling of Rectangular Steel Plates Supported along Their Four
Edges, International Journal of Solids and Structures, 31, (2), 1994, 219-230.
16.61 Gerard, G., Critical Shear Stress of Plates above the Proportional Limit, Journal of Applied
Mechanics, 15, (1), 1948, 7-12.
16.62 Coker, E.G. and Filon, L.N.G., Photo-Elasticity, Cambridge University Press, 1931, 601-611.
16.63 Marciniak, Z. and Kolodziejski, J., Assessment of Sheet Metal Failure Sensitivity by Method
of Torsioning the Rings, in: Proceedings of the 7th Biennial Congress of the International Deep
Drawing Research Group (IDDRG), Amsterdam, 1972, 6.1-6.3.
16.64 Sowerby, R., Tomita, Y. and Duncan, J.L., In-Plane Torsion Testing of Sheet Metal, Journal of
Mechanical Engineering Science, 19, (5), 1977, 213-220.
16.65 Tekkaya, A.E. and Pohlandt. K., Determining Stress-Strain Curves of Sheet Metal in the Plane
Torsion Tests, Annals of the CIRP, 31, 1982, 171-173.
16.66 Bauer, M. and Pohlandt. K., Fundamentals of the Plane Torsion Test for Determining Flow
Curves of Thin Sheet, Materialpriifung, 28, 1986, 220-225.
16.67 Murray, N.W. and Khoo, P.S., Some Basic Plastic Mechanisms in the Local Buckling of Thin-
Walled Steel Structures, International Journal of Mechanical Sciences, 23, (12), 1981, 703-
714.
16.68 Murray, N.W., Das Stabilitatsverhalten von axial belasteten, in der Langsrichtung ausgesteiften
Platten im plastischen Bereich, Der Stahlbau, 12, 1973, 372-379.
16.69 Munay, N.W., Ultimate Capacity of Stiffened Plates in Compression, in: Plated Structures,
Stability and Strength, R. Narayanan, ed., Applied Science Publishers, London and New York,
1983, 135-163.
16.70 Munay, N.W., The Static Approach to Plastic Collapse and Energy Dissipation in Some Thin-
Walled Steel Structures, in: Structural Crashworthiness, N. Jones and T. Wierzbicki, eds., Lon-
don, Butterworths, 1983, 44-65.
16.71 Munay, N.W., Recent Research into the Behaviour of Thin-Walled Steel Structures, in: Inter-
national Conference on Steel Structures, Recent Research Advances and Their Application to
Design, M.N. Pavlovic, ed., Elsevier, Amsterdam, 1986, 171-191.
16.72 Munay, N.W., Some Phenomena Observed in Thin-Walled Square Box-Sections under Axial
Bending and Torsional Loading, in: Proceedings of the International Conference on Steel Struc-
tures, Cardiff, Wales, 1987, 465-474.
16.73 Munay, N.W., Das aufnehmbare Moment in einem zur Richtung der Normalkraft schrag lie-
genden plastischen Gelenk, Die Bautechnik, 2, 1973, 57-58.
16.74 Khoo, P.S., Plastic Local Buckling of Thin-Walled Structures, Ph.D. thesis, Monash University,
Australia, Department of Civil Engineering, 1979.
16.75 Rawlings, B. and Shapland, P., The Behaviour of Thin-Walled Box Sections under Gross De-
formation, Structural Engineer, 53, (4), April 1975, 181-186.
16.76 Calladine, C.R., The Strength of Thin Plates in Compression, in: Aspects of the Analysis of
Plate Structures, D.J. Dawe, R.W. Horsington, A.G. Kamtekar and G.H. Little, eds., Clarendon
Press, Oxford, 1985, 269-291.
16.77 Ronda!, J. and Maquoi, R., Stub-Column Strength of Thin-Walled Square and Rectangular
Hollow Sections, Thin- Walled Structures, 3, 1985, 15-34.
16.78 Sobel, L.H. and Newman, S.Z., Plastic Buckling of Cylindrical Shells under Axial Compression,
ASME Journal of Pressure Vessel Technology, 102, Feb. 1980, 40-44.
16.79 Osgood, W.R., The Crinkling Strength and the Bending Strength of Round Aircraft Tubing,
NACA Report 632, 1938.
16.80 Gerard, G., Compressive and Torsional Buckling of Thin-Walled Cylinders in Yield Region,
NACA TN 3726, August 1956.
16.81 Moore, R.L. and Holt, M., Beam and Torsion Tests of Aluminum-Alloy 61S-T Tubing, NACA
TN 867, 1942.
16.82 Moore, R.L. and Paul, D.A., Torsional Stability of Aluminum Alloy Seamless Tubing, NACA
TN 696, March 1939.
16.83 Lee, L.H.N., Inelastic Buckling of Initially Imperfect Cylindrical Shells Subject to Axial Com-
pression, Journal of the Aerospace Sciences, 29, Jan. 1962, 87-95.
16.84 Horton, W.H., Bailey, S.C. and Edwards, A.M., Nonsymmetric Buckle Patterns in Progressive
Plastic Buckling, Experimental Mechanics, 6, Sept. 1966, 433-444.
16.85 Stodlkers, B., On Buckling of Axisymmetric Thin Elastic-Plastic Shells, International Journal
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
of Solids and Structures, 11, 1975, 1139-1346.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1408 Plastic Buckling Experiments

16.86 Johnson, W., Soden, P.D. and A1-Hassani, S.T.S., Inextensional Collapse of Thin-Walled Tubes
under Axial Compression, Journal of Strain Analysis, 12, (4), 1977, 317-330.
16.87 Gellin, S., Effect of an Axisymmetric Imperfection on the Plastic Buckling of an Axially Com-
pressed Cylindrical Shell, Journal of Applied Mechanics, 46, March 1979, 125-131.
16.88 Bomscheuer, F.W., Plastisches Beulen von Kreiszylinderschalen unter Axialbelastung, Der
Stahlbau, 50, (9), Sept. 1981, 257-262.
16.89 Tvergaard, V., On the Transition from a Diamond Mode to an Axisymmetric Mode of Collapse
in Cylindrical Shells, International Journal of Solids and Structures, 19, (10), 1983, 845-856.
16.90 Geckeler, J.W., Plastisches Knicken der Wandung von Hohlzylindern und einige andere Fal-
tungserscheinungen an Schalen und Blechen, ZAMM, 8, 1928, 341-352.
16.91 Bushnell, D. and Meller, E., Elastic-Plastic Collapse of Axially Compressed Cylindrical Shells:
A Brief Survey with Particular Application to Ring-Stiffened Cylindrical Shells with Reinforced
Openings, ASME Journal of Pressure Vessel Technology, 106, Feb. 1984, 2-15.
16.92 Murphy, L.M. and Lee, L.H.N., Inelastic Buckling Process of Axially Compressed Cylindrical
Shells Subject to Edge Constraints, lnternational Journal of Solids and Structures, 7, Sept.
1971, 1153-1170.
16.93 Mayers, J. and Wesenberg, D.L., The Maximum Strength of Initially Imperfect, Axially Com-
pressed, Circular Cylindrical Shells, Stanford University, USAAVLABS Technical Report 69-
60, U.S. Army Aviation Material Laboratories, Fort Eustis, Va., August 1969, AD 862102.
16.94 Wesenberg, D.L. and Mayers, J., Failure Analysis of Initially Imperfect, Axially Compressed,
Orthotropic, Sandwich, and Eccentrically Stiffened, Circular Cylindrical Shells, Stanford Uni-
versity, USAAVLABS Technical Report 69-86, U.S. Army Aviation Material Laboratories, Fort
Eustis, Va., Dec. 1969, AD 866199.
16.95 Mayers, J. and Meller, E., Material Nonlinearity Effects in Optimization Considerations of
Stiffened Cylinders and Interpretation of Test Data Scatter for Compressive Buckling, Stanford
University, USAAMRDL Technical Report 71-70, U.S. Army Air Mobility Research and De-
velopment Laboratory, Fort Eustis, Va., March 1972.
16.96 Reynolds, T.E., Inelastic Lobar Buckling of Cylindrical Shells under External Hydrostatic Pres-
sure, U.S. Navy David Taylor Model Basin, DTMB Report 1392, August 1960.
16.97 Boichot, L. and Reynolds, T.E., Inelastic Buckling Tests of Ring-Stiffened Cylinders under
Hydrostatic Pressure, U.S. Navy David Taylor Model Basin, DTMB Report 1992, May 1965.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
16.98 Lunchick, M,E., Plastic Axisymmetric Buckling of Ring-Stiffened Cylindrical Shells Fabricated
from Strain-Hardening Materials and Subjected to External Hydrostatic Pressure, U.S. Navy
David Taylor Model Basin, DTMB Report 1393, Jan. 1961.
16.99 Lunchick, M.E., Plastic General Instability of Ring-Stiffened Cylindrical Shells, U.S. Navy
David Taylor Model Basin, DTMB Report 1587, Sept. 1963.
16.100 Krenzke, M.A. and Kiernan, T.J., Structural Development of a Titanium Oceanographic Vehicle
for Operating Depths of 15,000 to 20,000 Feet, U.S. Navy, David Taylor Model Basin, DTMB
Report 1677, 1963.
16.101 Montague, P., Experimental Behaviour of Thin-Walled Cylindrical Shells Subjected to External
Pressure, Journal of Mechanical Engineering Science, 11, (1), 1969, 40-56.
16.102 DeHart, R.C. and Basdekas, N.L., Yield Collapse of Stiffened Circular Cylindrical Shells,
Southwest Research Institute, San Antonio, Tex., Sept. 1960 and May 1961.
16.103 Augusti, G. and D' Agostino, S., Tests of Cylindrical Shells in the Plastic Range, ASCE Journal
of Engineering Mechanics Division, 90, Febr. 1964, 69-81.
16.104 Simonen, F.A. and Shippell, R.J., Jr., Collapse of Thick-Walled Cylinders under External Pres-
sure, Experimental Mechanics, 22, Feb. 1982, 41-48.
16.105 Ross, C.T.F., Haynes, P., Seers, A. and Johns, T., Inelastic Buckling of Ring-Stiffened Circular
Cylinders under Uniform External Pressure, in: Structural Dynamics and Vibration 1995,
ASME, PD-vol. 70, B.A. Ovung, 1.1. Esat, A.B. Sabir and V. Karadag, eds., ASME, New York,
1995, 208-215.
16.106 Ades, C.S., Bending Strength of Tubing in the Plastic Range, Journal of the Aeronautical
Sciences, 24, August 1957, 605-610.
16.107 Gellin, S., The Plastic Buckling of Long Cylindrical Shells under Pure Bending, International
Journal of Solids and Structures, 16, 1980, 397-407.
16.108 Reddy, B.D., An Experimental Study of the Plastic Buckling of Circular Cylinders in Pure
Bending, International Journal of Solids and Structures, 15, 1979, 669-683.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1409

16.109 Sherman, D.R., Inelastic Flexural Buckling of Cylinders, in: International Conference on Steel
Structures, Recent Research Advances and Their Application to Design, M.N. Pavlovic, ed.,
Elsevier, Amsterdam 1986, 339-357.
16.110 Schilling, C. G., Buckling Strength of Circular Tubes, ASCE Journal of the Structural Division,
91, (ST5), Oct. 1965, 325-348.
16.111 Sherman, D.R., Tests of Circular Steel Tubes in Bending, ASCE Journal of the Structural
Division, 102, (STll), Nov. 1976, 2181-2195.
16.112 Corona, E. and Kyriakides, S., An Experimental Investigation of the Degradation and Buckling
of Circular Tubes Under Cyclic Bending and External Pressure, Thin-Walled Structures, 12,
1991, 239-263.
16.113 Jirsa, J.O., Lee, F.-H., Wilhoit, J.C. and Mervin, J.E., Ovaling of Pipelines Under Pure Bending,
in: OTC 1569, Proceedings, 4th Offshore Technology Conference, Houston, 1972, vol. I, 573-
579.
16.114 Johns, T.G., Mesloh, R.E., Winegardner, R. and Sorenson, J.E., Inelastic Buckling of Pipelines
under Combined Loads, OTC 2209, in: Proceedings, 7th Offshore Technology Conference, 1975,
vol. 2, 635-646.
16.115 Tugcu, P. and Schroeder, J., Plastic Deformation and Stability of Pipes Exposed to External
Couples, International Journal of Solids and Structures, 15, 1979, 643-658.
16.116 Korol. R.M., Critical Buckling Strains of Round Tubes in Flexure, International Journal of
Mechanical Sciences, 21, 1979, 719-730.
16.117 Bouwkamp, J.G. and Stephen, R.M., Large Diameter Pipe Under Combined Loading, ASCE
Tran~portation Engineering Journal, 99, TE3, Paper 9907, August 1973, 521-536.
16.118 Wilhoit, J.C. and Mervin, J.E., Critical Plastic Buckling Parameters for Tubing in Bending
under Axial Tension, OTC 1874, in: Proceedings, 5th Annual Offshore Technology Conference,
Houston, 1973, vol. 2, 465-472.
16.119 Sherman, D.R., Inelastic Local Buckling of Circular Tubes, in: Buckling of Shells in Offshore
Structures, J.E. Harding, P.J. Dowling, and N. Agelidis, eds., Granada, London, 1982, 365-
392.
16.120 Sherman, D.R., Supplemental Tests for Bending Capacity of Fabricated Pipes, Report, Depart-
ment of Civil Engineering, University of Wisconsin-Milwaukee, Sept. 1984.
16.121 Sherman, D.R., Bending Capacity of Fabricated Pipe at Fixed Ends, Report, Department of
Civil Engineering, University of Wisconsin-Milwaukee, Dec. 1985.
16.122 de Winter, P.E., Stark, J.W.B. and Witteveen, J., Collapse Behavior of Submarine Pipelines, in:
Shell Structures, Stability and Strength, R. Narayanan, ed., Elsevier Applied Science, London,
1985, 221-246.
16.123 Stark, J.W.B. and de Winter, P.E., Plastic Design of Submarine Pipelines, in: Behaviour of Thin-
Walled Structures, J. Rhodes and J. Spence, eds., Elsevier Applied Science, London, 1984, 287-
311.
16.124 Stephans, M.J., Kulak, G.L. and Montgomery, C.J., Local Buckling of Thin-Walled Tubular
Steel Members, in: Stability of Metal Structures, Proceedings of 3rd International Colloquium
on Stability of Metal Structures, Toronto, May 1983, 489-508.
16.125 Kulak, G.L., Tubular Members-Large and Small, in: Proceedings, Structural Stability Research
Council (SSRC) 50th Anniversary Conference, Lehigh University, June 1994, SSRC Lehigh
University, Bethlehem, Pa., 1994, 51-63.
16.126 Obaia, K.H., Elwi, A.E. and Kulak, G.L., Tests of Fabricated Steel Cylinders Subjected to
Transverse Loads, Journal of Constructional Steel Research, 22, 1992, 21-37.
16.127 Mohareb, M., Alexander, S.D.B., Kulak, G.L. and Murray, D.W., Laboratory Testing of Line
Pipe to Determine Deformational Behavior, in: Proceedings of 12th International Conference
on Offshore Mechanics and Arctic Engineering (OMAF), vol. 5, Pipeline Technology, ASME,
New York, 1993, 109-114.
16.128 Obaia, K.H., Elwi, A.E. and Kulak, G.L., Inelastic Transverse Shear Capacity of Large Fabri-
cated Steel Tubes, Structural Engineering Report 170, Department of Civil Engineering, Uni-
versity of Alberta, 1991.
16.129 de Winter, P.E., Strength and Deformation Properties of Pipelines, in: Deep Water. Second Stage
of MaTS Project PL-II-8, IRO, Delft, 1981.
16.130 Otey, N.S., Torsional Strength of Nickel Steel and Duralumin Tubing as Affected by the Ratio
of Diameter to Gage Thickness, NACA TN 189, 1924.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1410 Plastic Buckling Experiments

16.131 Fuller, F.B., The Torsional Strength of Solid and Hollow Cylindrical Sections of Heat-Treated
Alloy Steel, Journal of the Aeronautical Sciences, 3, (7), May 1936, 248-251.
16.132 Onal, Z. and Esin, A., Plastic Torsional Buckling of Thin-Walled Circular Cylinders, Material-
prlifung, 28, (112), Jan.-Feb. !986, 31-35.
16.133 Ramsey, H., Plastic Buckling of Conical Shells Under Axial Compression, International Journal
of Mechanical Sciences, 19, 1977, 257-272.
16.134 Ross, C.T.F., Plastic Collapse of Thin-Walled Ring-Stiffened Conical Shells under Uniform
External Pressure, Journal of Ship Research, 39, (2), June 1995, 166-175.
16.135 Ross, C.T.F. and Seers, A., Inelastic Instability of Thin-Walled Conical Shells under External
Pressure, in: Developments in Computed Aided Design and Modelling for Structural Engineer-
ing, B.H.V. Topping, ed., Civil-Comp Press, Edinburgh, 1995, 177-183.
16.136 Kiernan, T.J. and Nishida, K., The Buckling Strength of Fabricated HY-80 Steel Spherical
Shells, U.S. Navy David Taylor Model Basin, DTMB Report 1721, July 1966.
16.137 Costello, M.G. and Nishida, K., The Inelastic Buckling Strength of Fabricated HY-80 Steel
Hemispherical Shells, U.S. Navy David Taylor Model Basin, DTMB Report 2304, April 1967.
16.138 Krenzke, M.A. and Kiernan, T.J., The Effect of Initial Imperfections on the Collapse Strength
of Deep Spherical Shells, U.S. Navy David Taylor Model Basin, DTMB Report 1757, Feb.
1965.
16.139 Leckie, F.A. and Penny, R.K., Plastic Instability of a Spherical Shell, in: Engineering Plasticity,
J. Heyman and F.A. Leckie, eds., Cambridge University Press, 1968, 401-412.
16.140 Razzaq, Z. and McVinnie, W.W., Theoretical and Experimental behavior of Biaxially Loaded
Inelastic Columns, Journal of Structural Mechanics, 14, (3), 1986, 321-337.
16.141 Razzaq, Z. and McVinnie, W.W., Rectangular Tubular Steel Columns Loaded Biaxially, Journal
of Structural Mechanics, 10, (4), 1983, 475-493.
16.142 Dier, A.F. and Dowling, P.J., The Strength of Plates Subjected to Biaxial Forces, in: Behaviour
of Thin- Walled Structures, J. Rhodes and J. Spence, eds., Elsevier Applied Science, Publishers,
London, New York, 1984, 329-353.
16.143 Babcock, C.D., Plastic Buckling under Biaxial Loads, California Institute of Technology, GAL-
CIT, Pasadena, Calif., SM Report 85-11, 1985.
16.144 Tay, C.J., Steel, W.J.M. and Spence, J., The Inelastic Response of Pipes under External Pressure
and Bending, in: Behaviour of Thin- Walled Structures, J. Rhodes and J. Spence, eds., Elsevier
Applied Science, London, New York, 1984, 313-327.
16.145 Miller, C.D. and Vojta, J.F., Strength of Stiffened Cylinders Subjected to Combinations of Axial
Compression and External Pressure, in: Proceedings, SSRC 1984 Annual Technical Session,
SSRC Fritz Engineering Laboratory, Bethlehem, Pa., 1984, 159-168.
16.146 Giezen, J.J., Babcock, C.D. and Singer, J., Plastic Buckling of Cylindrical Shells under Biaxial
Loading, Experimental Mechanics, 31, (4), 1991, 337-343.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

16.147 Giezen, J.J., Plastic Buckling of Cylinders under Biaxial Loading, Ph.D. thesis, California
Institute of Technology, Pasadena, Calif., GALCIT SM 88-23, June 1988.
16.148 Madhavan, R., Babcock, C.D. and Singer, J., On the Collapse of Long, Thick-Walled Tubes
under External Pressure and Axial Tension, ASME Journal of Pressure Vessel Technology, 115,
Feb. 1993, 15-26.
16.149 Bushnell, D., BOSOR5-A Computer Program for Buckling of Elastic-Plastic Complex Shells
of Revolution Including Large Deflections and Creep, vol. 1, User's Manual, Input Data, LMSC
D407166; vol. 2, User's Manual, Test Cases, LMSC D407167; vol. 3, Theory and Comparison
with Tests, LMSC 407168, Lockheed Missiles Space Co., Sunnyvale, Calif., Dec. 1974.
16.150 Madhavan, R., On the Collapse of Long Thick-Walled Circular Tubes Under Biaxial Loading,
Ph.D. thesis, California Institute of Technology, Pasadena, Calif., GALCIT SM 88-24, March
1988.
16.151 Yeh, M.K. and Kyriakides, S., On the Collapse of Inelastic Thick-Walled Tubes under External
Pressure, ASME Journal of Energy Resources Technology, 108, 1986, 35-47.
16.152 Kyriakides, S. and Yeh, M.K., Plastic Anisotropy in Drawn Metal Tubes, ASME Journal of
Engineering for Industry, 110, 1988, 303-307.
16.153 Newman, J.B., Inelastic Column Buckling of Internally Pressurized Tubes, Experimental Me-
chanics, 13, 1973, 265-272.
16.154 Sobel, L.H., The Southwell Method for Predicting Plastic Buckling Loads for Elbows, Journal
of Pressure Vessel Technology, Transactions ASME, 105, 1983, 2-8.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1411

16.155 Horsfall, W. and Sandorff, P., Strain Distribution during Column Failure, Lockheed Aircraft
Corporation, Burbank, Calif., Report No. 5728, April 1946.
16.156 Singer, J., On the Applicability of the Southwell Plot to Plastic Buckling, Experimental Me-
chanics, 29, (2), 1989, 205-208.
16.157 Kollbrunner, C.F. and Meister, M., Knicken, Biegedrillknicken, Kippen, 2d ed., Springer-Verlag,
Berlin 1961.
16.158 Sobel, L.H. and Newman, S.Z., Comparison of Experimental and Simplified Analytical Results
for the Plastic In-plane Bending and Buckling of an Elbow, ASME Journal of Pressure Vessel
Technology, 102, 1980, 400-409.
16.159 Combescure, A., Upon the Different Theories of Plastic Buckling: Elements for a Choice, in:
Buckling of Shell Structures, on Land, in the Sea and in the Air, J.-F. Jullien. ed., Elsevier
Applied Science, Amsterdam, 1991, 448-457.
16.160 Blachut, J., Galletly, G.D. and James, S., On the Plastic Buckling Paradox, Proceedings of the
Institution of Mechanical Engineers, 210, 1996, 477-488.
16.161 El-Ghazaly, H.A. and Sherbourne, A.N., Flow and Deformation Theories for Plate Plastic Buck-
ling-An Engineering Approach, S.M. Archives, 10, 1985, 257-287.
16.162 Tugcu, P., Plate Buckling in the Plastic Range, International Journal of Mechanical Sciences,
33, 1991, 1-ll.
16.163 Tugcu, P., On Plastic Buckling Predictions, International Journal of Mechanical Sciences, 33,
1991, 529-539.
16.164 Fabian, 0 .. Elastic-Plastic Collapse of Long Tubes under Combined Bending and Pressure
Loads, Ocean Engineering 8, 1981, 295-330.
16.165 Bushnell, D., Elastic-Plastic Bending and Buckling of Pipes and Elbows, Computers and Struc-
tures, 13, 1981, 241-248.
16.166 Corona, E. and Vase, S.P., Buckling of Elastic-Plastic Square Tubes under Bending, Interna-
tional Journal of Mechanical Sciences, 38, (7), 1996, 753-775.
16.167 Inoue, T. and Kato, B., Analysis of Plastic Buckling of Steel Plates, International Journal of
Solids and Structures, 30, (6), 1993, 835-856.
16.168 Inoue, T., Analysis of Plastic Buckling of Rectangular Steel Plates Supported along Their Four
Edges, International Journal of Solids and Structures, 31, (2), 1994, 219-230.
16.169 Inoue, T., Analysis of Plastic Buckling of Steel Plates in Shear Based on the Tresca Yield
Criterion, International Journal of Solids and Structures, 33, (26), 1996, 3903-3923.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
17
Influence of Holes, Cutouts
and Damaged Structures

17.1 Effect of Holes and Cutouts on Plates and Shells


17.1.1 Introduction
Holes and cutouts are essential elements of most practical structural elements, and their effect
on the buckling of plates and shells has therefore been the subject of extensive studies, many
of them experimental. The buckling experiments on plates and shells with holes and cutouts,
however. employed very similar test setups and test techniques to those used in unperforated
plates and shells. The treatment of these experiments here will therefore be rather brief,
discussing only a few examples and assuming that the reader has had the opportunity to
study the relevant tests on unperforated plates and shells, considered in Chapters 8, 9, 12,
13 and 14, respectively.
The additional topic of buckling and strength of damaged structures, in particular damaged
or dented shells, will on the other hand be treated in more detail (in Section 17.4) since it
involves some special test techniques. The effect of delamination on the buckling behavior
of composite structures will also be dealt with.

17.1.2 The Effect of Holes and Cutouts in Plates


The primary effect of holes in plates and shells investigated in the fifties, sixties and seventies
was the stress concentration occurring around them (see for example [17.1]-[17.7]). When
the holes and! cutouts in plates and shells were small, their influence on the buckling behavior
of these structural elements was essentially that of a local imperfection.
In plates these local imperfections hardly affect their buckling load, due to their stable
postbuckling behavior. When subjected to axial compression, bending or shear, plates there-
fore usually continue to sustain substantial buckling loads even in the presence of fairly large
cutouts (see for example Section 4.2.5 of [6.3], [8.54], [17.4] and [17.8]-[17.22]).
The influence of circular holes on the buckling of simply supported square metal plates
subjected to axial compression is shown in Figure 17 . I. In this figure k, is the buckling
coefficient for a plate of width w, having a central hole of diameter d, while k,., is the
buckling coefficient of the corresponding unperforated plate. The experimental values of Yu
and Davies [17.4] are compared in the figure with predictions by the method of Kawai and
Ohtsubo [J 7.23], a combination of a Rayleigh-Ritz procedure and a finite-element method.
It can be seen that for circular holes the reduction in buckling load is indeed not very large.
However, for square holes of the same size (with width h, equal to the diameter d), the
reduction in the buckling load! is more substantial, as shown in Figure 17 .I b. In this figure
Buckliug E.x!Jerimems: E.xperimeutal Methods in Buckling of Thiu· Walled Structures: Shells, Built-Up Strucwres, Composites
and Additioual Topics - Volume 2. J. Singer, J. Arbocz and T. WeUer Copyright © 2002 John Wiley & Sons, Inc.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1414 Influence of Holes, Cutouts and Damaged Structures

1.50

~
Uniform
dispbc~nte•t
l25
!::...
k..
LOO
kawai ••d Ohtsubo Q


o.n Unironn
S4rtu

~
0.50 • Be•m lest
o Column tut

o.2~ol--4-----.o.-l,.z.--~o>:---~o.,...4---cuk---.tt--,ff------t
(a) d/w

0· 2~o!-(-b-)--::o!-:-J--~o.'=-2---;!ol:-3_h,_l_w-;;}o_47---....-:~';---;!o.e;---;!o.1

Figure 17.1 Buckling of simply supported, axially compressed square plates with central holes-
influence of hole size and shape on buckling coefficients and comparison of test data with
predictions (from (17.4) and Chapter 4 of (6.31): (a) circular holes, (b) square holes

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
the experimental results are compared to predictions by a finite-element method, based on
that of Yall!g [ 17.24]. Here k, is the buckling coefficient of the plate with a square hole and
k", that of the corresponding unperforated plate.
A similar presentation of the influence of circular and square central holes on the buckling
of simply supported and clamped square, mild-steel plates subjected to shear is depicted in
Figure 17 .2, based on tests carried out by Narayanan and Chow at University College,
Cardiff, in the mid-eighties (17.1 2]. In the figure Ac is the area of the central circular or
square hole, A is the area of the unperforated plate and K~ and K? are the shear buckling
coefficients of the perforated and unperforated plates, respectively. A comparison of Figures
17.1 and 17.2 shows that the reduction in buckling load due to holes in the case of shear is
nearly twice as large as in the case of axi.al compression. It should be noted that the exper-
imental values for compression, in Figure 17.la and b, were obtained in short-column tests
and beam tests on typical structural elements (see [17.4]), whereas those for shear loading
in Figure 17.2 (see [17.1 2]) were obtained on plates fixed in a typical picture frame test rig
(discussed in Subsection 8.4.2 of Chapter 8). However, comparison of the results of Figure
17.2 with those observed in shear tests on beams (see for example [17.4]), indicates that the
reduction in critical shear stress is of similar magnitude in both types of shear tests.
For all the loading cases, most of the earlier experiments and analyses on perforated plates
considered square plates, and only more recent tests also included rectangular plates (see for

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Effect of Holes and Cutouts on Plates and Shells 1415

t ·

0.9
e

0.8 •
CD

~- t:· 0.7 • 0

., .,
~ ~ •
~ 0
~ " •
()_6
~
cd
§ 0.5
... ... o ..

.,""""
..li .,..
~
0.4
~-
~
~" ~
03

02 ·
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

0 1


0 0.1 02 OJ 0.-4 05 06
Ac
A
Figure 17.2 Buckling of clamped square steel plates with central holes subjected to shear-influence
of hole size and shape on buckling coefficients and comparison of test data with predic-
tions by the finite-element method or an approximate empirical formula (from r17 .12])

example [17.4], [17.11] and [17.19]). In rectangular plates, however, the increase in hole
size may caUJse a change in the buckling modes. For example, in the axially compressed
rectangular steel plates of aspect ratio 2, tested at Strathclyde University, Glasgow (see
[17. I I]), circular holes whose diameter exceeded 0.35 times the p.l ate width precipitated a
change in buckling mode from two to three half-waves.
As emphasized in Chapter 8, plates, because on account of their stable postbuckling be-
havior, usua]ly sustain loads considerably in excess of their initial buckling loads before
failure. This applies to both axial compression, bending, shear or combined loadings. Stiff-
ened plates therefore collapse, in general, at loads significantly above the local buckling o f
their plate e lements. The influence of holes on the postbuckling behavior and ultimate
strength of plates is therefore of primary importance.
In the postbuckling range, plasticity and material properties affect the behavior of plates
as well as the influence of ho les on that behavior. See for example the collapse modes of
square perforated steel plates subjected to uniaxial compression (Figure 17.4 and !17.26]).
Hence, recent studies on the effect of holes on the ultimate strength of plates sometimes
employ rigid-plastic mechanism solutions (see for instance Figure 17.1 8 and [ 17.26)) and
often include also comparisons between metal and composite plates (see for example [ 17.20]
and [17.25]).
Figure 17.3b shows the influence of central holes on the maximum load sustained by
metal and composite plates subjected to axial compression. For comparison, the influence o f
central holes on the buckling loads of the same plates is shown in Figure 17 .3a. Both plots
also represent a comparison of this influence for metal plates, here made of 2024-T3 alu-

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1416 Influence of Holes, Cutouts and Damaged Structures

::i.-.i_ .........
1. 1

_R_ •. 8

P.. •. b

•. 4 ..... "&'- .iutlmJM t !P•o ..· ·JJso 1bs 1 ·

0 2
-13 ..
G~iEPOXY I :.P..~.~ : ._3150 lbs I
-A- GRIPfH ti Ppo • 75li I bs l

8. I a. 1 e.J 1. ~ ~.5 &. b f. 7


HOtE D!A. I PLATE WIOHI
(a)

t.Z

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
El

.. ~ ... . .. ....... ' .. ..

~-e- AIIJM[NU/1 1 Pu! o • \8. \'39 lbs l


e. 2
' -& GR,EPOXY I p,;~u • 2&. 546 l bs l
' .
:-C,.- GR/P(EK I Pupo • I 0. l.11 I bs l
4-~~~~~~-r-·~~--r,-~-~~~,-~~
I. I 8. 1 I. ] I . ·4 i. 5 8. b II. I
HillE OIA. I PLATE 1~; Ill! I
( b}
Figure 17.3 Wichita State University experimental studies on the influence o f central holes on the
buckling loads and maximum sustained loads of metal and composite square plates sub-
jected to axial compression (from [17.20]): (a) influence of hole size on the buckling
loads, (b) influence of hole size on the maximum load sustained by the plates. The plots
are average values of three panels of each material

minum alloy, with that for two types of composite plates, one made of graphite/epoxy (AS4/
3501) and one of graphite/PEEK (AS4/PEEK). The graphs are reproduced from [17.20]
and [17.25] (the latter is essentially an updated version of the former) and are based on tests
carried out on 45 square plates (15 of each material) at the National Institute for Aviation
Research of Wichita State University, Kansas, in 1989. The aluminum alloy plates were
made from commercially available 0.08-in. ( - 5.4-rnm) thick flat plates, whereas the com-
posite test plates were laid up to form symmetric 16-ply laminates with a stacking sequence
of [45/90/ - 45/0/45/-45/45/ - 45].. The dimension of the test section was 10 X 10 in.
( - 254 X 254 mm), and the loaded edges were clamped while the unJoaded ones were simply
supported.
The centrally located circular holes were found to influence the buckling characteristics
and ultimate strength of the metal and composite in a similar manner. For all three materials
the buckling loads (see Figure 17.3a) first slightly decreased as the small holes increased in
size, but then beyond hole diameter-to-plate width ratios (dlb) of about 0.4 the buckling

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Effect of Holes and Cuto uts; on Plates and Shells 1417

load tncrca,ed wuh hole ,,,e, e\,ccuing thai of the ~:urrc,pondwg unpcrfor.ucd plate. The
effect of hole size on the ultamatc load of the p<UII!), hcc Figure 17.3b) wa-. ugutn '>imilar
for the three material-;. TI1e ultimate load remained fnirly con,tant up to .u hole dtumeter-to-
width ratio (t//b) of about 0.5 und then decreu~ed with mcrca~ing hole size. But it <.hould
be remembered that because of the Mable postbud.ling behavior of plates. the ultimate loads
arc abuut four to eight time' their corresponding huci.Jing load'i. with the htghcr ratio
rclatmg to the <.muller holes
Though the general mnucnce ol central holes on the hud..ling and uJtimatc load<; of eom-
p<Ntc plate.:' i-. 'imilar to that uf mctaJ one!>. thctr po<.tbud.hng beha\ ior al\o 'tron!!IY de-
pcnch on the plate onhmmp) hec for example (17.2211 b~pcnmemal re ull\ mdtc:llcd that
mo't ol the compo~itc plate' tC\tcd ""ith centr.ll cutuut' c\htbllcd le'>s poo;tbuclhng \ltffne<;
than the corresponding un~rlorated plates. llowever. \Omc of the highJy onho troptc com-
p<l,IIC plates with holes (te,h.: d in (1 7.221> showed htgher pmtbuckling_ :-.tiiTncsll than the
corre,ponding unpcrfor:ucd phlle,, i.e. the complex interaction of cutout ~• t.c und phttc or-
thntropy on the postbuckling load distribution couiJ 111 certain cases lead to an increase in
po'tbuclJing. ·trcngth. Therefore. the cutout 'i7e and the <.t,tcking ~cquence ldctem1ining the
mthotrUp} l could be tailored Hl orumt/C the po\tbuclltng ~tiiTne"s of perforated compo,ime
plate\ runher thcorcucal and e'perimental \IUd) OJ thl' modaJ interaciJon. hO\\C\Cr. i
needed for reliable dc:,tgn method\
A' ha' been evident from thc dt,CU'>'>ion in Chapter 1-1. the comprcso;ion bcha\tor of
lanun.tted composite panel,, without and wilh hole~. i' 'trongl} dependent on the dctaiJ.. of
the lay-up and the resultant ani,otropy. It has therefore been 'ubject to many tnvc~ugatton<,
but ~till insufficient expcrimcruul study. in particular lor curved panels (sec Section 14.3 a nd
117.13).117. 16] and 11 7. 171).
One may further note that where:." the local buckling lnad for a perforated 'tiffcncd plate
\tructurc '' often more aflccted h) a 'quare hole th.m b)- .1 circular one (a, indtcated in
1-tgurc 17.11. the postbuclhna; o,trcngth of perforate(.! 'tntcturc with a <.quare hole and a
ctrcular one were found to be ncarl} the .;arne tf the dtJmctcr of the circular hole equal!. the
"tdth ul the 'quare hole ('>cc 117.4 )I
Funhem10re. it should <ll'n ll\: pmnted out that though mu'l tc 1 on perforated plate<. dcaJt
wuh central holes, Lhc innucncc of eccentric cm:ular and 'quare cutout~ on the bucllmg and
ro~thuclllng hchavior of plat~' wa-; al\o in"c~tigatctl hy \omc rcsearchc11- (sec for example
Figure 17.5 for lhc case of :-.hc;lr loading and 11 7.12). 117. 16 I aml l J 7 .26J).
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Fi~:urc 17.4 Buckled ~qunl'\: 'led plutc' wilh central!) localcd holl.:'. lc~ICd 10 latlurc nt Untvct-.;lly
College. Cardiff (frum 117.261. courtc~y of Profc,,or R. Narayanan). Nmc th.JI the 1mlure
paUcm \Uj!jlC'l' n 'IOlfliC jlJ.t,lit: mcchallt\lll

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1418 Influence of H oles, Cutouts and Damaged Structures

..
~.
0)

(a)

(b) (c)

Figure 17.5 Ruck led square steel plate~. "nh ccccmncally located ctrcular hvle,. ~UbJCCted to ~hear
and tc\ted to failure. as ob~el"\ed an the Unilcr,it} College, Cardiff. e., pcrimems (coune~>
of Profe~sor R. 'arayanan} (a) plate\ \~1th one eccentric bole in tcn.,ion diagonal. (b)
plate \\ ith two eccemric hole'> 111 tcn,tnn diagonal. (c) plate "uh l\\0 cccenltlc hole,. 10
compre,\ion dtagonal

17.1.3 The Effect of Holes and Cutouts in Shells


In the case of shells. contrary to that of plates. the load imperfection which the cutout
represem~ mny ~ignilicantly affect the buckling load hccau~e of their imperfection sensitivity,
unless the hole is very minute indeed. The influence of an imperfectilln in the fom1 of a
cutout on the buckling of cylindricnJ shells under different types of londing has therefore
been tudicd by l>everal investigators (see for instance 117 .271- 117.411).
The comprehensive experimental and anal} tical \tudy by Stame~ at Cahech in the late

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
ixtics ( ec 117 .38]. [ 17.30] and [ 17.3 1)), and then at 'ASA Langley in the early \cvemies
(!lee [17.32]), ~ummarize the knowledge of the innucncc of unreinforced cutouts on lhe
buckling of cylindrical shelL in the seventies. llio; e"<pcriments co,crcd a broad range of
circular cylindrical shells. with (R/f) - 400- 800. and cutouts wuh (r/R) 0-0.5 (where R
is the radius of the shell. r is the radius of the circular hole. or a similar charactcri~tie hole
dimension for other shapes. and 1 is the wall thickness of the shell). which were ~uhject to
axial compression, bending or torsion loading (sec 117.38] and [17.32)). The test specimens
were made of Mylar polyesner film. were 8 in. (- 203 mm) in diameter and 10 in. (- 254
mm) long ami had a wall thickness 1 - 0.005- 0.0 I0 in. (- 0.13- 0.25 mm).
The ~hel h were loaded to buckling before any hole!> were cut into their ~ideo; in order to
provide reference buckling loads. A ~eric~ of concentric circular. square or rectangular cut-
out wa<> then cut into the cylinder wall at midlength. and the !>hell bud..ltng load was
mea!>ured for each cutout size and dec;ircd loading condition. Thi' procedure "a" continued
until the large. t desired hole wa), cut into the \hell and all desired loading conditions had
been appltcd. Buckling was detected by 'isible and audible means.

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Effect of Holes and Cutouts on Plates and Shells 1419

The results of the experimental and theoretical studies revealed that the buckling loads
were related! to the nondimensional geometric parameter
r= (r/VRt) (17. 1)

r was therefore used as the governjng con·elation parameter.


Central axial compression loads were applied to 16 shells with circular holes in their sides,
and the buckling loads were measured for increasingly larger cutouts. The effect of the
cutouts is summarized in Figure 17.6 by the curves representing the boundaries of the ex-
perimental results. The measw-ed experimental buckling loads, P, are given as a function of
r and haVe been nondimenSiOnaJized by the claSSiCal theoretical buckling load, p CL• for a
cylinder without a cutout, to provide a constant reference for comparing results from different
shells. The cylinders tested included specimens with ( Rit) = 400, 533 and 800, and the
results of a particular cylinder with (Rit) = 400 are shown by the two symbols in the figur,e.
The dashed curve depicts the results of a finite-element analysis with the NASTRAN code .
Starnes pointed out [ 17.32) that one can distinguish between four ranges of r according to
their different buckling characteristics:
I. "For values of r less than approximately 0.5 a circular cutout had no appreciable effect
on shelLs loaded by central axial compression. The stress concentration due to the holes
in this range of r is apparently too small to cause buckling before the shell buckles .into
the general collapse mode due to some other imperfection. The buckling mode for this
range r was always the general collapse diamond pattern with the hole randomly oriented
with respect to the diamond ridges.
2. For values of r between approximately 0.5 and 1.2, the buckling loads always dropped
sharply as r increased. The general collapse d iamond pattern was still the mode of
buckling, but the hole was located on a ridge of the diamond pattern or the intersection
of these ridges. As is well known, a cylindrical shell becomes sensitive to the slightest
disturbance when the stress level in the shell is near its buckling stress. Apparently, the
prebuckl ing stress concentration around the hole in this range of is of sufficient mag- r
nitude to cause the hole region to snap into a local buckling configuration. This local
snap buckling could, in turn, provide enough of a disturbance to the shell at these high
applied stress levels to cause general collapse of the shell. [Indeed , the imperfection
sensitivity of the shells was clearly demonstrated in the experiments.]
3. Values of r between about 1.2 and 2.2 represent a transitional region, where the character
of the effect of cutout size on buckling load changes from the sharply declining general

.9 I
I
I
I REPRESE~TAT I VE
SHELL WITH RA • 400
\ ----NASTRAN 0 LOCAL BUCKLING
1 RESULTS t> GENERAL COl lAPSE
I
I
l> I
\
\

80\JNOARIES Of EXPERIMENTAL
.4 RESULTS FOR 16 SHELLS
J 0 0
.z
.1

Figure 17.6 Cahech tests on Mylar ciJ·cular cylindrical shells- the influence c,f circular cutouts on
buckling under central axial compression (from [17.32)). Note that from the transitional
region (r = J .2- 2.2) and beyond, there appears first a stable local buckling mode (shown
by the symbol Q ). which only after a further increase in load turns into a collapse mode
(shown by the symbol L::.)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1420 Influence of Holes, Cutouts and Damag~ Structures

collap c hucl.hng loads. ob~encd for ~tween 0.5 and I 2. to the much rmldcr additional
decrease rn hucl.hng loads found lor r equal to ahcmt 2.2 or greaiCr. In thr' tran~itional
range a 'hc11 would either bucl.lc mto the general co11ap'c diamond p.:utcm or a stable
h.xal bucl.hng mode located around the hole. The ptl'thucl.hng fonn of thr' 'table local
l'luckhng mode [ hown in Figure 17 71 W.ll> appmximatcl) that of an e11ip'c The senu-
major .rw. of thr<. ellipse imuall) appeared to be cuhcr tangent to the 'hc11 crrcumfercnce
pa..smg through the bole center or along a ~I.e\\ or hchcal d1rcctJOn, and then rotated
IntO the ClrcumfcrcnuaJiy tangent p<>'>lllOn. The ma\llllUill depth o£ till' lOC<II buckling
mode occurred at the cutout edge" [and wal> uwu11y on the order of 25· 50 time~ chc
shell wall thr cl.ncs~]. "Whenever till\ stable /om/ lmcWnx occurred. additional load
could he apphcd to the shell cau,mg the length ol chc cllip\e ~cmi major .1.\l'l to incrca_..c
unulthe general collapse occurred. In the range ot r hctwccn 1.2 and 2 2. the difference
between the bucl.hng loads ass~X.ratcd w nh the 'table local and general collap'e buckling
modes ".1' U\ually larger than for r equal to 2.2 or greater.
..t. For value.\ of r Xft>ater thtm 2.2. the test hell\ <tlwuy' snapped 1nto the ~table local
buckling mode which was fo11owcd by gencml cnlh1p1.c when additionol load was ap-
plied. Appan:ntly. the stre s concentration around <1 crrcular hole rn thl' rangl! of; i of
-.uflictent ma~nuudc to cause hlCal buciJrng to occur ~fore the apphed load \\.C. large
enough to rai'c the heU''i gcncr.LI ''re" le,cl to a \Jiue that would m.tl.e the 'ihcll
\Cn~iU\ e to a ~eneraJ collap..e cau rng dJ,turbancc ~rnc.:c the hell wa~ no longer sen-
\lllvc to latcr.tl di,turb:mces once th" 'wble local l'luc.:l.hng had occ.:urrcd. the concen-
tr:ued strc~~c' around the hole were apparently reduced rn magn1tudc a' a result of a
•MI!'>~ rcdt'. trrhUtJon. Buckling load~ conlmucd tu dcchllc "' th1'i r.mgc of ,., but the rate
r
of decline wa.. much le. than for between 0 5 .tnd I 2. \\here a \Cry \harp decrca: c
wa'i ob..ct'\cd ..
One hould note that for small hole!. the ctreular cutout at:t<; ao; a fonn ul 1mperfecuon.
\\hcrcas for la'l!cr hole' the cutout change' the bucl.ltng mode to a ... rahle local buckling
mode, which only after additional loadmg tum'i into wllap ...c The actual rcduc.:uon 10 bud-
ling load due tu the hole is, howe\er, very "gnificant (up to 65 percent). Starnc...' 1.ummary
cmpha'ii7e'i the tran,itiun (see Figure 17 6) from one type ut behavior to the other (the stable
local buciJang 'hown in Figure 17 7) wtth incrca.~e tn r II may further be pomtcd out Lhar
u \lmJiar tran 1tion and large reducuon' rn buckling ltMd Cup to 55 percent) \\ere obsened
h) Tcnn)o;on rn ht\ "pun-cast cpo\y Circular C)hndncal 'hell ( cc [12.271>. foda [17.36]
and [ 17.37] oh,ervcd a !>imilar behav1or 1n lm Mylar poi)'C\tcr ~hell'i \\ 11h c1rcular cutouts.
Funhcm1orc, 111 a parallel study on the effect of elliptic cutout<;. Toda oh,crved a similar

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

tl~tu~ 17.7 Sralllc loci.ll buckhng modi: ftlf 11 .:•n:uiJt ~.:urou1 rn .a ~:u'l:ular qhndmal 'hell loaded b)
ccnlr.ala\ial compres~1on. 3\ ob-..:J\o:d in 1hc C';~llc.-~h '"'''' 1-..:e 117 ':!II

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Effect of Holes and Cutouts on Plates and Shells 1421

behavior and in particular a similar stable local buckling mode for large elliptic holes (see
[17.35]).
Starnes also investigated the effect of eccentrically applied axial compression loads on the
buckling of cylindrical shells with circular cutouts (see [17.32]) and found that in this case,
r
too, four characteristic ranges of could be defined in a manner similar to that in centrally
loaded cylimders.
The influence of circular cutouts on the buckling of circular cylindrical she lls loaded by
torsion or combined central axial compression and torsion was also included in Starnes'
investigation (see [ 17 .32]). The effect of the cutouts was again a decrease in the buckling
r
stre ngth of the shells as increased, but a s maller one than for axial compression loads.
The effect of square and rectangular cutouts on the buckling of cylinders loaded by central
axial compression forces is another topic studied by Starnes (see [17.32]). The effect is
summarized by the results of 12 test cylinders which are bounded by the curves in Figure
17 .8. These cylinders had (RI t) = 400, and representative results for a typical shell with
square cutouts are also shown in the figure by the two indicated symbols. one for the stable
local buckling and one for the collapse that follows after further increase in load. The dashed
curve in Figure 17.8 again represents the results of a finite-element analysis with the NAS-
TRAN code.
Most of the cylinders tested consisted of specimens with increasingly larger concentric squares or
rectangles cut into their sides, and a limited number had both square and rectangular cutouts with
a common center. One shell was tested with both circular and square cutouts with common centers,
where the square side lengths were usually equal to the circle diameters. The rectangles cut into
each shell were oriented with their long sides either in the shell longitudinal or circumferential
direction, with a given orientation maintained for a particular shell. The rectangles had aspect ratios
ranging from approximately 1.2 to 2.0 for each configuration orientation, with the majority equal
to 1.5 or greater.
r
The characteristic hole dimension in the correlation parameter was taken to be equal to one-
half of the side length for the squares and the average of the half-side lengths for the rectangles.
Even for such diversely different configurations as longitudinal and circumferential rectangles with
aspect ratios equal to 2.0. all of the experimental results for the cylinders with square and rectan-
gular cutouts fell within the relatively nan·ow scatter-band [of Figure 17.8] when plotted with
respect tor.

From these results it was again possible to identify approximate ranges of with different r
buckJing characteristics.

r
For values of less than approximately 1.2, the buckling behavior of the cylinders with square and
rectangular cutouts was much the same as the behavior of cylinders with circular holes [see Figure
17.6]. For values ofr greater than about 1.2 the buckling loads continued to decrease with increas-

REPR£SENTAIIIJ[ SHELL
.8 WITH SQUARE CUTOUTS
0 LOCAL BUCKLING
t>. G£~[RAL COLlAPSE

p
Pet
.4

.2

Figure 17.8 Caltech tests on Mylar circular cylindrical shells-the influence of square and rectangular
cutouts on buckling under central axial compression (from [17.32)). Note that here there
is no transitional range, but otherwise the behavior resembles that caused by circular holes
(shown in Figure 17.6)

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1422 Influence of Holes, Cutouts and Damaged Structures

ing r, but at a much milder rate of decline than that for the small values of r. For rectangular and
square cttlOrtiS there was no rransitional range as there wa.~ for circular cutouts, and the general
collapse of tire cylinders was always preceded by a stable local buckling mode with an inward
postbuckling deformation panem. The stable local buckling mode wa.~ followed by general collapse
r
when loading was continued. For between approximately 1.2 and 1.6. the Stable local buckling
mode always occurred in a symmetrical pauem approximately in the form of an ellipse with its
r
semi-major axis tangent to the shell circumference. For larger than about 1.6 the symmetrical
stable local buckling mode was usually preceded by an asymmetric mode aligned with one of the
cutout diagonals which either snapped into its postbuckling form or just began growing inward as
a large local deformation. For square and longitudinal rectangular CUtOUtS with greater than r
approx imately 2.5, noticeable symmwic outward prebuckling dcfonnation of the longitudinal cut-
out sides preceded the above asymmetric deformations. For <:iJcumferemial rectangles with greater r
than aboUit 3.2. noticeable symmetric outward prcbuckli ng deformations of the circumferential
cutout sides preceded the asymmetric patterns. In all cases, the difference in the load~ for the first
noticeable local buckling and general collapse was small.
r
For less than approximately 4.5, there was no detectable difference (beyond normal experi-
mental scatter) between the general collapse loads for longi tudinally and circumferentially oriented
rectangles for the limited number of cylinders tested with each cutout configuration. For greater r
than 4.5, the general collapse loads of the circumferentially oriented rectangles tended to occupy
the lower ponion of the scatter-band [in Figure 17.8). This phenomenon appears to be a result of
the large prcbuckling deformations and stable local buckling mode for the cylinders with circum-
ferentially oriented rectangles, which cause extensive stress redistribution to occur farther around
the cylinder circumference. This would bring about a general overall weakening of the shell at
lower applied loads than for square or longitudinal rectangular cutouts.

The detailed observations of the different buckling c haracteristics have been reproduced
here at leng th since they emphasize the physical influence of cutouts on the buckling modes,
which are responsible for the ch anges in buckling loads.
'·Generally, for a given value of r,
the gene ral collapse loads for cylinders with sq uare
and rectangular cuto uts were found to be slightly lower than those for cylinders with circular
cutouts." In Figure 17.9 the collap se loads for similar circ ula r and square c utouts (with the
half side len gth of the square equal to the radius of the circular ho le) in the same cylindrical
shell are compared as an example. The differences are shown to be indeed s mall, with the
reductions due to square cutouts being slightly larger than those due to c ircular c uto uts.
Toda [ 17 .39] compared the results obtained by Mylar specimens with those on metal shellls,
whereas Starnes also tested s ix electrofonned copper cylindrical shells w ith circular cutomts
(see [ 17 .3 1]) and found the results to compare well with those of similar Mylar sheDls.
Furthermore, Gavrish et al. [ 17.40] carried o ut a comprehensive test program on 200 rela-
tively thick steel shells with (Rit ~ 60) to extend the results to plastic buckling . The con-
clusions on the influence of unrei nforcecl cutouts were also ex tended to composi te shells
with the test carried out by Schulz on glass-epoxy cylinders at the Tec hnical University
Karlsruhe (see [ 17.34]).

.s

p \ CIRCUlAR CUfOUTS
PCl .4 ' 0:

.2
[ sQ~ARE c.;;~;s--- ----

Figure 17.9 Caltech tests on Mylar circular cylindrical shells-the innuence of circular and square
cutouts on the buckling under central axial compression of a typical shell (from [17.32))
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Eff&ct o f Holes and Cutouts on Plates and Shells 1423

In the early seventies a series of buckling tests on axially compressed 6061-T6 alurninu m
cylindrical shells with large rectangular cutouts wa conducted at the Lockheed Palo Alto
Research Laboratory (see [ 17.29]). This wa~ an excellent example of a careful experiment
closely correlated with a nonlinear computer analysis. Details of the tests are discussed ~ n
Section 17.3 below. Here it is only pointed out that both unreinforced and reinforced cutollts
were investigated. wilh an a. sessment of the reinforcernenl benefits.
The studies of the influence of unrcinforccd cutouts on the buckling of metal and com·
posite shells have continued in the last decade. For example, extensive te t were carried out
at J ASA Langley Research Center on graphite-epoxy cylindrical shells with circular cutouts
(sec Figure 17.10 and 11 7.42]). Another experimental program at NASA Langley Research
Center included severaltesL~ of compression-loaded 6061 •T6 aluminum singly curved panels
with a central circular cutout (see [ 17.43]). in which the buckling behavior was carefuliy
monitored with the aid of the corrc ponding shadow moire pauerns.
Whereas the earlier investigations (such a [17.27], [ 17.30]-[ 17.35) and [ 17.38)-[ 17.41 D
approached Lhc effects of holes as a form of imperfection, later studies focused mainly on
Lhe effectiveness of reinforcement~ in alleviating the reduction in buckling load caused by
the cutout ( ec for example I 17.29]. [ 17.34), [ 17.37) and [ 17.44]- 11 7.48]).

17. 1.4 Reinforcements


The focus on the effectivenel> · of Lhe reinforcements of the cutouts was indeed logical. since
whenever cutouts were required. the designers u ually tried to compensate for the resulting
reduction in buckling strength by suitable reinforcement of the holes.
A good example is the passenger window of a modem commercial airplane (see Figure
17.11). Here the essential cutouts. the windows, are reinforced by relatively heavy window
frames wilh the aim to annul the reduction in buckling trcngth, caused by the windows.
The test specimen of the window belt hown in Figure J 7. II is similar to Lhe corresponding
ponion of the Boeing 777 fuselage and includes window cutouts, stringers and frames. The
test specime n is loaded by various combinations of compression and shear. The window
reinforcements were here relatively heavy forgings. but at maximum combined loading Lhcy
experienced substantial twisting. After removal of the load numerous diagonal shear buckJes
remained throughout the panel.
Another example is the tests on reinforcing of openings in civil engineering shells sub-
jected to axial compression and combined loading. carried out at the Chicago Bridge and
Iron Co. in the early eighties [1 7.471. The test specimens were 15-in. (- 381-mm) diameter
Mylar shells. (sec Figure 17.12) and 18-in. (- 457-mm) diameter Lexan (a plastic material
similar to Pe rspcx) shells. The ~pecimens were reused with different opening and reinforce-

Figure t7.10 NASA Langley graphite-epoxy circular cylindlical ~hell ~ with unreinforccd c ircular cui-
out for investigation of the innucncc of cutouts un the buckling !-.trcngth (coune;y of
NASA Langley Research Center)
--`,`,`````,`,````,,``,``,,`-`

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1424 Influence of Holes, Cutouts and Damaged Structures

Figure 17. I J Typical p:t''CII!!Cr winc..low~ of a modem commcn:tul atrplane. the Bocmg 777 The te\t
specimen t>f a fuii·\C31c "''"do"' belt panel. "h1d1 mcludc~ windll" ~:uluul' 'l!.llh rein-
fom:mcnl rr.unc... 'tnngcr\ :llld frame' and tllJt "'a' loaded b) rornbmcJ ~:~lmpre,\ion
and ~hear tcoune'y ut tht: Boeinl! Commen:t.J \trplanc Compan) I

men!\ to permit pammclric 'ludic~. The mol.ltntere,llng ·''('Ccts 'tudicd there ~ere the effect~
on buckling of adding rctnfon:cmcnt~ to the cylinder \\:til both nom1al <and p<amllcl to the
axts. the influence of MiiTcncrs rmher than pads or ned. reinforcements nnd the influence of
combmcd axial load~ •Uld cxtcmal pressure on the effect~ of reinforcements.
A further example i' the extensive test~ conducted hy Schul£ at the Technical Uni\crsity
Karlsruhe f 17.34] to evalut\IC the cffectivencs!> of reinforcement in his perforated glas •
epoxy cylindrical . hcll<,. llt' carefully controlled cxpcnrncnt<., (<>cc Figum.• 17. 1'l)mcludcd
round. '>quare and diamond-,hapcd cutou~ and cm·en:d a \Hdc range of hole i1c' (lrl Rl =
0.286 1.285. "here R i~ the radtu' of the cyhndcr •.m<.l r the characteri"illC hole t.limenc;ton
(e.g. the radtus of a ctrcular ~utout).
A~ m the CBl tc 1~ in Chicago, the Karbruhc 'hell"' were fir..t carefull) hud.lcd unper-
foratcd in axial compres~ion. l11c load was released and a required hole cut. with the cylinder
slightly loaded in the test rig to prevent any shifting in position that might alter the hounclary
c<>nditions. Then, after rcpc:ucd huckling. the rcinfurccmcnt was glued on and the ~hell
buckled again. Next a larger hole was cut and the procedure repeated. l n thi' manner u,cful

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
parnmctric de!>ign infom1ation w:l\ obtained. in addition to the hac;ic 'ludy of the behavior
ol the !>hells.

(a) (b)
l"i(lurc 17. 12 Chicago Bridge nnd Iron Co. tcsh on circulat cyltnc..lncal Mylar ~hclh. with •cmforccd
circular cutoul\, lnnded by axial compre>~•on (courtC\)' of Dr. C.D. Miller): (a) bud..lcd
shell Wtlh frame nr neck reinforcement amun<l cutout. (b) ring· \liffcncd bud.lcc..l \hell
with longitudmul Jnd ctrcumfcrential rcmfnrc~m~nls nt•:u cutout

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Effect of Holes and Cutouts on Plates and Shells 1425

Fij!ure 17. 13 Karl,ruhc cxpcnmcnl\ on the mllucncc of Circular. d1amond-shaped and square cutouts
on the buckling of gl:t\~·fabric reinforced Jl(llyc"cr \hell' \Ubjecled 10 axial compression.
The tc'" alw c' aluatcd the cffccti,cnc\~ of glued-on reinforcement frames. The bole
we' co,crcd the range (2r/R) - 0.286-1 285 (from [17.34)1

Al~o. for design mfonnation. another practical problem. the effecti\ eness of a.'(ial and
circumferential reinforcing stiffeners for cylindrical \hell~ with fai rly large rectangular cut·
out\ again~t buckling under axial comprer;<;ion or bending, was swdied by Jung a1 the Fach-
hochschule Aachen in the late seventies (see [ 17.34)). He concluded that if longitudinal
\tiffener\ of sufficient length were provided, thi~ reinforcement sufficed to compen ate for
the hole and resulted in practicaJly no reduction in the buckling capacity of the perforated
'hell.
In practice. however. large cutouts usually occur in l>hell!. Miffened by rings and stringers,
:mc.J they also cut these stiffeners. Palchevski and Polyakov 11 7.451 studied this problem for
closely minger-stiffcned. welded aluminum alloy shells (with Rlt = 400) under axial com-
prc~~ion in order to assess the relative effectiveness of the different reinforcements. They
concluded that stringers arc more ctTcctivc than rings but that combined stringer and ring
reinforcement is most effective. which agrees with resu Its of earlier optimization studies for
~t iffcned ~he ll ~ without cutout.
The effect of cutout reinforcements in clo~ely ~tiffened circular cylindrical shells. where
the holes al~o cut the stiffeners, were al~o <>tudied in the late seventie. by Cervantes and
Palatotto at Wright-Patterson Air Force Base. Dayton. Ohio (~ee [1 7.48)). Theirs was a
numencal optimization study with the STAGS A code [2.53]. an example of what i'l usually
tenned IIIIIIU' rtcal e.,perimems. Though the e were an cxten~ive parametric stud}. they can
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

onl} serve as a prelimin3I) indication. and additional physacal experiments are rcqu1red 10
validate their conclusions. This limitauon m mmd. the main conclu~ion in this cao;e (17.48j
were:
"I. A reinforcing fmme equal in volume to the material remo,ed for a cutout and placed
adjacent to that cutout, can increase the buckling load of the shell to or above that of
a 'hell without a cutout.
2. l-or either a stringer or a ring and stringer qin'cncd cylindrical shell with rectangular
cutout'>, the optimum position for a reinforcing frame is along the cutout edges. Yet.

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1426 Influence of Holes, Cutouts and Damaged Structures

the designer must be aware that in certain situations a frame positioned far removed
from nhe cutout edge may reduce the buclcling load below that of an unreinforced shell.
3. Positioning a frame away from the cutout edge, in most cases, drastically reduces the
buckhng load from that obtainable by placing the same volume of reinforcing frame
adjacent to the cutout.
4. For very light frames there may be a frame position away from the cutout edge which
is equally effective as the position adjacent to the cutout."
Actual experiments (see for example [17.44]) have indicated that cutouts with appropriate
reinforcements can indeed be designed Lo cause only minor reductions in the buckling
strength of stiffened cylindrical shells (see for instance Figure 17. J4 ).
The study of the effectiveness of cutout reinforcements led to attempts to optimize them.
These optimization efforts were accompanied by significant advances in the measurement
and calculation of stress concentrations (see for example [ 17 .49]), and considered not only
buckling but also strength, fatigue and fracture mechanics.
One of the early approaches to optimization of cutout reinforcements was the neutral holes
in plane sheet. Neutral holes are reinforced holes that are elastically equivalent to the uncut
sheet for one type of loading system. They were introduced by Mansfield at the Royal
Aircraft Establishment (R.A.E.), England, in the early fifties (see for example [17.50]-
[17.52]). Att the time, these investigations received further impetus by the conclusions of tihe
famous, wide-ranging Comet accident inquiry that " the cause of the accident was the struc-
tural fai lure of the pressure cabin brought about by fatigue" (see [17.53] and [17.54]) and
that the failure originated at the windows. Though most of the benefits of the lessons learned
from the Comet accidents went to our understanding of fatigue, optimization and design of
reinforced cutouts against buckling was also a notable beneficiary.

17.2 Experiments on Plates with Holes and Cutouts


17.2.1 Metal Plates and Webs
The earliest experimental study on the influence of cutouts on the buckling of thin plates
under unidirectional compression was probably that of Kumai at Kyushu University, Japan,
in 1951 [17.55]. This investigation included tests on 30 plates, with hole diameter-to-plate
width ratios of 0.1-0.65, and were accompanied by an analysis with an energy method.
Professor Kumai pointed out "that since no experimental data are dealt with in both (then
available) investigations [17.56] and [17.57] it is very ambiguous to apply the results in

96.5"1·

7,,,.,. §8Q'/2:.o:~
§
1118 "/.

a:!;; a § ,-cu10UT

0 -~
~
-~ - -
...12 IS
rr

z
2 w (
43.8'/.

-~
':(
•-
-"''i' -
0:
"'cz "'
w <( 1 <(

§
~---~ c-il
0
_! 0
-~
If

0
-~
0:
"'u.
0
- ~
If
...
0
OQJIUM

5 w
l!l w
"'00 0
w
~'
::> <( ~
0
• 8<( ~

Figure 17.14 Effects of a cutout and different reinforcements on the buckling behavior of a Lex.an
isogrid model, representing a McDonnell-Douglas Astronautics Company Thor Delta
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Launcher interstage cylinder (from [17.44))

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Plates with Holes and Cutouts 1427

practical design." He then embarked on his careful experiments, comparing their results with
his calculations and those of Levy et al. [ 17.57]. He obtained good correlation bet ween his
experiments and calculations. His main conclusions were that whereas in simply supported
square plates with a central circular hole the critical load ratio decreased monotonically with
diameter ratio, in similar clamped plates a minimum critical load was reached at a diameter
ratio of 0.2, after which the critical load again increased, eventually beyond that of an
unperforated plate. One should note the reciprocal reliance on theory and experiment in this
study.
From the sixties onward, extensive experimental studies on the buckling and postbuckling
of perforated metal plates we re carried out at many research centers in several countries.
Some typical examples are investigations conducted in the sixties by Schlack at the Univer-
sity of Wisconsin [8.54] and {l7.8]; the tests performed by Yu and Davies at the University
of Missouri-Rolla [17.4] in the early seventies; the studies carried out by Ritchie and Rhodes
at Strathclyde University, Glasgow in the mjd-scvcnties [ 17.11]; the tests on the ultimate
load-carrying capacity conducted by Narayanan and Chow at University College, Cardiff, in
the eighties [17.26); and the more recent experiments of Nemeth at NASA Langley [17.58].
These five programs considered uniaxial compressive loading, whereas other parallel inves-
tigations dealt with shear loading, as for instance the shear tests on perforated plates carried
out at University College, Cardiff, by Narayanan, Rockey and their co-workers in the eighties
(see [17 . 10], [ 17 .12], [ 17 .59] and [17.60]), and the experiments on girders with holes in
webs, conducted by Ravinger at the Institute of Building Constructions and Architecture of
the Slovak Academy of Sciences in Bratislava [17. 19] in the same period. The more recent
investigations for both loading cases focused on the postbuckling behavior and ultimate load
of the plates with cutouts.
Schlack's exper.iments in the sixties [8.54] and [17.8] were widely quoted since they were
carefully executed and concentrated on the influence of large holes, up to a diameter of 0.7
the width of the square plate. As pointed out in Subsection 8.2.2 of Chapter 8, a series of
independently acting needle bearing blocks (see Figure 8.19) were employed to ensure pre-
cise definition of simple supports on all edges. Rigid side supports, reinforced by three pairs
of horizontal rods to help prevent in-plane deformation of the side supports during loading,
assisted in the definition of the boundary conditions. The plates tested were 19 X 19 x
0.125 in. ( - 483 x 483 X 3.2 mm) in size, made of 2024-T3 aluminum Alclad, and repre-
sented typical thin plates (bit = 152) for structuraJ applications. Out of a group of 20
machined square plates, the 3 most nearly flat plates were chosen, and a complete set of
tests for all hole sizes was perfom1ed on each of these 3 specimens. The deflections at and
near the center of the plate were measured by dial gages and recorded up to 50 percent
above the estimated buckling load. After an initial set of tests on each unperforated plate,
the holes were carefully cut (with the plate securely blocked and lightly preloaded) and then,
upon completion of the test, progressively enlarged. The technique of progressive enlarge-
ment of the cutout was later also adopted in tests on perforated shells (see for example
[17.31] and (17.34]).
All the buckling loads for each plate were within ± 20 percent of the average values
recorded for the set of the three specimens. Local buckling occurred in various regions of
the plate for different loads (see [8.54] for details). The differences were quite small and
depended primarily on local plate imperfections, material properties and residual stresses (all
the three quantities were, however, not measured). In Figure 17.15 the average of the buck-
ling loads, recorded at various locations for each hole size, are depicted. All the experimental
points lay within the shaded area indicated in the figure, which provided some guidance for
designers. The theoretical curve in the figure represents an approximate Raleigh-Ritz solution
(see [17 .8]).
As pointed out in Subsection 17.1.2, the more recent investigations on the influence of
cutouts in plates focused on the postbuckling behavior and ultimate strength. For smaller
specimens, however, fairly simple test rigs were often employed (see for example Figure

Copyright Wiley
--`,`,`````,`,```

Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1428 Influence of Holes, Cutouts and Damaged Structures

.."'g
Q

~ tooo r--r-----t-----;~~~r---~
...ii'
~

u
...0
·~ --4-----~-----+-----+-----1
'!Q.U: AU. UPtiiiiO[MTAI. POINTS
L it: WITHIN l'H{

1000
CIOOSI-~ATC~£0 • 01010 t---j

I
~
~r--+-----+-----1------t----~

o~~----~----~------~--~
0.1 o.l o.s 0 .1 o.tt
r• AAnO Of' HOI.E DIAIIIETEII . TO
PLATE DIMENSION

Figure 17.15 Schlack's tests on uniaxially compressed 2024-T3 Alclad perforated square plates-
experimental buckling loads as a function of the ratio of central hole diameter to plate
dimension (from [8.54])

17.16 and [17.16]), which were similar to those discussed in Subsection 8.7.2 of Chapter 8
for unperforated plates. The emphasis here was on the application of accurate w1iform uni-
axial compression to plates of various aspect ratios. This was ensured by relatively stiff fiat
linear bearilngs and guides along with an adjustable, padded bottom platen to facilitate uni-
form loading. The load uniformity was monitored by many strain gages, while the deflections
were measured by an array of dial gages to yield the buckling loads. The test rig has been
employed in an extensive test program, on both metal and composite plates, at Paisley
College of Technology, including about 600 specimens (see [17.16] and (17.18}).
Another type of test rig, capable of applying uniaxial compression on plates, was the
horizontal desk-mounted setup employed at University College, Cardiff, for the study of the
ultimate load of perforated plates (see Figure 17.17 and [ 17 .26)). This self-contained rig is
suitable for experiments on small-scale models (up to 300 rrun wide and 300 mm long) in
a laboratory. In concept it somewhat resembles an earlier University College, London, t'est
rig, discussed in Subsection 8.2.2 of Chapter 8 (see Figure 8.1 6). It also employs beadng
blocks, which contain sets of needle beari ngs, to provide simply supported edges, but here
also for the unloaded edges. The loads on the plate are applied by hydraulic jacks actitng
against the bearing blocks, and they are measured by calibrated load cells. Axial loads up
to 90 kN can be applied in both directions.
The axial shortening of the specimens is measured by a pair of horizontally placed trans-
ducers, whi.le their initial imperfections and out-of-plane deflections are determined by ver-
tically placed linear displacement transducers (see Figure 17 .17). All the transducers were
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

connected to digital voltmeters, which had been directly calibrated to yield displacement
measurements accurate 10 0.00 I mm. The tesl' rig is shown in 1he figure with a specimen,
having a large square cutout, tested to failure.

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Bcperiments on Plates wi th Holes and Cutouts 1429

rou 5UI'~rs (d)

.,..., .....
QCJ

~·'
(c) GUO£$

H J:urt' 17. 16 Pat\lc) Cnll~gc '""'' ng lur uma"<iall) cnmprc"ctl rc.:t.mgular plate' \\llh cutout' !from
(17.1611 The ltlad "apphcd h) a hydraulic J••d.; 1.1) \ia a prccalibr.ued pru' '"8 nng tb).
\\htle a w1 of llal hnc.u ll.:;mng~ l.:l .mc.l Jll <tc.ltu'tJhle padded bullum pi.Jtcn cn,ure
accur.ue unta\Jal ~.:umpre,,i<IO ICI.Idm~ un r~'tunguiJr pl.llc\ of \J.nou' ;1\(X"Cl rauo~.
lntcrchangc:ahlc 'UJ'I't>rh (til \imulo~h! c11hcr 'unrk •uppon' of cl.unpcc.l cc.lgo:'

\ • pmntcJ out m Suh\4!~tinn 17.1.2. rigtd pJa,lll' llll'\:hani'm ,oJuuon' (dt,c.:u,,cd tn Sub-
'cc.:llon 16.2 5 of C"haptcr 161 \\Crc 'omclimc' u\4!d hi c'umatc the ulltmatc <.trength tlf ''eel
pcrloratcu plate . With \uch a mcc.:ham<.m. the plate matcn.tlt-. a''umed to bcctlmc perfect!}'
piN1c 111 one step. and thereafter all dcfonnation' tal..c pl.tcc ill the hinge'>. The hulurc pancm
uh,crvcd tn the tests for centrally pcrfomted plate' w11h '4Lmrc and circular cuwut<. (see
Figure 17.4) :-uggcstcd n relative I) simple model of collup'c mcchani~m. 'hnwn in Figure
17. 11<:1 torn ~;quare plate "ilh ~~ ..quare hole. It con~"'' ol fuur yield line!-. which qan from

l'i~tun• 17. 17 llmvel'>ll) College. Cardtff. honzomal dc'k nlnliiiiO:d IC'I ng for 'mall unw~t311) corn-
pre~'ed pl.1u:' \\ilh cutout\ !from [17 .:!61 cuurto:') nl Prufc"or R J)Dran.llll

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1430 Influence of Holes, Cutouts and Damaged Structures

r:::~ a
+ +
•+ + ~
s
T a:

0
lo

E
l
t t t f t t t t
a --j 1- · (I

(a) (b)

Figure 17.18 Rigid plastic mechanism for collapse of centrally perforated square steel plates (based
on Figure 5 of [ 17.26)): (a) for square holes, (b) for circular holes. Note that the mech-
anism consists of four diagonal yield lines and four edge yield lines, with four quadrantal
segments between them

the four corners (A, B, E. Fin the figure) and end at the cutout boundary (C, D, H.!), with
four rigid segments, like ABCD, between the yield lines (AC, BD, etc.). These yield lines
and the edges of the plate (like AB, etc.) are the axes of rotation of the rigid segments, and
the plastic .rotation of the yield lines represents the main deformation of the perforated plate.
For circular cutouts the mechanism is similar (see Figure 17.18b), with a' replaced by the
equivalent chord (0.707d for a square plate).
The rigid-plastic analysis presented in [ 17.26] yields a convenient, approximate method
for calculation of the ultimate load of simply supported perforated plates, subject to uniaxial
compression. The ultimate strength predictions correlated well with the experimental vall!les
(see Figure 17.19). The test data and the predictions in the figure indicate that the degrading
effect of square cutouts on the ultimate strength was slightly more pronounced than that of
circular holes, whereas earlier studies had found the reduction in ultimate load to be nearly
the same for both types of cutouts (see for example Subsection 17.1.2 and [17.4]).

0·5 t •77.40
EXI'IRJI1EH1Al VAl U£S
• (JRCULAR t«U
• SQIJAR E HOlE
- P~EOICTIOH F<JI CIAGLlR I-O.E
0-4 --PREDICTION FOR SQUARE ttCI..E

O·J +----------.-----...------,,------.----~0·5
0 0·1 0·2 Q.] ~ 0·4

Figure 17.19 University College, Cardiff, rest results and predictions with the aid of a rigid plastic
mechanism for uniaxially compressed, perforated plates with square and circular cutouts
(from (17.26])

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Plates wfth Holes and Cutouts 1431

A typical example of a more recent inve~1iga1ion of 1he influence of cutouh on compre<;-


~inn loaded metal plates i'i !he seric~ of 1es1s carried ()ut hy Nemeth m NASA Langley in
the late eighties. The extensive tests on isol!opic 6061 T6 aluminum plates. with centrally
located circular. ~quare and elliptical cutoul<> (see 117 .58]). fun her clarified the effects of
cutout size. cutout 'hapc and pl:ue :l'>pcct rotio on plate (Xhtbuckling behavtor. while ervmg
ao, a reference for \ trntlar e'lperiment'> on lammotcd compc->-tte. perforated plate-..
The pecimen' were machined out of 6061-T6 'heet' ol nominal Lhidmcss 1.59 mm
(0.0625 in.) into rccwngular plate' havmg a<>pect ratio~ I. l and 5 and wtdth· to-Lhicknes).
ratios of 64-160. The cutouts were muchined into the panel<; with a mi11ing machine. em-
ploying for the elliptic cutouts a ~ix-nx i~ numerically controlled milling machine. The spec-
imens were tel>tetl under axiall compre"').ion in a NASA Lunglcy plate te l li'\ture C<.ee Figure
17.20) placed m a ()53 MN ( 120 l..tp) capacity hydraulic te"'ung machine. In this fixture

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
(\\ htch i -;ome\\ hat 'imilar to that ~htm n m Figure 14 32 or Chapter 14) the loaded end~
\\ere clamped whilc the unJoaded edges were simply supponcd. The plate\ were loaded to
approximately I\\ icc the buckling load.
One side of the ~pecimens was paanted white 10 rcnect light for usc of the moire fringe
method to monitor the out-of-plane dcfonn:uions. Str:tin~ were measured by clectricaJ resi~­
tancc ~train gages. whtle axial lli!>piJccments and normal (out-of-plane) dl!.pluccments were
rnc:bured by dtrcct current differential tran.sformefl>. The clectncaJ signaJ, from this instru-
mentalion. and the corresponding apphcd loads. were recorded on magncltt tJpc! at regular
tntcf\aJ during tC'>IIng.
The NASA Langley test setup and procedure reprc!.cnt a fairly typical ,tandard approach
and have been clt~cus~ed here only bccau'>e of the wide-range testing. primarily on perforated
laminated compo:-.ite plates. that cmpluycd them.
A, mentioned in Subsection 17.12. the experiment\ on perforated plate' have al<.o been
C\tcndcd to shear loadmg. Much of the IDtllal \\Orl.. on the mOuence of cutout' m the wcbo;
on the uiLim<llC \hear capacity or pl:uc gtrdcr-. originated at t:ni\ersity College. Cardiff. in
the early eighuc~ hee (17.10]. [17.1:!1. 117.59] and [17.601). where it formed part of the
'-Ub\tanlial theurellcal and experimental mvestigation., on the posthuckling ~trcngth of shear
panels OJnd plate girde r~. carried out by II.R. Evans, Rockey. :'-Jnrayanan and nthcr<.. discu. sed
in Section R.4 of Chupter 8. A ~pecwl test rig. capable of applying a maximum load of 100
wn' onto plate gtrdcr). up to 20-m 'ran. wos built at Cardiff for the shear strength studies.
Sccuons of mis rig were used for the perforated web qud•c,, "ith lhe loadmg bcmg appJjcd

figun! 17.20 NASA L.mgley rectangular plate test fixture fur a~,ial \:ompression lo:u.hng. with a typical
'<JUM.' pctfornted ~p.:cnnen 10 i1 (coune y of NASA Longley Re.can:h Ccnt~r)

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1432 Influence of Holes, Cutouts and Damaged Structures

by means of a servo-controlled hydraulic jack, operated by a control system that permitted


load application by both load or deflection control.
The test setup and instrumentation are presented in Figures 17.21 and 17.22, or in more
detail in [17.59). Figure 17.2 1 shows the elements of the girders tested.. Two series of tests
were carried out on steel girders with webs containing central circular cutouts of various
diameters. In the first series the webs were all b = 750 mm wide, h = 500 mm deep and
t = 2 mm thick. Hence, their (hit) ratio was 250 and the diameter ratio of the holes was
(dlh) = 0.1-0.8. In the second series all the webs had b = 720 mm, h = 720 mm and t =
2 mm. Their (hit) ratio was therefore 360, and the diameter ratio of the openings ranged
between (dlh) = 0.25-0.67. Each girder consisted of two test sections with openings of
different diameters (A and 8 in Figure 17.21).
As can be seen in Figure 17 .22a, dial gages were used to measure the vertical deflection
of the flanges and electrical resistance strain gages to monitor the strains developed in the
flanges. Their positions are shown in the figure. At the end of a load increment the strains
were monitored by a data logger and processed by a computer. A low-voltage digital trans-
ducer was used to monitor the travel of the jack. The output was fed directly into an x-y
plotter so that a continuous visual display of the applied load versus vertical deflection of
the girder was available. This also helped in the selection of the most appropriate rate of
loading, especially when the ultimate load carrying capacity of the girder was approached.
The deflected shape of the web plate and of the flange plate was measured by a ripple
scanner (see Figure 17.22b).
The ripple scanner consisted of a carriage moving on an aJuminium beam (1000 mm long) which
could be flxed at any position over the girder. The transducer gauge A, mounted on the aluminium
carnage at right angles to the length of the beam, automatically recorded the displacements of the
web plate or flanges as the carriage was moved along the beam. At the same time, the transducer
B recorded the position of the carriage on the reference bar. The output from the transducers was
fed directly imo an x-y plotter, so that the deflected shape of the web/flange. was obtained directly.

The instrumentation detailed in the previous paragraphs is typical of present-day plate


girder tests (with or without web openings), though often the dial gages would now be
replaced by displacement transducers that permit automated recording of the displacements.
The test procedure here included a very simple but useful device: The aim was to test the
web with the larger opening (A) first. A 100 X 100 mm timber strut was therefore fixed
diagonally across the web that was not under test (B), in order to prevent it from buckling
under the applied loading.
Before testing, the web panel being tested was cleaned and sprayed with two coats of brittle lacquer.
The test girder was then mounted on roller supports set at t500 nun (or .1440 mm) apart and the
loading j ack positioned exactly over the central stiffener. Care was taken to ensure that the girder
was correctly positioned in the test frame; any gaps between the tension flange and the two roller

100 x 8 ot· lOO x 15 All stifleners


2 mm tr'\ICk web (nominal) { !lange (see Table 1) 49 mm x 12 mm !hick

-zmmhll el
welds

Figure 17.21 Univl!rsity College, Cardiff, shear tests on perforated steel girders-details of specirnens,
(h/1) = 250 and (dlh) = 0.1-0.8 (from [17.59))

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Plates with Holes and Cutouts 1433

All dimen~ions 1n mm _ ..ll!

St . S2 ...$24 Eleclric.al resjstance suain gauges ~~


01, 02, 00, 04 Oial gauges s~ale,ol 1'" 1 ·-
25 87 ' 88 87 88 • 1 0 '
ls't ss 59 SIJ S• 7
S2 56 510 514 5 18

e
SUDPOII 03

(a)

(b)
Figure 17.22 University College, Cardiff, shear tests on perforated steel girders-instrumentation used
(from [17.59]): (a) typical instrutnentation for plate girder tests, (b) the ripple scanner

supports were shimmed to ensure that the web plate was in a vertical plane in line with the centre
line of the j]ack applying the load.
A small trial load (not exceeding 20 percent of the expected ultimate load) was then applied
slowly, so that the shims would be compressed and the girder properly seated. This also helped to
check that tlile strain gauges and dial gauges were functioning conectly. The load was then removed,
reapplied and again removed slowly. The girder was then tested to failure. The load was applied
using small increments of load and the strains occurring in the flanges and the vertical deflection
of the girder were recorded at the end of each load increment. When the load on web A had
reached its ultimate value. and the panel had started load shedding, the girder was unloaded to
prevent unnecessary disto11ions in web B.
The diagonal wooden strut, reinforcing web 8, was then removed and the permanent
defonnations of web A and the flanges were measured.
The wooden reinforcing diagonal strut was then placed across web panel A, and web B
was tested in the same manner as web A before, except that this time the loading continued
to much larger distortions and clear occurrence of plastic hinges. After unloading and mea-

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1434 Influence of Holes, Cutouts and Damaged Structures

surement of displacements, the reinforcing strut was removed from web A and the girder
was reloaded to produce larger distortions in web panel A. After further unloading, the final
deflections of webs A and B and the flanges were measured with the ripple scanner. Note
the artifice of doubling the number of effective specimens tested (which somewhat resembles
a similar technique, using a stiff dummy section, that was employed in the Technion repeated
buckling tests on shear panels, discussed in Subsection 8.4.4 of Chapter 8).
The Cardiff shear test on perforated webs (reported in I 17.59]) showed that the ultimate
capacity of girders dropped. linearly with increase in the diameter of the circular cutout.
Similar results were also reported in other shear studies on perforated plates (for example
[17.10] and [ 17.61]).
Further shear loading tests were carried out at University College, Cardiff, on perforated
webs having rectangular cutouts. Some of these openings were reinforced by strengthening
strips on either side of the hole, which restored the strength of the girder to that of the
corresponding plate girder without cutouts (see [17.10] and [17.60]). This paper also d is-
cussed the suitability of small-scale models for shear loading tests and showed that such
models are capable of reproducing the postbuckling and ultimate load characteristics that
appear in full-scale tests.
In passing, it may be of interest to mention the complex spatial buckling pattern resembling
a Maltese cross, that has been observed around holes or cracks in thin plates subjected to
tensile loads (see I 17 .62]). This Maltese cross buckling is a local instability caused by the
second principal stress component (which is compressive) in plates under global uniaxial
tensile stress, which appears in four domains arou nd a hole or a crack. This usually beautiful
buckling pattern somewhat resembles a Maltese cross and provides a striking qualitative
visualization of the compressive stress field around a defect, such as a hole or a crack. It
therefore represents an indicator of the influence of the hole or defect on the response and
strength of the tensioned plate and can be considered to be a more general case of the
Georgia Tech experiments on thin plates with slots or cracks (discussed in Subsection 8.3.2
of Chapter 8).

17.2.2 Composite Plates


As pointed out in Subsection .17.1.2, the postbuckling behavior of perforated composite plates
strongly depends on the plate orthotropy and its interaction with the cutout. Hence, the
influence of orthotropy (detem1ined by the stacking sequence) has been widely studi,ed,
including also tailoring as a means of optimizing the strength of plates with cutouts (see for
example [ 17.13), [17.1.4], [17.17], [17.2l], [17.22], [17.63], [17.65]-[17.67]). These studlies
were analytical, using an approximate analysis developed at NASA Langley [ 17 . 14), nu-
merical, elil1ploying a finite-element analysis with the EAL code [17.64], and extensively
experimental.
The tests carried out by Nemeth at NASA Langley on symmetrically laminated angle-
ply, graphite-epoxy plates with circular holes [ 17.21 ], are a typical example of the many
such experimental investigations performed at various research centers during the eighties.
The specimens were tested under axial compression in the same NASA Langley plate test
fixture (and placed in the same hydraulic testing machine) employed for the 6061-T6 alu-
minum plates (see Figure 17 .20), discussed in Subsection 17 . I .2, and similar test procedures
were applied.
The specimens were made from commercially available (Hercules AS4/3502) tapes, which
were laid up to form 24-ply-thick laminates having [( ± 30),,], and [( ± 60)6 ], stacking se-
quences. The laminates were cured in an autoclave, after which they were ultrasonically C-
scanned to assess their quality. The specimens were then machined to their dimensions. 254
mm ( - -10 in.) long by 254 mm (- I 0 in.) wide, and their loading edges were machi ned flat
and parallel to permit uniform compressive loading. The effective lenglh and width W (the
dimensions representing the portion of the plate between edge supports of the test fixtUJre)

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Plates with Holes and Cutouts 1435

were 241.3 mm (-9.50 in.). Central holes were machined into the square panels, using
diamond-impregnated core drills, the effective hole diameter ratios (dl W) ranging from 0'-
0.66. Nine specimens, five with [(±30),J, and four with [(± 60)hL were tested.
As in the case of the metal plates, the loaded edges were clamped while the unloaded

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
edges were simply supported, and the specimens were loaded slowly to twice the buckling
load. The instrumentation andl measurement techniques used were also practically identical.
The experimental results and the approximate analysis results are compared in Figure
17.23. The analytical results are based on the effective length and width of the plate between
the edge supports (which deforms out-of-plane at buckling), as well as on the average value
of several thuckness measurements made on each specimen, and were obtained by the NASA
Langley approximate method of [17. 14). Good agreement was earlier found between the
finite-element results and those yielded by the approximate analysis. the largest difference
in the results being about 10 percent for d I W = 0.6. The experimental buckling loads for
these plates were obtained from the intersection of the linear extrapolations of the prebuck-
ling and postbuckling branches of the load-shonening c urves in each case (method no. 6 in
Table 8.2 of Chapter 8).
The experimental and analytical average buckling strains (Pc,.l EA) are shown in Figure
17.23 for displacement-loaded, clamped [( ± 30)6]. and [( ± 60)(,], graphite-epoxy square
plates with circular openings. The results of the approximate analysis in the figure agree
well with experiments for holes with (dl W) :S 0.56, in the case of the [( ± 30)cJ <plates, and
with (dI W) ::=; 0.63 in the case of [( ± 60),),. plates, and even for somewhat larger-diameter
cutouts the analysis indicates the trend in the buckling behavior.
Once the agreement between theory and experiment had been demonstrated, a detailed
parametric study beyond the actual range of the experiments was carried out by Nemeth.
This study indicated that the o rthotropy, geometry (hole size and plate aspect ratio). loading
conditions and boundary conditions play a significant role in the influence of the cutout on
the buckling behavior. Both the experimental results and the analysis, which extended its
range, showed a clear trend of increasing buckling load with increasing hole size.
One may note the procedure of first establishing a good con·elation between experiments,
an approximate theory and a finite-element analysis in the limited range of the experiments,
which include an assessment of the importance of the different parameters. On this sound
base a reliable, wider parametric study could then be carried out. In other words, a series of
experiments accompanied by extensive theoretical study led to a good understanding of the
physical elements of the buckling behavior, on the basis of which a wider-range simulation
study could be justified.

p
cr Approximate
.010 Experiment Analysis
- w- • [( ±301J 5
Average· 008 A k±60) ) -
buckling 6 5
strain, .006
Pcr
1A . 004
• 002
---·------ _..... ----------'.....•~
0~~--~--~--~--~--~·~,~
.l .2 . 3 d .4 .5 .6
. .7

w
Figure 17.23 Nemeth's NASA Langley tests on symmetrically laminated angle-ply graphite- epoxy
square plates with circular cutouts- comparison of experimental and approximate av-
erage buckling strains for displacement-loaded clamped f( ± 30),,). and [( ± 60),], plates
(from [17.2 1))

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1436 Influence of Holes, Cutouts and Damaged Structures

Nemeth continued his experimental studies at NASA Langley on square compression-


loaded graphite-epoxy plates having circular cutouts, with the emphasis on the postbuckling
behavior (see [17.22]). Unidirectional [0 11,1 and [90 ,.1], plates, [(0/90):,].. plates, as well as
[(±30)6 L and [(± 45)6 1, ang le-ply plates, were tested . Some pert"orated 6061 -T6 aluminum
plates were also included in the experimental program. The circular holes covered a diameter
ratio range d/W = 0-0.66. Most of the specimens were loaded gradually to twice their
buckling load, but some were loaded to failure. Extensive moire fringe patterns were pho-
tographed and analyzed in relation to the postbuckling behavior of the plates. T he experi-
mental results indicated that most of the plates with cutouts exhibited less postbuckliing
stiffness than the corresponding unperforated plates (see for instance Figure 17.24 showing

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
the nondimensional load versus end shortening for the [( ± 30)1J. square plates with central
circular cutouts). However, an important finding o f the entire experimental study was that
even plates wi.th large-radius cutouts preserve some postbuckling strength.
Examples of similar experimental studies on composite plates with cutouts are the VPI
investigation, which focused on the postbuckling failure of composite plates with centrally
located circular cutouts [17.65]; the VPI study of the effect of slots on buckling and post-
buckling behavior of laminated plates [ !7.66]; and the more recent VPI investigation [17.67],
which considered a new concept of fabricati ng flat plates containing cutouts by incorporating
curvilinear fiber trajectories to transmit loads around the hole. Additional examples are sev-
eral experimental studies on the postbuckling behavior of composite p]ates loaded in shear
(for instance [17.68]-[17.70]).
Recently, Nemeth [17.71] reviewed the research on buckling and postbuckling of lami-
nated composite plates with cutouts carried out during the last 25 years and summarized the
main behavioral characteristics. His paper also included a fairly comprehensive bibliography.
Among the important overall findings, it was emphasized that plates with cutouts can buckle
at higher loads than the corresponding unperforated ones and also exhibit substantial post-
buckling load-carrying capabilities, and Lhat laminated construction, coupled with cutout
geometry variation, has great potential for tailoring for optimum structural response. To
realize this potential, however, extensive experimental and theoretical studies are still needed.
As mentioned earlier, moire fringe patterns were ex tensively used in most of the tests for
assessment of the out-of-plane displacements. Other optical techniques have also been de-
veloped. For example, a variation of holographic interferometry was developed at VPI for
measurement of thickness changes with high sensitivity (see [ 17 .72]). This optical technique
utilizes a relatively simple means of holographic interferometry on both sides of the perfo-

2.0

1.5
.60
Nondimensional .66
loa~ng, 1.0

~
.5
• Buck.ling
I failure

0 1.0 2 .0 3.0 4.0


No-.dimensional end·shortening, 6/.1~r

Figure 17.24 Nemeth's NASA Langley tests on square compression-loaded graphite-epoxy plates with
circular cutouts-nondimensional load versus end-shortening experimental results for
r
[( ± 30),.), plates (from 17.22])

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
E.x,perim~nts on $hells with HQ!es and CutQuts 14:J7

rated plate specimen, fo1lowed by additive moire, in order to obtain the thickness changes
as the sum of the out-of-plane displacements.
Before leaving this section, the reader may wish to briefly review the main characteristics
of composite structures testing discussed in Sections 14.2 and 14.4 of Chapter 14.
One shou]d note as well that stress raisers in laminates, which occur under compressive
loading in the presence of holes, can initiate and accelerate failure by a local buckling process
that involves fiber microbuckling. This local buckling phenomenon and the failure that ensues
also involve delamination buckling and delamination growth, which are further discussed in
Subsection 17.4.3. Here it may suffice to mention that this failure process has been the
subject of several investigations (such as [ 17.17] and [ 17 .73]-[ 17 .77]), and direct the reader
in particular to the careful experimental study of Waas, Babcock and Knauss [ 17.74]. In
these experiments the damage initiation and propagation was followed throughout the entire
load history via real-time holographic interferometry and photomicrography of the whole
surface, and the results were verified by post-experiment ultrasonic and optical examination.

17.3 Experiments on Shells with Holes and Cutouts


17.3.1 Metal Shells and Curved Panels
As pointed out in Subsection 17. 1.3, the experiments on. cylindrical shells with large recta n-
gular cutouts carried out by Almroth and Holmes at the Lockheed Palo Alto Research Lab-
oratory in the early seventies [ 17 .29], represent an excellent example of meticulous tests,
closely coordinated with nonlinear numerical analyses. The tests warrant further discussion.
The II thin cylindrical shells tested were machined from 606 l-T6 al uminum tube stock
to the dimensions shown in Figure 17.25 (which are all given in inches). The twofold purpose
of the relatively heavy end rings was to promote unjform load distribution and to serve as
attachment rungs for the interior buckle depth-limiting mandrel. This mandrel was empl.o yed
to confine the stresses in the buckled specimen to the elastic range and th.us petmit repeated
buckling, since in view of the imperfection sensitivity of axially loaded cylinders it was
deemed necessary to test each specimen first without c utouts to establish a reference level.
In the preliminary tests on the unperforated shells, they were loaded repeatedly with the
buckle restricting mandrel in place. It was generally found that the initial test caused some
damage to the shell but that subsequent tests did not add to this damage. Consequently, for
the second and following tests the buckling load was more or Jess repeatable. This repeatable
buckling load was considered to be the appropriate reference level for assessment of the

1 - - -IZ.OO
--jl~l-- 9.00 - 1
I

Figure 17.25 Almroth and Holmes's Lockheed Palo Alto axial com-
pression tests on aluminum cylindrical shells with large
t ~ rectangular cutouts-dimensions of test specimens
~0.430 (inches), where. the wall thickness t was the only vari-
~·~-,- able within the set of eleven shells (from Ll7.2Yj). T he
DETAiL Q cutout angle was 45° for all shells bur one

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1438 Influence of Holes, Cutouts and Damaged Structures

influence of cutouts. Buckling in the presence of a buckle depth-limiting mandrel (and [he
resulting behavior) was discussed in Chapter 9 (see Subsection 9.2.1 and Figures 9.12 a nd
9.13).
Figure 17.25 also shows the cutout dimensions, as well as one of the four different types
of reinforcements used in the test program.
In the experiments, two rectangular cutouts were made after completion of the preliminary
tests. These were centered at cylinder midheight and 180° apart on the circumference.
The cutouts were made by drilling 0.062 in. diameter holes at each corner of the proposed cutout,
and then sawing along prescribed lines with a high-speed dental wheel. The wheel is driven by a
hand-held Dremel motor. The cylinder is held in a felt- lined wood cradle and the operator's hand
is braced on a bar fastened to the cradle. Some cleaning up and deburring with a swiss file is
necessary. Because of the high speed of the abrasive wheel, almost no tool pressure is required.
The size of the cutouts on ani cylinders was 45• of arc by 3 inches in the axial direction (except
cylinder No. I which had cutouts with a 30" arc).

All the reinforcements consisted of thin angle sections a nd were installed on the outside
of the shells (except in one specimen). Only axial stringer re inforcements were used, except
one shell that was tested with a full picture frame around the cutout (as shown in Figure
17.25).
Extensive calculations were carried out here both for pretest analysis, for optimum pro-
portioning of test specimens, and for post-test analysis, for better understanding of the ex-
perimental results. The computer program used was STAGS A, a program for the nonlinear
analysis of shells of general shape (see [2.53]), which was developed by Almroth and his
co-workers at Lockheed Palo Alto (see also [ 17.28) and l3.37)).
The pretest analysis guided the design o f cutout reinforcements. T he initial reinforcing
configuration of a solid rectangular stiffener, attached like a picture frame around the cutout,
was found to be very inefficient. Physical considerations and further computations indicated
that the rei nforcing stitfener at the edge of the cutout should have significant bending stjff-
ness, but its cross-sectional area should be as small as possible, to minimize stress concen-
trations. Hence, a thin angle-section stiffener (as used o n six of the seven shells with rein-
forced cutouts) appeared su.perior to one with a solid rectangular section. Furthermore, it
seemed likely that for axial compression stiffening along the curved edge of the cutout would
be of little value and that it would therefore be more efficient to extend the stiffeners in the
axial direction at the expense of the curved edge reinforcement (which could be omitted).
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

The calcula tions validated these intuitive trends a nd led to the very thin machined angle-
section reimforcements with tapered ends used in the test shells, which all had 45o cutouts
(see Figure 17.26). The computed effect of reinforcements around a 45° cutout in she lls,
with a wall thickness 1 = 0.356 mm (- 0.014 in.), is shown in Figure 17.27 for reinfor•ce-

1-I!- TJPO I ~

Q02() 0.150
Tg .o-.o--" •• ' "a 0.010 0 . 1150
_L_~_[ c QO IO 0.080
p•
1--aZS~-! T 0.020 0 .1 50
'SH ~It 17. 2S

Figure 17.26 Lockheed Palo Alto axial compression tests on aluminum cylindrical shells with large
rectangular cutouts- details of stringer reinforcement used (from (17.291)

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Shelfs with Holes and Cutouts 1439

~,r-----~----~------~----------~

0.1

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
0 0 . 00!1 0.010 O.OZO O.OZll
JWNFOfiC[IoiENT TMICICNESS, lr

Figure 17.27 Lockheed Palo Alto axial compression tests on aluminum cylindrical shells with large
rccmngular cutouts-eomputcd effect of reinforcements around n 45" cutout in shells
wilh a wall Lhicknc's of 0.014 in. (from [17.291)

ments of various thicknesses t,. Only three data points were available. but they indicated a
clear trend that a reinforcement using only axial stringers (type A or B in Figure 17.26) is
more efficient than the equivalent picture frame (type P in Figure 17.26).
A geneml view of a typical shell with an unreinforcetl cutout (no. 3) after buckling i s
shown in Fig ure 17.28. The specimen buckled at 2050 lb (- 9. 12 k.l'J) but went on to carry
2120 lb (-9.43 kN). The ignificant size of the cutout is immediately apparent. Note also

Figure 17.28 Lockbeed Palo AJto axiaJ compression tests on aluminum cylindrical helh with large
rectangular cutout~-general view af1er buckling of a 1ypical 'hell, wi1h an unre inforced
c ulout (specimen no. 3) (from [17.29))

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1440 Influence of Holes, Cutouts and Damaged Structures

the massive upper and lower, 2-in. ( - 50.8 mm) thick, square aluminum end plates, whose
contacting surfaces were machined to a flatness better than ±0.0005 in. (- ±0.013 mm) to
ensure load uniformity.
In the correlation between experiment and theory and in the evaluation of the experimental
results, Almroth and Holmes introduced a quality parameter cJ> (commonly known as the
knock-down factor). Since they compared the buckling loads of the shells with cutouts to
the repeatable buckling load (obtained with a buckle-restricting mandrel) of the correspond-
ing unperforated shells, this quality parameter of the specimen was included in their reference
load. In the discussion on the benefit of the reinforcement of cutouts, they pointed out that
if the quality parameter </> 1s relatively low, the unreinforced cutout may represent a less
severe imperfection than those present in the original unperforated shell. Jn that case, rein-
forcement ·Of the cutouts will not increase the design load of the shell, whose buckling load
will be determined in any case by the initial imperfections of the unperforated shell.
In general, the agreement between experiment and theory was very good and provided
support to the trends obtained by the computer-based nonlinear analysis (STAGS A). The
important conclusion was that careful experiments, closely coordinated with extensive cal-
culations, can provide design guidelines; though Alrru·oth and Holmes pointed out that more
experiments and analysis were needed to yield reliable design procedures. They thus provided
one of the foundations of the recently proposed design methods for shells, involving com-
prehensive imperfection measurements, as well as imperfection and boundary condition data
banks (discussed in Chapters 10, 11 and 15).
As mentioned in Suhsection 17. 1.4, several practical design prohlems related to the ef-
fectiveness of cutout reinforcement were studied in the seventies (for instance [17.45],
[17.46] and [ 17.48]). Another example from that period is the experiments on torsion-loaded
thin-wall cylindrical shells with cutouts performed by Strenkowski and Witmer at MIT (see
[17.78]). The shells were fabricated from 76.2-mm (-3-in.) outside diameter 6061-T6 alu-
minum alloy rubing to produce specimens with (R I h) = 90.7-97.8 at their central part,
which was milled down at a thinner test section to coax buckling to occur there. The tests
are of interest because they compared the load-deflection behavior under pure torsion of four
kinds of predarnaged shells: (I) shells with a 25.4-mm diameter circular cutout, (2) with a
round-ended rectangular slot, 9.5 mm wide by 76.2 mm long, oriented either axially, cir-
cumferentially or midway between axially and circumferentially (see Figure 17.29), as well
as of corresponding initially undamaged shells.
The buckling and postbuckling behavior up to large angular deflections (well into the
plastic range) was studied, with attention to the occurrence of secondary buckling, tearing
and pronounced changes in the defonnation pattern (see for instance two progressive buck-
ling patterns for a shell predamaged with a circular hole in Figure 17.30). All the types of
cutouts significantly reduced the incipient buckling of the shells as compared to the unda-
maged ones (down to 39 percent for the circular holes and down to 45-51 percent for the
rectangulali slots in the various orientations) as well as noticeably affected the large-angle

~ 0 J p =
o• slot
6
(a) 1-tneh Diameter Hole (b)

~ ~ 6 ~ ~ 6
(c) 45° Slo~ (d) 90° Slot

Figure 17.29 MIT tests on torsion-loaded cylindrical shells with cutouts-schematic of cutouts in
prcdamaged shells (from [17.78])

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Shells w;th Holes and Cutouts 1441

(a)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(b)
Figure 17.30 MIT tc~b on IOr,inn-lnaucd cylindru.al 'hell~ 1\Hh cuwuh typical budding paucnt\ for
predamag:ed 'hell wub ciJ'Cular hulc Cl mm 117 7RJI Ia I amu:ll hud.hog pauem at 2" t\\ "'·
tbl progre,"on ,,f p.~uc:m at s '"'''

behavior. The MIT re earcher:-. however. pmntcd out the need for much lunher re earch in
order to provide databases for <tppropriale muucling and dc~i gn .
Indeed...evcrnl further C\pcnmental studic\ of the infllilcncc of cutouts and thetr reinforce-
men" on the buciJmg behavaor of metal 'hell' were earned out to the etghuc' and nineue-.
hce for tn'>tancc 117.79)- [17.811 and [1 7 ..HJ) The fi~t of thc..c 117.79) rcpon ' on buckling
tests of ~ix ' tccl '>hell!. v.. ith CR/1) - 500. hnvtng nuclear cont.unment-lil..c feature~. Four of
the shell<; hat! reinforced circular cult)uts (rcinl'urccd in accordance with ASMF. code rules).
the larger ones cutting one lCI three of the ring ' tiffener-.. Geometrical imperfections were
mea urcd and recorded but were not fully currdated wuh the calculation'\. The !>hell" were
loaded \\tth ..m olf<;ct axial luau and bucL.ku pla,ticall). The tc,ts and the analy-.es . howed
that the larger cutouL.,, even though reinfon:cd, rcdut"ed the buckling capacll) of the shclb
In the mtd-cightics a comrm·hcn<.ive ~eric~ of hrgc-scalc 11.:,1!> on a stiO"cncd cylindrical
'hell with rcctnngular cutout' wa~ carried out at NASA Langley Re~carch Center (~cc
[ 17.80]). ThC' 'hell w<fi a lO:l· m C10-fl) di<lmcter aluminum alloy integrally 'tiffcncd hell
~ASA LA- I. aln:ad) di~CU'i'-CU 111 Section I0.5 of 01apter I() 111 connection with 1t~ complete
tmpcrfecuon -.ur. C) (-;ee h gure 10. 15 and 110.1211. and had completed another tco;t program
<~I 1\'ASA Mar, hall. After a re(.;t.mgular opcntng w~ cut into ''' ring- and c;tringer-st.iffenec.l
wall. the shell wa~ loaded to bud.ling failure by an end-bending moment. The moment wa'
.tpplied in the large ASA Langley bending tc't rig. To facilitate the tran,fcr of the bending
load., to the q hnder "all. lal}!e ~teeI loauing rings were machmed and llohcd to the end'

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1442 Influence of Holes, Cutouts and Damaged Structures

of the cylinder, while individual steel backup plates were bolted to the inside of the shell
wall for the same purpose. The steel loading rings were in turn bolted to a pair of conical
shaped loading heads. One of these heads was bolted to the rigid vertical wall mounted to
the laboratory floor (the backstop), while the other conical head was attached to a steel end
plate, which in turn was attached to the usual triangular loading frame (see Figure 9.75b of
Chapter 9 for a similar loading setup, but without the conical loading heads needed here for
the larger diameter specimen).
As pointed out in Section 10.5 of Chapter 10, the shell wall integral stiffening was mill
machined into flat aluminum plates, which were then roll formed to the shell radius. Three
such rolled plates were finally welded together to form the test shell, and the three equally
spaced weld joints could later be recognized in the initial imperfection pattern.
The position of the first rectangular cutout in one of the segments was centered relative
to the weld joints and to the ends of the cylindrical shell. The first cutout was 24 X 27 in.
(-610 X 686 mm) and was cut in a manner such that two of the shell wall cutout sides
were stringer supported while two were ring supported. In other words, the opening was
delineated by eight stringer spacings (8 X 3 = 24 in.) and by six ring spacings (6 x 4.5 =
27 in.). The second cutout was a 36 x 36-in. (-914 x 914-mm) opening made by enlarg ing
the first cutout area and covered an area of 12 stringer spacings by 8 ring spacings. The
strain gage readings and visual examination of the shell wall from th e first 24 X 27 in.
cutout test revealed that the plastic permanent damage to the shell wall was confined to a
local area around the cutout. Hence, it was reasoned that a 36 x 36-in. cutout would remove
all the yielded wall material left from the first cutout test. The third cutout was an 18 X 18-
in. opening made in the center of another of the three shell segments. For the third test, the
36 X 36-in. cutout opening was covered by an equivalent extensional stiffness aluminum
patch plate whose dimensions were two stiffener spacings wider than the cutout to allow
room for bolting the patch to the shell. Apparently, the patch plate fully compensated the
shell for the loss of stiffness caused by the cutout.
The test shell was densely instrumented: back-to-back strain gages were located around
the edges of the cutouts, and many LVDT's for deflection measurements were placed near
the top shell generator and around the cutout edges. AJl the data were recorded by the NASA
Langley automated data acquisition system.
The stiffened cylinder was tested in bending three times, in each case with one of the
three rectangular openings in the compression side of the shell wall. Failure occurred by
buckling of the shell wall around the cutout in all three cases.
The experimental results on the large-scale integrally stiffened shd l loaded in bending
confirmed the trends observed on smaller models and other loads. The square openings
significantly reduced the buckling capacity of the shell, and the magnitude of this reduction
depended as before on the nondimensional cutout parameter r = (r/YRt), given by lEq.
(17.1) of Subsection 17.1.3, where r is the cutout area characteristic dimension (here half
the side of the square or rectangle). Though the buckling moment of the undamaged shell
is not reported in [17.80] (and hence exact comparisons are not possible), it appears that the
buckling moments for the shell with cutouts, when plotted against the cutout parameter r,
would roughly fit within the curves of Starnes's results for small isotropic shells under axial
compression (Figure 17.8 of Subsection .17.1.3).
Another example of comprehensive tests on the influence of cutouts on the buckling of
cylindrical shells is the experimental program recently carried out by Jullien at INSA Lyon,
France [17.81]. The aim of this study was to provide data for the development of a reliable
physical model and hence the deduction of a general design rule. The lNSA tests were
performed with great care and included many (over 100) specimens.
The shells were all nominally identical, with R = 49.5 mm, t = OJ 75 mm and length
L = 104 mm, i.e. (Rit) = 283 and (L/R) = 2.1, which could buckle in the elastic or elasto-
plastic range. They were made from steel sheet by rolling and electrical welding along a
generator. Both operations were automated to ensure that the resulting geometric imperfec-

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Shells with Holes and Cutouts 1443

tions were practically t'he same for all specimen~. The openings were always located on the
opposite side of the weld and included six type!. of cutouL~ (shown in Figure 17.3 1).
For each specimen the geometric imperfections were measured at 25,000 points by a
noncontact sensor auaehed to a special automated scanning system. The initial imperfections
were mainly located in the vici nity of the welded generator. The reproducibility of the te~ts
was verified by a comparison of the buckling loads measured for eight unperforated shells.
which exhibited a standard deviation of only about 5 percent of the mean.
The governing nondimcnsional cutout parameter r "' (r/VRt). proposed by Stame~ (see
Eq. ( 17.1 ) of Subsection 17. I .3). was found to apply al~o to Jullien's results. As a matter of
fact, the 1:-.ISA results on steel shells, with (Rir) = 283. compared rather we ll with the earlier
one obtained by Starnes on Mylar ~hell~ with (R/r) = 400 (sec Figure 17.6 and [17.3 11
and [17.32]). In Figttrc 17.32 the ratios of the experime ntal buckling load!i to that of the
corrc ponding perfect unperforated shell are presented versus the nondimcnsional cutom
r
parameter for some of the JNSA shells a~ well as for the Caltech ones with circular holes.
II appears that in both test series there wa~ a coupling between the effect of geometrical
imperfections and that of the cutout for small openings. whereas for large ones the cutout
dominated the resultant buckling load anu behavior.
The last example of metal shells is the recent ASA Langley tests on compression-loaded
6061 -T6 aluminum singly curveu panels with a central circular cutout (see [ 17.43]. which
were tmeny mentioned in Subsection 17.1.3.
Curved cylindrical panel tests under axial compression present a serious boundary con-
dition problem along the unloaded edges, as well as difficulty in ensuring uniform axia~
di phwcment. (end shortening) over the whole panel. The~e two prohlems were discussed
in detail in Section 13.6 of Chapter 13, the vertical boundary conditions in particular in
relation to the Lockheed-Berkeley largc-scaJe panel test of the mid-sixtie and the end-
shortening unifom1ity with respect to the Lockheed Palo AJto test~ in the mid-eighties (see
Figures 13.69-13.72 and [ 13.132]- 113.1341 and [ 13.138]).
The problems of curved panel tests, which may be even more severe in composite struc-
tures, were discussed in Sub ection 14.4.1 and 14.4.2 of Chapter 14 for laminated composite
panels. They arc also referred to in the next subsection. which discus es the composite shells
and curved panels with cutouts.
The • ASA Langley aluminum panels had a nominal radius of curvature R = 1524 mm
<- 60 in.) and a nomina] thickness r 2.54 mm (- 0.10 in.). Hence. their (Rir) = 600 and
they were rather thin-walled. Their length and arc width were 374.7 mm (- 14.75 in.) andl
368.3 mm (- l-l.5 in.). respect ively.
The panels were loaded slowly in axial compression by unifonnJy displacing the opposite
curved edges with a hydraulic test machine. "The loaded ends of a panel were clamped and
the unloaded edges were simply supported by a test fixture. The unsupported length L and
arc· width W of the panels in the test fixture were both 355.6 mm ( 140 in.). Electrical rcsis-

ease S casc 6 ca~e 7

Figure 17.31 INSA Lyon test~ on the eiTcct of cutouts on che buckling of ~ted cylindrical shells
subjected 10 axial compression-shapes and locOition~ of openings (from r17.81))

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1444 Influence of Holes, Cutouts and Dam aged Structures

.: • .. .
• •
0._--~--~----L----

8 • 10
r. r/'/Rt
Figure 17.32 fNSA L)nn ''''' on the effect of cutuul\ nn the ll\Jclo.hng of 'teeI c) hndnc:;ll ,hcJh
'uh~·dedto 3\131 compt'h 100~omp.mw11 "' I 'iA rc~ull\ ~ llh earhcr uoc,, nt>t;.uned
h} StilfOC' IIL'~21un M)lar ,~u, 11"''" 111 XII•

l,m~:c ,lrJin gauge \\Crc u\Cd 10 mca.-.ure \lr.lln,. .tnt! thrccl c:urren1 diltcrcmial lr.tn,lonner;
'~ere u'cd 10 mca .. urc J.\tal tll\placemcnt' and tll\pl.t..:cmcnt' nonnal to 1hc panel 'urfacc.
Sh,ttluw mom~ imerfemmcll) Wlh abo u~ctl 10 monllor dl\placcmcnl\ nonnal to the panel
'urfacc " This rcprc.,cnh l)'plcal prc,cm-da) pracucc
'I he c\pcrimcnlal rc\ull\ arc prc,cntcd in figure I 7 l l ·" curves of nondtmcn,IMJI load

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
PIP: l \-Cr'u~ nonduncn,tonal end 'honemng t.l/t). '~here P, '' 1hc hncar htfurcauon load
lor ..tn unp!:rforatctl panel. '" h•ch wa' oblaim:d h) 1hc 51 \GS code. The tla,hctl hnc tn the
tirurc corrc,pond, to a panel \\ ithout a cutout and the 'oht.l hnc' corre,pontl to pane), "ith
cu1out tltamctcr-to-panel-\\lthh rauo (t//\V) =0.3. 0 4 untl 0.5
The C\penmemaJ CUI"\ c' an Ftgure 17.33
mdi~Jie thm the ch:aractcr of 1~ nonhnc3r rc'pon-.c nl a pJnd change' "gmhlantly a' the ...-utoul
'lie mcrca,c\. 1-'or c'nmple. the re>ulu. mdteaiC that the p.mel' wnh diiV 0 and (ll exh1h11
hucklmg bcha\Jor thnt tn\ohc\ n U)llUilliC change flllm «lllC \tahlc equ1hbmun conligurJIWn to
annthcr S1milar re~uh\. nul ~hown in the figure. wen: nhtmm:d fur panel~ ~ ith til IV 0.1 and
() 2. The I'C\Uh m the figure nhn indicate that the pund' '"uh tillY 0.4 and () 'i do nm e».h1b11
thl' type (lf bcha~ior. but c~ubu 'table. munotum,all) mcro:,l\111£ nonlinear re'J'!Oil"' ·n1c re\ull'
,,,.~ that tbc mten\11) ul the U)DJIUK bud.Jmg pn-..:e" 4kcrca'" \Ub,tanuall~ ll.\ J / 1\ mc;n:J\C'
I rom u \aluc of 1cm til 0 ~ 1llc mtclbll) of tllc d)nJilliC lllk:l..hng rc'pon-.c " md1~.1tcd b) the
l.hfto:rencc m the budJmg lc•aJ and the IO\\~t tal:llc fl'hthu.l..hng 11'341.

I0

01
,
..:, u

,.
fig~ 17.33 -.:1\S \ l...mgk) te't' on a\lall) wmprc"cd 6061-
.16 .llummum cul'\cd paneh ~ 1th a ~:cntrJI cir-
cular t uhlut C'l:penmemal n<lmllmen\IUn.ll load
\ 'CI\ll' end \hOIICillll[l CUI'\C\ lor \ UrtiiU\ <:UIOUI·
d1arnct•r tn panel ~1dth rallo'. till\ (), () J. O..t
and !l.'i llrum I17.4J]l

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Shells with Holes and Cutouts 1445

The shadow moire patterns on the convex or outer surface of the panels (see (17.34]),
provided additional insight into the effect of cutout size on the character of the nonlinear
response of the panels. These patterns indicated the significant prebuck.ling deformations
around the cutout, which dominated the response of the panels with (dl W) ;=: 0.4, while
there was no dynamic buckling response.
This example of cw·ved panels with medium- and large-sized cutouts illustrates the un-
expected nonlinear response caused by openings. This response would usually not be con-
sidered by designers, who would thus overlook some potential gains inherent in such be-
havior. ew NASA design recommendations should therefore identify some selective
experiments to be conducted in conjunction with advanced nonlinear analysis in order to
provide better guidance to the designer.
In their 1997 paper [ 17 .43], Nemeth and Starnes also made some more general suggestions
for future experiments, which warrant a brief quotation here, though they obviously apply
also beyond shells with cutouts. As pointed out in Section 1.3 of Chapter I, meaningful
knock-down factors for shell buckling depend greatly on high-fidelity experimental results,
and "the knockdown factor may be overly conservative when questionable test results" are
used that correspond to relatively low buckling loads. Nemeth and Starnes therefore presented
the following timely recommendations:
To obtain high-fidelity experimental results, there are several issues that must be addressed and
several tasks that must be performed. Prior to conducting an experiment, initial geometric imper-
fections of the shell surface, the thickness distribution, unevenness of the loaded edges, and the
material properties should be measured. Knowledge of these quantities is extremely important for
obtaining good correlation between results from theory and experiment. The instrumentation for a
test should be planned adequately to facilitate the correlation between results from theory and
experiment, and should be planned to provide enough data to help understand the expected behavior.
The data sampling rate should be adequate to represent the shell response. The instrumentation
should include back-to-back strain gages for monitoring bending strains and local nonlinear de-
formations, direct-current differential transfonners (DCDT's), or other similar devices, for moni-
toring displacements normal to the shell surface, and shadow moire interferometry for qualitatively
monitoring buckle patterns. The amount of instrumentation can be determined from preliminary
analyses in many cases.
For experiments that involve load introduction by displacing a platen of a loading machine,
proper alignment of the platens should be verified, and DCDTs, or other similar devices, should
be used to define the plane of the loading platen and to detennine if there is a load introduction
anomaly. The loaded edges of compression-loaded shells should be measured to make sure that
the edges are as close to flat and parallel as possible. A loading rate should be selected that is
consistent with the goals of the test.
Details of the test fixture, and its relationship to the desired boundary conditions, should be
clearly defined when reporting test data, and the locations of all instrumentation that correspond
to the results that are presented should be clearly indicated.

These recommendations are quoted here (maybe somewhat out of place) because they
clearly stress and reconfirm similar guidelines, emphasized by us in several other chapters
of this book-guidelines that warrant repetition!

17.3.2 Composite Shells and Curved Panels


Just as in plates (see Subsection 17.2.2), or even more so, the buckling and postbuckling
behavior of composite shells and curved panels strongly depends on the material properties
(partly determined by the stacking sequence), the cutout size and shape and the nonlinear
interaction between the resultant out-of-plane deformations and in-plane stresses near the
opening. These effects have therefore been widely studied (see for example [17.42] and
[17.82]-[17.91]), including many experimental investigations.
The recent experimental research on buckling and postbuckling of composite shells and
curved panels with cutouts has mainly been carried out at NASA Langley, Wright-Patterson

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1446 Influence of Holes, Cutouts and Damaged Structures

AFB, in particular the Air Force Institute of Technology (AFIT) there, the University of
Michigan Composite Structures La boratory and the University of lllinois Center for Com-
posite Materials Research.
Cylindrical composite shells with cutouts were tested a t NASA Langley (see Figure 17 .IO
and [17 .89']-[ 17.91]). T he specimens were graphite-epoxy cylindrical shells with rectangular
cutouts. Six of these e ight-ply laminated shells
were fabricated on an aluminum mandrel using 12.0 in (-305 mm) wide unidirectional Hercules,
Inc. AS4/3502 pre-impregnated graphite-epoxy tape. The shells had a nominal radius of 8.0 inches
(-203 mm) and a length of 16.0 inches (-406 nun). The specimens have (+45/0/90], and [+451
o•
02 ]s stacking sequences with plies aligned in the axial djrection of the shell. The nomjnallamina
ply thickness was 0.005 inches ( - 0.127 mm), resulring in a total wall thickness of 0.04 inches
(-1.02 mm).
After the shells were cured according to manufacturer's recommended procedures, they were
ultrasonically inspected to insure initial quality of the shell. A C-scan technique was employed for
this task using a pulse-echo method. No sigruficant defects were identified in the shells. A rectan-
gular cutout was then machi ned in each of the cylinders at e = o•, at the shell mid-length.

The cutouts tested included 1.0 X 1.0-in. (-25.4 X 25.4 -mm) squares, 1.0 x 2.0-in.
circumferentially aligned rectangles, and 2.0 x 1.0-in. axially aligned rectangles. Each of
the cutouts had re-entrant corners with a 0.05 in. ( - 1.27-mm) radius.
To ensure that the ends of the specimen remained circular during testing and did not fail prema-
turely, the ends were potted in an aluminum-epoxy casting material, HYSOL TE-5467 aluminum
filled epoxy with 1-lD-0111 hardener. The potting el\tended approl(ifll<l!ely 1.0 inch in from botltl
ends of the specimen resulting in a total unsupported shell length of 14.0 inches (-356 mm). The
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

ends of each specimen were then machined flat and parallel.

The test shells were simillar to the specimen of an e arlier NASA Langley program shown
in Figure 17 .I 0, but slightly larger.
Three sens of initial imperfections were measured in order to characterize the geometry of each
specimen. The measurements included the results of a three-dimensional survey of the inner and
outer shell surfaces to characterize the initial mid-surface geometric imperfections, thickness var-
iations, and material property variations in the cylinder. The surface measurements were taken
every 0. 125 inches (-3.18 mm) along the unsupported length of the specimen and every degree
around the circumference. Measurements of the loading surface geometry were obtained to char-
acterize nonuniform load introduction into the specimen, and measurements of the potting thickness
were taken in order to characterize any nonuniformities in the boundary conditions.
Surface strain measurements were made with back-ro-back electrical resista.nce strain gage pairs
mounted on the inner and outer surfaces of each specimen. Strain measurements were taken at
over 80 locations on each specimen with the majority of the gages positioned near the cutout.
Displacements of the test frame loading platens were measured with direct current differentia l
transformers (DCDT"s) at three locations on the top loading platen.

All elect rical signals from strain gages, DCDT's e tc. were recorded during the tests, o n a
magnetic tape data acquisition system, at one-second intervals.
The specimens were teste d in a 300 kip ( - 1330 kN) capacity hydraulic loading frame.
"Before each test, a load balancing procedure was u sed to insure that the loading platens
would apply the load to the specimen as uniformly as possible . The load was applied to the
specimen via a controlled ax ial end-shorte ning displacement. All specimens were loaded to
the point of global collapse."
P1ior to testing, the response of the six perforated laminated composite shells (with two
stacking sequences and three cutout sizes and shapes) was predicted using the STAGS C
finite element analysis code [14.89]. This "nonlinear analysis included the effects of initial
measured geometric imperfections, nonuniform load i ntroductio n, boundary condition non-
uniformities, shell wall thickness variations, and material property variations."
The correlation between numerical and experimen tal results was studied in de tail a nd
found to be fairly good . N umer ical results could therefore be used to complement the ex-

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Shells with Holes and Cutouts 1447

perimental ones for better understanding of the buckling and postbuckling behavior of the
shells.
[The] results indicated that a nonlinear interaction between out-of-plane defonnations and in-plane
stresses near a cutout in a compression-loaded shell caused a local buckling response to occur near
the cutout. Additional load could be applied to the shell in the postbuckling range before collapsing
into the general instability mode. Each buckling event was accompanied by a reduction in the
effective ax.i<tl stiffness due to the formation of large deformations in the shell. The compression
response was stro ngly inlluenced by the local displacements and internal load distribution near the
cutout. These local displacemems and internal load disu·ibutions were affected by the site of the
cuto ut, material properties, and imperfections in the shell.
[The1results indicated that an increase in the size of the cutout caused a reduction in the buckling
load of the s hell and the magnitude of the load reduction associated with the buc kling response.
Numerical results predicted that the buckling load of a compression-loaded axially stiff orthotropic
shell was less sensitive to the addition of a cutout than the buckling load of a quasi-isotropic shell.
The experimental results seemed to indicate that the buckling load of an onhotropic shell with a
cutout mi.ght be more sensitive to other imperfections in the shell than the buckling load of a quasi-
isotropic she ll with an identical cutout.

Most of the other experimental and analytical studies of curved composite clements with
cutouts, considered curved panels (see for example [ 17. 82]-[ 17.88]), though several addi-
tional test series on unperforated composite cylindrical shells were perfonned (see the pre-
sentation in Subsection 14.4.lb of Chapter 14, or for instance [14.186], [14.189) and
[1 4.1 9 1]).
As pointed out in the previous subsection, curved panel tests under axial compression
present serious boundary condition problems. These may interact with nonlinear behavior
near the cutout, or with the panel orthotropy. These boundary condition problems are there-
fore even more difficult to accommodate in composite panels tests. They were discussed in
detail in Subsection 14.4. 1b of Chapter 14 for unperforated curved panels (see for example
Figures 14.137- 14.147 and the related references quoted in that subsection) and are the
subject of additional studies for curved panels with cutouts (see for instance [ 17.85]- [ 17 .87]),
that need not be reiterated here.
.Palazotto's 1988 experimental study [17.87) presents a typical example of a test fixture
for curved panels, based on the accumulated experience (whose details were described in
Subsection 14.4.1b of Chapter 14). The specimens were graphite-epoxy cylindrical panels
of (Rit) = 305, eight plies of orientations (0/ ±45/90)5• and with central square cutou ts of
(d/W) = 0.167. As shown in Figure 14.147 (and discussed in Subsection 14.4.lb) and in
[17.87], the vertical edges of the panels were supported by side knife-edge supports (nearly
simple supports), while their horizontal edges were clamped by a wedge technique. The
instmmentation used (strain gages and LVDT's) was also similar to that employed in the
tests for nonperforated panels, discussed in Chapter 14, except that six LVDT's were added
around the edge of the cutout.
The nonlinear etTect of the c utout on a typical panel as the load increases, can be seen in
Figure 17 .34, whi.ch shows the deformed panel at 80 percent of the collapse load and then
at collapse load.
The test fixture, as modified by Palazotto for curved panels with cutouts, represents a test
rig of reasonable accuracy. The concurrent finite-element analysis (with STAGS C) wa<; very
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

useful there irn the determination of certain features of the test rig.
The NASA Langley mid-eighties experiments on axially compressed graphite-epoxy cy-
lindrical panels with circular cutouts [ 17.82], represent a good example of a test series on
composite curved panels employing present-day techniques.
The nine specime ns tested in this investigation were fabricated from commercially available uni-
directional Tlhornel 300 graphite fibe r tapes preimpregnated with 450 K (350 F) cure Nannco 5208
thermosetting epoxy resin. ... T he tapes of 0. 14 mm (- 0.0056 in.) nominal ply thic kness were
laid up to form 8- and 16-ply q uasi-isotropic laminates with [ ± 45/90/0ls and [ ±45/90/02 /90/
+45]., stacking sequences. The laminates were cured in an autoclave using the resin manufacturer' s

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1448 Influence of Holes, Cutouts and Damaged Structures

(a)

(b)
Fl~ture 17...34 PalaL.ouo·, ,\fIT •~"'' on :n:ia.ll} C(>llljlfe'>-cd cul'\ed compo,tlc panel' ~llh -.quare cut-
ou~ a tyruc.JI dc:formed panel as the loJd 1ncn:a-e' (from [ 17 !i7)): Ia I at 110 percem
of collap-.: loild. (h) :11 collapse load

rccummcndcd procedure\, rollowing cure. the laminate' were uhra~onically m~pcctcd 1<1 eMabh'h
'i>''cimcn quality und 1hen \h!I'C cut into lest specimen' I he end' of tho.: '>pccimcll' were moctuned
nat and parallel to permit umfonn compn:\Si\c load1ng Circular hole' wen: rnuchincd IntO lhe
\pcCimens \1 itb diamond unrrcgn:ucd con: drill'>.
The com·e" side ol ca~h spccnnen was pamted \\hllc to reflect light 10 order to facaliaate
1hc u-.c of a moire fnn!!c tcchmque 10 morutor rada.tl d"placemenl'> dunng IC\llOg The
'>JlCCitnCn!> te!.led had a :\56-mm 1- l+in.) -;quare pruJcclcd planfonn. \\llh c11hcr a 1524-
mm c 60-in.) or a 381 -mm 1 ~ 15-in.) radau~ ol curva1urc. They had calher nn hole o r a
'mglc 13-mm ( - 0.5-in.), 25-mm (- 1 in.) or 51-nun ( 2-m.) diameter central circular hole.
The appara1us and test procedure were ~imilur 10 1ho~c employed in the NASA Langley
test~ on aluminum curved panels. di~cu sed in the prcviou' ~ubsection. where the extcn ive
concurrent calculationo; were carried out w11h the STAGS C cornpu1cr code.
Each specimen te~ted bud,led into a mode with large radial displacement and c~hibitcd
'>Orne p<Nbuckling strength. A' i' 1} p1caJ of C} hndncal -.hello; or curved panct... the a:~tiaJ
load 'upponed by each 'JlCC.:IIncn decreased r.tp1dl) at bud..ting. The '(>C~lmcn' walh holes
had nonlinear prebuclllng luau-<,hortemng re~pon'c (.Uf\c,. ucfonned locally ncar the cutout
and hucllcd at lower applied load\ 1han the (Om!<.pondang unpcrforated cuf\ed pane!... M oire
fnngc palterns reprc~cnling the radial displacemenl.; were lhCd 10 identify the buckling mode
~hapc of each specimen .mu 10 monitor vi<;ually 1hc change in Lhe posthucl..ling radial dis-
plnccmcnt pauems of the specimen~ (sec Figure 17.36).
" Load-shortening rc~uhs for the 16 ply specimen' :lrc shown in [Figure 17.351 a-. a func-
unn ol' the applied load. The mcao;urcd end shoncnin~ 1'l of each specimen '' nomlahtcd by

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Experiments on Shells with Holes and Cutouts 1449

. 003 If£f~ d.O

ICPll
.002
p
fA
. 001

0 .002 .CK'A .006 . 008


(a) 6/L

• FAILURE

. 003

Figure 17.35 NASA Langley tests on axially compressed g raphite-


. 002
p epoxy cylind rical panels with circular c utouts-
IA experimentally determined load-shonening results for
16-ply quasi-isotropic specimens (from [ 17.82]): (a)
specimens with a 1524 mm panel radius and a 2.3
mm wall thickness, or (R i t ) = 663, (b) specime ns
.002 .004 . 006 . 008 with a 38 1 mm pane l radius and a 2.3 mm wall thick-
(b) 6/t ness, or (R!t) = 166

the specimen length L and the applied load P is normalized by the specimen initial prebuck-
ling extensional stiffness EA. The resulting nondimensional parameters f>f L and PI EA can
be interpreted as measures of an applied displacement and an applied load, respectively."
The experimental results shown in Figure 17.35a indicate that the load-shortening response
curves for the specimens with a 1524-mm panel radius are nearly identical for various hole
diameters. Note that some of the specimens with the larger panel radii were loaded by
uniform end shortening to nearly eight times their end shortening at buckling before failure
occurred. "The presence of a circular hole in specimens with the larger panel radius had!
little effect on either the buckling load or the postbuckling response of the panels. All
specimens buckled into a local mode shape around the hole at relatively low levels of the

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
normalized applied load." This local mode shape appeared to be embedded in a global mode
shape at buckling. The experimental results shown in Figure 17.35b indicated that the pres-
ence of a hoEe in specimens with a 381-mm panel radius had a significant effect on the
response and failure of the panel. The specimens buckled at relatively high levels of the
normalized applied load, and their postbuckling behavior depended on the hole size.
Photographs of typical moire fringe patterns of the radial displacements for specimen CP4,
with (R/1) = 1385, are shown in Figure 17.36, before buckling (a), after buckling (b), and
after secondary buckling (c). The first reduction in the normalized applied load corresponded
to a local deformation pattern around the hole embedded in a global buckling mode shape,
as depicted by the moire fringe shown in Figure 17 .36b. " This pattern is radically different
from the pattern before buckling shown in Figure l7.36a, and is typical of the initial post-
buckling response of specimens with holes. The local deformation pattern. shown in Figure
17.36b has two local regions of inward radial displacements near the hole that are separated
by nodal lines that intersect the hole. Some of the specimens failed in this initial postbuckling
response mode. Other specimens subsequently buckled into a secondary buckling mode sim-
ilar to the one shown in Figure 17.36c as the applied load is further increased." When
secondary buckling occurs, a second reduction in the load occurs, and the radial displacement
pattern changes to an overall curved panel mode shape, dominated by the hole with inward
radial displacements, as shown in Figure 17.36c.
The experimental results indicated that even low values of the applied load caused the
development of high local strains near the hole. "The level of the local strains typically
approached the strength limits o f the graphite-epoxy material system. It appears that the high

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1450 Influence of Holes, Cutouts and Damaged Structures

...

Figure 17.36 NASA Langlc) tc't' on a'ially comprc~'ed graphite-


cpox) C) lindrical paneh wnh circular cutouts-typ•cal
mont fnngc pattcm5 of 'pec•men CN \\llh !R/t) ~
ll!!5 !from [ 17.821)

local suain-; a-;sociated with severe local bending around the hole lead to local failures. such
ac; dclaminmionc; ncar the hole.'' As mentioned in Section 14.1 of Chapter 14. local delam-
ination failures of thi type are unique to laminated compol.itc structure!. and can propagate
to cause fai lure of the laminate. The resulls of th1s inve~tigation indicate that the specimens
with the smaller panel radius are subjected to higher strains than those with the larger radius
and are therefore more likely to fail locally ncar the cutout. after buckling.
The rC\Uit' of the other studies on compoc;ue cuncd panels. mcmioncd in the beginning
of thi' subsection. in general confinn the bcha' 10r ob!.en·ed by Knight and Starnes in their
NASA Langley inve~tigalion.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
17.4 Stability and Strength of Damaged or Dented Shells
17.4. 1 Buckling and Strength of Damaged Structures
In the eightie~ conc;idcruble effort w:tc; devoted in many labomtories to the study of the
buckling and collap:-.e behavior of thin-walled ring and stringer-stiffened shell'> of geometries
relevam to those in oO\hore engineering practice. TI1c, c inve tigation~ included the effect of
accidental damage. such a that inflicted b) rn1nor collision~ between an ofl'~>hore ~tructure
and anendant ve'>scls. Since relati\ely little wa' known on the influence on -.hell ~trcngth.
behavior and ~tructural integrity. re~ulting from operational accidenb hec for example
[17.92]- ( 17.94 )). cxtcn-;ive studie<. have been carried out at several rc<;carch center" on the
strength and buci..Jing behavior of damaged and dented ~hell<> (sec for cx<unplc [ 17.95]-
(17.1 12]).

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Stability and Strength of Damaged or Dented Shells 1451

These studies considered the damage process (as in [17.92], [17.93], [17.99] and {17.94]).
proposed plastic denting mechanisms (see for example [17.95], [17.98] and [17. 10 I]) and
involved many tests on damaged and dented model shells (see for example [ 17 .95]-[ 17.98] ,
[17.100)-[17.104], and (17.107]-[17.109]). Most of the experimental investigations dealt
with ring- and stringer-stiffened shells that were initially damaged and then subjected to
combined compressive axial and extemal pressure loading (see for example [17.96], [ 17.98],
[17.101), (17.102] and [17. 108]).
Shells represent a type of structure that Jacks an overall structural redundancy and are
therefore very sensitive specimens for assessment of the effect of local damages. Hence,
most damage-effect-tests were on shells and stiffened shells, and only very few dealt with
plates (as for example part of [17.98]) and other structural elements.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

In [17.104] it was pointed o ut that one type of accidental loading that is "of pm1icular
relevance to thin-walled shells is the situation in which a collision of moderate energy occurs
between the platform and an attendant vessel. The energy is too low to cause extensive
damage to the structure but adequately large to result in local damage to the relevant struc-
tural member." It has been shown [17.94] " that this happens sufficiently often to be a source
of potential hazard to the structures now on-station in offshore locales." Because of their
imperfection sensitivity, shell elements of offshore structures may, however, be vulnerable
to these low-energy collisions.
"Damage effects are difficult to define precisely, since they can be the result of a multi-
plicity of events within a collision. The ship can strike the structure leg with its bow, its side
or its stern. The ship structure can deform and absorb part of the energy, or it can be relatively
rigid and the less rigid structure may suffer all the deformation."
In most of the experimental studies in the eighties, the damage was simulated primarily
by a knife-edge dent imposed midway down the shells, at right angles to its axis (see for
example [17.98]- [ 17.1 07]). The knife-edged dent was chosen because it seemed an appro-
priate modeling of an accident in which the colliding ship was infinitely rigid compared to
the shell, in which case the struck stiffened shell structure would absorb nearly all the
collision energy. The use of a rigid knife-edge indenter was considered to inflict damage on
the shell that would be more severe than that in the corresponding practical collision.
One objective of the research carried out was to develop simple methods by which de-
signers could estimate the collision resistance (or residual strength) of orthogonally stiffened!
cylindrical shells, such as those employed in fixed or ftoating offshore structures. Plastic
mechanism analyses, which were verified by model tests, were therefore developed to provide
such design methods (see for example (17.98], [17.101] and [17.102]). T he details of the
denting tests are discussed in Subsection 17.4.2, which follows.
In the development of such a mechanism analysis. Ronalds and Dowling fust considered.
the stiffened cylindrical shell to be made up of longitudinal stringer beams, each consisting
of a stringer and its corresponding strip of shell, spanning across ring frames (see [ 17 .l 0 1]).
Then the response of the ring components was dealt with separately, and fi nally the combined
response of the beam and ring components of the stiffened shell was calibrated, using their
test results on dented short three-bay models subjected to axial compression.
Besides the offshore structure-boat collisions, there are other cases where local dents can
strongly influence the buckling strength of structures. For example, damage in the form of
dents has been found to be one of the most common causes of deep-water pipeline failures
(see [17.110]-[17.114]). As pointed out in [17.111], "dents can occur due to impact by
foreign objects as well by other causes during transportation of the pipe, during installation
and trenching, and during the operation of the line. Dents reduce the local collapse capacity
of the pipe and can be the initiators of propagating buckles [discussed in Section 18.3 of
Chapter 18]. In addition, for gas lines under high internal pressure, dents, sometimes accom-
panied by gouges, are often the prelude to burst failures" (see for instance [17.113] and
[17.110]).
Hence, several experimental studies on the collapse capacity of dented pipes have been
can·ied out. Various types of indenters were employed in these experiments to pre-dent the

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1452 Influence of Holes, Cutouts and Damaged Structures

tubes that were then subjected to external pressure: Kyriakides et al. [17. 11 5) used point,
plate and knife indenters on steel tubes, Estefen et al. [17.116] used knife indenters on
aluminum tubes, and Park and Kyriakides [ 17.Ill] used spherical head indenters of two
different d iameters.
This experimental study of Park and Kyriakides (Ref. 17.11 1) represents a good example
of the careful application of present-day test techniques with concurrent numerical simulation
using nonlinear elastoplastic shell analyses. It therefore warrants a more detailed discussion.
The experiments were conducted on commercially available stainless steel (SS) 304 seamless tubes
with nominal diameters of 1.25 in. (-31.8 mm) and nominal Dlt values of 18.9. 24.2 and 33.6.
The tubes came in lengths of approx imate ly 24 ft (- 7.3 m). The test specimens were cut from
these tubes to le ngths corresponding to approltimately 29 tube diameters. Tihe diameter (D) a nd
wall thic kness (I) were measured at several positions along each test specime n and their variatio ns
were established.

Each test specimen was dented at mid-length with rigid indenters with a spherical head
of diameter d (that was smaller or larger than the tube diameter D).
The indention was carried out in a standard, screw-type electromechanical testing machine with
the test specimen resting on a rigid, flat plate. A thin (0.063 in. - 1.6 mm) rubber pad was placed
between the tube and the rigid plate. to help prevent rotation of the tube during the early stages
of the indentation. Test specimens from each of the three Dlt value~ studied, were indented to
varying degrees to final indention depths in the range of c\, < 0.4D (where So is the maximum
dent depth after unloading). Indenters with diameters d = 0.4D and 1.60 were employed. ln each
ca,~e. a sufficient number of tubes (8-12) were indcll!cd with the indenter with d - 0.4D to ensure
that the trend of the behavior was captured with accuracy. An additional tllree tubes from each
family were indented to varying degrees with the indenter with c/ = 1.60. Figure [17.37] shows
three tubes indented to varying degrees with the larger diameter indemer.
The prescribed displacement (o) and the required load were recorded for each indention test and
used later to check the accuracy of the corresponding analysis. An eMmple of one s uch force-
displacement record, from a tube with Dlt of 24.2, is s hown in [Figure 17.38] (note the good
correlation between eltperiment and analysis). In this case, the tube was indented to a final value
of o0 - 0.330. (The small nonlinearity observed in the unloading part of this response, to a large
extent is due to the presence of the rubber pad.) ... [T]he majo r geometric parameters of each
dent are its maltimum depth o., a nd the measure of the ovality of the most deformed cross section,
Ao<~• defined as follows:
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

( 17 .2)

where D"'"' is the maximum distance across the deformed cross section and D,.i, is the minimum
distance across the convex part of the deformed cross section.

These dent parameters were measured and recorded for all the specimens. A more detailed
record of the geometry of each dent was obtained by using a circumferential scanning device
to record several cross-sectjons of the dented tubes in the neighborhood of the dent. The
scanning device was based on the same principles as the earlier Caltech initial imperfection
measurement system, discussed in Section 10.3 of Chapter 10. After suitable data reduction,
sequences of cross-sections were recorded, together with their position along the most de-
formed generator of the tube, as shown for a typical specimen in Figure 17.39. Though the
larger-diameter indenter (with d = 1.60 ) was used in this case. the dent was quite localized,
extending primarily over a length of three tube diameters. The small asymmetry observed
in the figure in profiles 8 and 9 was apparently caused by some misalignment between the
tube and the indenter.
Returning to Figure 17.37a. showing the indentation with the larger-diameter spherical
head d = 1.6D, with indentation degrees or depths in the ratios I :4.1:7.4: the sequence of
the tubes in Figure l7.37a and 17.37b is the same, indicating that the elttent of collapse in

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Stability and Strength of Damaged or Dented Shells 1453

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Figu re 17.37 University of Tcxa.o; tests on dcmed cylinders under external prcssurc- photogmphs of
three dents produced by the latge spherical head indentor, with d - J.6D, o n the tubes
with Oft - 18.9 (from f 17.111 ]): (a) the three dent~ (indentatio n depth~ in the ratios I:
4.1 :7.4), (b) the tube~ of (a) after collapse by external pressure. Note that the sequence
of the tubes i~ the .,arne in (a) and (b). and that the extent of collapse in (b) is the inverse
of the site of the dent; in o ther words. the collapsed section is larger f<)r ,maller dents

2.0

F $S-.3G4 8 F
(Idol) ~ •24 2 ()H)

I
1.5
~·0.4
I &

1.0

0.5

-
. ~
.AI'O/)'M
!.
D

0
0. 1 02 O.J o.• 0.5
510

Figure 17.31l Univcr,ity of Texas tests on dented cylinders under external pressure- typical indentor
force-displacemem record. T1te indentation force-ma)(imum depth response shown is for
the ;mailer spherical head indemor. with diameter d = 0.40. Note that the analysis
predictions correlate well with the experimental point; (from 11 7. Ill ])

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1454 Influence of Holes, Cutouts and Damaged Structures

888
88~
~ o -..J
88c:J

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Figure 17.39 University of Texas tests on dented cylinders under external pressure- cross-sections of
a dented tube, indented with the large diameter indentor of d = 1.60 (from r17.111))

(b) is the inverse of the size of the dent in (a); or in other words. the collapsed section of
the tube is larger for the smaller dents!
To explain this apparent paradox one should recall that because of the high stiffness of
the pressurization system and the nearby volume-controlled pressurization scheme used, the
collapse is local in nature. affecting sections of the tube that are only a few tube diameters
long. Tubes with relatively small dents have higher collapse pressures and, as a result, more
energy is transferred in the case of rhe smaller dent~ from the pressurizing system to the
collapsing tube. Thus, the extent of the collapse is larger for the case of smaller dents, as
seen in Figure 17.37b.
Park and Kyriakides drew the following main conclusions from their comprehensive ex-
periments o n 37 specimens:
" I. For all three Dl1 tube families used, denting was found to reduce the collapse pressure
quite significantly. Dents with values of tJ.(Jd of approximately 0.1 c ause a reduction in
the collapse pressure of the tube which is on the order of 50 percent. The trend is that
as 6.011 increases the collapse pressure tends to approach the buckle propagation pressure
(P).
2. When the measured collapse pressures, from each D I 1 fami ly. were plotted against the
dent parameter tJ.,ld, the results from the small diameter and the large diameter indentors
were found to have an excellent correlation. In view of the large d ifference in the two
indentor diameters, this indicates that !lOt/ is a successful measure of the severity of l'he
dent as it affects the collapse pressure. If we couple this with similar observations
made in [ 17 .115) for a much broader range of dent types and geometries, this geometric
measure of the severity of dents can be considered to have universal applicability for
local dents.
3. The experimental results generated in the present study, as well as those of Kyriakides
et al. [17.115], indicate that local dents with flo,1 < 0.5 percent are quite benign from
the point of view of their effect on the collapse perfonnance of a pipe and can be
neglected. Their effect on other performance characteristics of pipelines should, how-
ever, be evaluated independently."
Concurrently, the tube denting and collapse processes were simulated through a finite-
element shell analysis of the problem. The fonnulation used allows for finite defonnations
of the shell. The measured geometric and material parameters of the tubes tested were
adopted in these simulations, and the results of the analysis were found to correlate well
with the experimental ones.

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Stability and Strength of Damaged or Dented Shells 1455

17.4.2 Damaged Stiffened Shells


As part of the comprehensive U.K. Department of Energy S tiffened Shell Study for Offshore
Structures, initiated in the late seventies (and discussed in Sections 13.2, 13.3 and 13.4 of
Chapter 13), extens ive experimental investigations on the buckling and postbuckling behavior
of damaged stiffened shells were carried out at Imperial College London and the University
of Surrey in the eighties. These investigations are good examples of careful test programs,
closely CotTelated with substantial numerical computations. The Imperial College tests dealt
with axial compression loading (see [17.95]-[17.98). [17.101]-[17.103] and [17.107]-
[17.109]), while the accompanying University of Surrey tests treated the case of external
pressure and combined external pressure and axial compression loading (see [17 .99],
[17 .104]-[17.106]).
Both programs employed small-scale welded steel stringer- and ring-stiffened shells, sim-
ilar to those tested intact and discussed in Subsections 13.2.2 and 13.4.1 of Chapter 13. As
pointed out there, the Surrey welded shells (originally developed at University College,

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
London; see Figure 13.18) and the Imperial College ones (see Figure 13.21) were fabricated
by somewhat different techniques, but were of similar quality.
The tests in both programs involved three phases:
I. Measurement and recording of the initial geometric imperfections of the specimens
2. Introduction of a controlled dent
3. Testing to failure of the damaged models, under compression at Imperial College and
under external pressure or combined loading at the University of Surrey
In the initial Imperial College program [17.101]
four cylindt·ical model shells were built with longitudinal stiffeners, separated into three equal bays
by two intermediate ring frames. All specimens had a diameter of 320 mm (- 12.6 in.) (R It) =
267 and bay length L = 96 mm (- 3.8 in.). Two shells had 40 stringers and two only 20. The
models were fabricated [see [ 13.70) of Chapter 13] by welding shell and stiffener elements together
while they were restrained to minimize distortion. The complexity and expense of the jigging meant
that only very limited variations in the geometry were possible. For the same reason both rings
and stringers were fiat-bar stiffeners. When the models were complete they were stress relieved.
Although the steel was extremely thin, 0.6 mm. a special roiJing process combined with later heat
treatment gave the material mechanical properties representative of structura] steel.
The shells were then damaged or dented in the denting rig shown in Figure 17.40. Lateral
load was applied centrally though a 60° knife-edge wedge with a 3-mm tip radius, lying
across the lo·ngitudinal stringers. The denting process was slow because the strain gauges,
displacement transducers and a load cell were monitored after each small deflection incre-
ment. Heavy end rings on the test cylinders were bolted to thick e nd blocks, which were
supported on rollers and sliding bearings (see Figure 17 .40). This arrangement ensured that
the model ends were completely free to pull in longitudinally and rotate. However the pla-
narity enforced by the blocks gave the dent ends flexural fixity and considerable axial re-
straint. The resulting longitudinal pull-in of the ends of the dent was measured using trans-
ducers.
The residual compressive stiffness and strength of the dented shells were determined ex-
perimentally by testing the specimens under axial compression. These tests were carried out
in a relatively stiff test frame developed for the earlier intact small-scale welded steel stringer-
and ring-stiffened shells (see Figure 13.40 and the relevant parts of Subsection 13.3.2 of
Chapter 13 a nd [17.100]). As a result of the denting process, the model ends were not
perfectly straight and parallel, and therefore a spherical bearing was inserted in the test rig
above the test model to take up any unevenness and allow further rotation during the com-
pressive load.ing. The screw jack of the test frame was driven by a variable-speed electric
motor through a low-ratio gearbox, which allowed the very slow loadjng rates needed for
the denting process. The buckling deformations, as well as the initial geometrical imperfec-

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1456 Influence of Holes, Cutouts and Damaged Structures

Figure 17.40 Imperial College axial compression tests of dented stiffened shells- knife-edge deming
rig (from [17.96])

tions, were measured with the aid of a bank of 40 radially aligned transducers, set up around
the model shell, that could also be moved vertically.
A total of 18 damaged stiffened shells were tested in the various phases of the Imperial
College experimental investigation (see [17 .96], [17.1 00], [17.101] and [ 17.108]), nine of
these were ring stiffened only, whereas the remaining 9 had both stringers and rings. The
specimens in all the phases of the program were 320 mm in diameter, between one and five
bays in length, their (Rit) = 133- 267, their (LIR) = 0 .15-1.08, and they were dented to
residual central dent depth ratios 8D<,!R = 0.021-0.107.
The observed residual strengths of the shells P,/P 0 (where P,1 is the collapse load of the
dented shell and P 0 is that of the corresponding intact one) are plotted in Figure 17.41 as a
function of their central dent depths ratios (8rx/ R). The strength losses due to damage were
found to be less severe than might have been expected.
The manner in which the axial compression is applied is important, and there are signif-
icant differences between load control (shown in Figure 17.41) and displacement control,
the suitable method depending on the actual boundary conditions. Though the knife edge
dent was local, the resulting damage spread over a number of bays, unless the ring stiffeners
were rather stiff.
The buckling mode shapes were studied in detail (see Ref. 17.108). Both local and general
buckling modes were observed, but often the dent promoted general buckling.
In the University of Surrey program similar small-scale dented ring- and stringer-stiffened
welded steel shells were tested under a combination of external pressure and axial com-
pressive loading. The test specimens were again identical to the small stiffened shells em-
ployed in earlier experimental programs on intact specimens (discussed in Subsection 13.4.1
of Chapter 13).
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`-

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Stability and Strength of Damaged or Dented $hells 1457

0..0
-~~-:.-.:::__A...~--._-
o.."
<;'

~
0 0 -------::--

-----
;;; o·
..,·;;g K<"'
!>'IIIIHU£~ STIH"Ii:N£C

~ t. OC45 . ;:,ce6 .:;c

"

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
~lNG Srlf: F!N£0
- K'"'~O
-- 0<>5
00
- ·- 1)'0

006 012
Central d<nt dept:1, 6,1~

Figure 17.41 fmperial College axial compression tests of dented stiffened shells-residual compressive
strength of damaged shells versus central dent depth ratios (from [17.108] and [ 17. 109)).
P,1 is the collapse load of the dented shell, while P0 is that of the corresponding intact
one

The geometries of the six test specimens are presented in Figure 17 .42. Two were plain
ring-stiffened cylinders (Rl and R2), two were T-ring s tiffened cylinders (R3 and R4) and
two were ornhogonally stiffened cylinders (RS and R6). The models all had at least three
inter-ring bays, and the knife edge damage was imposed at midheight.
''The intention of the model design was to provide information on a wide spread of
parameters, within the limited numbers of tests, and to el.iminate as far as is possible the
effects of the boundary conditions on the buckling behaviour of the cylinders in the vicinity
of the damaged zone" [17.105].

Shells R1 8. R2

Shells R3 8. R4

Shell RS INs: 40 )
Shell R6 ( ~= 20 I

INfi ~:I •-t~ I...,g


Figure 17.42 University of Surrey tests on the strength of damaged ring and orthogonally stiffened
shells-geometries of test shells (from [17.105)). All dimensions are in mm

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1458 Influence of Holes, Cutouts ;;~nd Damaged Structures

The shell geometries were designed to be fairly re presentative of the thin-walled com-
ponents of semi-sub and tension leg platform (TLP) structures in offshore practice of the
eighties.
Particularly, the plain ring stiffeners were dimensioned such that elastic torsional buckling (tripping)
should not occur. The particular interest in testing models Rl and R2 was to detennine the effect-
iveness of the plain ring stiffeners in preventing the localised damaged from extending (under
pressure loading) into adjacent regions of the shell.
... The geometry of the T-rings in R3 and R4 was such that it had about the same value of
second moment of area as the plain rings in Rl and R2. The objective in the tests on cylinders R3
and R4 was to determine the possible improved efficacy of the T-rings in arresting the developmentr
of the damage zone. This improvement could result from the greater torsional stiffness of the T-
ring compared to the plain rings, even with the same flexural rigidity. l n an attempt to widen the
range of parameters studied, the inter-ring bay length was reduced in R3 and R4 compared to Rl
and R2. However, to offset this factor the damage was applied to T-ring stifl'ene.d cylinders at the
junction of a ring and the skin. It was considered that this would create locally a longer bay-length
and thus provide a reasonable comparison with RJ and R2.
_The design of the orthogonally stiffened cylinders was also intended to reduce the influence
of the boundary conditions on the buckling behaviour. The stringers' end rings were dimensioned
such that elastic tripping would be eliminated and these members should attain their full plastic
strength. The shells R5 and R6 differ in geometry significantly in that the former had 40 stringers
and the latter only 20. Previous test geometries of undamaged shells have shown that with 40
stiffeners the shells behave in a quasi-ortbotropic manner, with the buckles which formed during
collapse encompassing several stiffeners. On the other hand, shells with 20, or fewer, stiffeners
buckled such that the stiffeners acted as barriers between buckles and the major dcfonnntions were
restricted to the longitudinal panels between the stiffeners. The primary objective of the tests on
shells R5 and R6 was to determine if the presence of significant damage would effectively alter
this pattern of deformation."

Each cyLindrical shell had its ends securely attached to re latively rigid end rings, attached
to the shell either with a very strong sand-Araldite mixture or by welding.
"The ring stiffened models were manufactured by rolling sections of sheet steel into
cylinders and welding the seam. The cans thus fonned were welded to the ring stiffeners to
form the complete shell. A stress relieving procedure was performed subsequent to this
welding operation."
Similarly, the stringer- and r ing -stiffened shells "were fabricated using a technique in
which all the residual stresses, incurred during the welding of the stiffeners to the skin panels,
were eliminated by stress rel ieving. This operation was performed prior to welding the shell
to the end rings. Hence some level of residual stresses could exist at the junction between
the shell and the end rings,'' but it was expected to be insignificant on account of the
geometry of end rings.
"The inevitable geometric imperfections of the models were measured by rotating the
completed shells on a turntable relative to a fixed linear voltage displacement transducer
placed at various levels down the shell length." The data were reduced to the corresponding
Fourier coefficients, as detailed in Chapter 10. The specimens were found to have very small
imperfection values.
The damage was imposed on the sheUs (after their fabrication imperfections had been
measured and after strain gages were affixed to skin and stiffeners at the appropriate loca-
tions) in the form of a flat-bottomed knife edge dent. T he denting apparatus is shown in
Figure 17.43 and was such that the load itnd depth of dent could be carefully controlled and
monitored.
The knife-edged indenter was very slowly pressed on to the surface of the shell, the line to the
knife-edge being normal to the centre-line of the shell. At appropriate values of deflection, the
loading was decreased and the indenter removed from the shell. Thus several values of loading
and corre:sponding levels of applied and residual dent depth were obtained. In one or two shells a
layer of brittle lacquer was deposited on the outer surface of the shell in the region of intended

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Stability and Strength of Damaged or Dented Shells 1459

Fi.:ure 17.43 Um,erstty of Surrey tc">l\ on the ,trength of damaged ring and orthtlgonally ~uffened
~hell\ t.lcnting apparatu' (frotn (17. 10511

t.lamagc. Tht' provided Htlunble information on the manner 111 \\hich the damaged tone extended
n' the dent depth wa' incrca cd.
Suh>cqucntto the dcnung. the shell~ \\ere funhcr \tram gaged at locatton\ where 11 wa\ expected
buckling would o.:cur.

The model\ were then mounted in the Surrey h) pcrbaric chamber (descrihed 10 Sub ection
13.4.1 of Chapter 13: see Figures 13.-17 and 13 48). which can apply :vaal compres tve
loading to the \ hell' ~imultaneously with e\ternal pre,.,ure.
In their two paper:. 11 7.1051 and [ 17. 106) Wall.er, McCall and Thorpe analyL.ed the col-
lap e behavior of their six shells in great detail and provided useful infonnation for hoth
experimenler.., and de~igncrs.
In their )>Ccond paper [ 17. 1061. they drew a number or conclusion~ from the result~ of the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

tesL~. mo\1 of which arc reproduced bclo'':

"1. The plain ring '>tiffened model' were much less able to resist the effect of impact
loadmg than the T-ring stiffened mo<kl-.. TI1c effect of denting wa-. to cause local
tripping to occur in the plain nngs. \\hach <.uh<,equentl} undennined the ~helb ·capacity
to suppon prc<,<,urc loading.
2. E'cn for very low Je,eb of denting load. the plain ring ~tiffened model exhibits plastic
str.unin£. Moreover. the effect of the plain rings was negligible m re,traming the de-
velopment of a pla\tic damage l.tme. Tht!> ts evidenced by the correlation between the
predictions of a ~i mp l ificd analy~i!>. ha~cd on an un~tjffened shell, and the results of
the plane ring-stiffened shell.
3. The pl;1in ring-~tiffened shell collapsed locally when subjected C\en to moderate ex-
ternal prc">urc loading. The pJa.,ticuy mcurrcd during the denting process had so un-
dermmcd the radial load-bearing capaCII) of the rings Ihat con iderablc load wa~ placed
on the adjacent ring .... When the} in tum collap">cd. due to local tripping. a ca~cadc
effect wa<. <.et off. If the shell had been longer. and incorporated more ring stiiTeners.
the local buckle would have extended through all the ring stiffeners to span the full
length of the \hell.

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1460 Influence of Holes, Cutouts and Damaged Structures

4. The extent of the initiaE buckling of the plain ring-stiffened shells was limited only by
the provision of adequately stiff and strong end rings. It is recommended that care
should be taken in practical design to provide such a support to the shell at appropriate

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
positions along its length.
5. The plain ring stiffeners were adequately stiff and strong in the undamaged state to
prevent general buckling when the shells were subjected to the combined pressure and
axial loading used in the tests. In other words, the failure of the shell was associated
by local inter-ring buckling, as would be intended by the design engineer. This intention
is evidently undermined completely by the presence of even slight damage, and it is
recommended that attention be given to an investigation of the effects in shell strength
of accidental damage to the plain stiffeners themselves.
6. The behaviour of the T-ring stiffened cylinders was much more satisfactory from a
practical viewpoint. The T-rings used in the models, although having approximately
the same value of second moment of area as the plain ring stiffeners in other shells,
had a much higher torsional stiffness. No evidence was noted of any local plastic
buckli.ng and the shells were able to support a much higher loading than the corre-
sponding plain ring-stiffened shells, while collapse was associated with local inter-ring
buckling with no significant extension of the original damaged zone.
7. The design ofT-rings incorporated in the test models was obviously adequate to prevent
local buckling, etc. However, a more optimum design would be desirable and certainly
engineers require more information to be able to dimension the geometry of the T-
stiffeners with assurance. There is a requirement for more research to be performed to
generate this information, and it is recommended that fu.rther analysis and some testing
be ins.tituted.
8. The orthogonally stiffened shells were fabricated with plain stiffeners which were ad-
equately compact and strong to ensure that elastic tripping should not occur and that
buckling was limited to inter-ring deformation. However, it may be concluded from
the results of the tests that the presence of damage overloaded the rings locally, which
collapsed and this potentially would give rise to the cascading effect noted above.
9. Only two orthogonally stiffened models were included in the test programme, and these
were subjected to relatively large dent depth compared to the ring stiffened shells.
More information is required on the effect of depth on the strengtlh of the shells. It is
recommended that further research, analytical and experimental, be initiated to deter-
mine if a critical dent depth exists, below which the presence of damage of the shell
strength is negligible and. therefore, can be ignored in practice."

Walker and his co-workers caution about the limited validity of their conclusions, since
they were based on a few tests only.
Notwithstanding, readers may wish to consider this form of detailed conclusions for re-
porting on their own work.
Before we conclude this subsection, it may be of interest to consider briefly some recent
buckling experiments on locally damaged cylindrical shells carried out by Hambly and Cal-
ladine at Cambridge University [17.117]. Their study focused on the effect on the buckling
of a cylindrical shell of the local damage or imperfections caused by some s011 of impact.
They envisaged an event that produces a severe indentation of the shell and that is followed
by art attempt to restore the shell to its original fonn by pushing out the indentation somehow
or other. The behavior of a typical empty drink can (usually having R It 350) when held =
in the hand and pushed in on the side by a thumb is a suitable example. " In this case, the
deformation has a strongly ' inextensional' character, in the sense that if the indentation is
pushed out, the can returns to much the same shape as before. There is little sign of wide-
spread surface straining, and yet highly localized regions of damage are evident."
The indentation produced by the thumb has the form of a crease (see Figure l7.44a) with
some localized compression (crinkling) damage at its ends. When the shell is restored again

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Stability and Strength of Damaged or Dented Shells 1461

(a) (b) (c )

Figure 17.44 Cambridge University buckling experiments on locally damaged thin-walled cylindrical
shells- effect of thumb-pressed indemation (from [17.117]): (a) empty drink cans with
a thumb-pressed indentation of width s, (b) after the indentation has been pressed out,
damage in the form of two dings (shown schematically by double lines representing
localized creases a nd small dimples or depressions beyond them) remains at the ends of
the original indention, (c) after rotation of the can and application of an eccentric ax.ial
load F, to ensure initiation of buckling at one ding; with increase of the load F the
dimple grows and becomes a diamond-shaped dent

to its original shape, these localized regions of crinkling damage " dings" remain (see Figure
17.44b).
Such localized damage is basically different from the usual geometrical imperfections
caused by regular fabrication processes, which can be decomposed into harmonic compo-
nents (as discussed in Chapter 10). The Cambridge exper imental investigation aimed at
clarification of the difference in the influence on buckling between these two types of im-
perfectio ns, the usual geometrical ones that cover most of the shell and the localized ding,
such as that left after severe impact damage.
With the aid of a simple test setup that allows eccentric loading to ensure that buckling
w ill initiate a t a ding, the difference in behavior between damaged thin shells and undamaged
ones could be evaluated. It was observed that as the axial compression increased, the local
ding or dimp le g rew and became a diamond-shaped dent. In general, the ding closer to the
principal gemerator (which i s itself determined by the eccentric location of the load), being
a region of h igher compressive stress, beca me the focus of buckling.
Hambly and Calladine carefully followed the buckling process of many locally damaged
and undamaged steel cans (wi.th R = 33 mm, t = 0.095 mm and L = 95 mm). As explained
in Subsection 9.3.6 of Chapter 9, such drink cans, bec.ause of the quality control required
by the ir a utomated filling and sealing process, represent inexpensive specimens with accu-
rately repeat,ed dimensions.
From their observations they concluded that

the most obvious point to be made at this stage concems the sharp contrast between the behaviour
of damaged and undamaged nhin-shell specimens under load. The damaged shells reached, on
average, about one quarter of the buckling load predicted by the classical theory, while the unda-
maged ones reached about one half. But much more strildng than this quantitative comparison
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

between the two sets of tests i.s the qualitative difference. Thus, while the undamaged specimens
gave no indication that a catastl'ophic, unstable failure was imminent, the damaged specimens gave
ample warning, by virtue of the steady growth in the size of the dent as the load increased. And
indeed, the rule that catastrophic failure occurs when the diamond-shaped dent has reached a unique
size enables us to predict the point of instability, while a test is proceeding, with some certainty.

The Cambridge researchers emp hasized the need for further study, but their work again
points out that ofte n even s imple experime nts can significantly broaden our understanding
of the buckling behavior. They closed their 1996 paper by stating:

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1402 Influence of Holes, Cutouts and Damaged Structures

A final point raised by the present investigation concerns the suddenness of the buckling failure of
the specimens. The undamaged cans failed abruptly, and without warning, whereas the damaged
cans had a phase of stable growth of a dent before the catastrophic failure ensued. Now for some
engineering purposes it is deoSirable to have visible, early warning of a failure, even if this is
associated with smaller ultimate strength. There is scope here for investigating the deliberate in-
troduction of "damage" of some sort in order to make the process of budding of thin shells more
benign.

17.4.3 Buckling of Delaminated Composite Panels and Shells


Delaminations defects are generally considered to be the prevalent type of damage initiation
in laminated composite structural elements. A variety of causes, such as manufacturing flaws,
high interlaminar stresses, foreign object impact (bird impact, tool drops, flying runway
debris) and internal fiber kinking or microbuckling due to excessive internal stresses, may
result in the formation of a delamination. If the loading on the delaminated part of the
laminate is compressive, delamination buckling and subsequent growth of delaminations may
occur and provide the mechanism for spreading the damage to the undamaged regions (see
for example [17.74]-[17.77]).
The compressive strength and stiffness of structures made from laminated composite ma-
terials may be substantially reduced by the presence of interlaminar delaminations. The
delaminations may influence the structures in one of two ways: as a plain material imper-
fection or by local delamination buckling. The latter phenomenon usually occurs when the
initial delamination allows the regions bounded by a delamination and the free surface of
the laminate to buckle locally and thereby create conditions that promote delamination
growth (see for example Section 14.1 of Chapter 14 and [17.74], [17.76], [17.77], and
[17.119]-[17.121]).
Delamination buckling and delamination growth have been the subject of extensive study
in the last two decades because of their important influence on the safety or damage tolerance
of laminated composite structures. Because of the inherent difficulties (such as the simulation
of the original delaminations and the large number of parameters that affect the initiation of
delamination buckling), only a small portion of the work has been experimental. A detailed
discussion of these investigations, even just the experimental ones, would require a separate
and voluminous chapter. However, since much of the work on delamination buckling and
growth is still in progress, it has been decided to leave the topic for a future edillon of the
book and present here only a selected list of references that may guide the readers in their
initial acquaintance with delamination buckling and delamination growth.
Fairly extensive reviews may be found in [17.74]-[17.77] and [17.122]-[17.127]. Some
of the earlier experimental studies appear in [17.73], [17.74] and [17. 128]-[17. 134], while
some more recent ones are presented in [17.75]- [17.77] and [17.135]-(17.145]. Many of
these papers and reports include also comprehensive theoretical and numerical investigations.
However, there are many more analytical studies, not accompanied by tests, that are not
listed here.
We concnude this subsection with the observation that our understanding of delamination
buckling and growth and its analysis are still incomplete and definitely require much addi-
tional careful experimental study.

References
17.1 Savin, G.N., Stress Distribution Arou11d Holes, Naukova Dumka Press, Kiev, Ukraine, 1968 (in
Russian), translated by NASA, NASA TT F-607, Nov. 1970.
17.2 Budiansky, B. and Yidensek, R.J., Analysis of Stresses in the Plastic Range around a Circular
Hole in a .Plate Subjected to Uniaxial Tension, NACA TN 3542, Oct. 1955.
17.3 Sobey, AJ., The Estimation of Stresses around Unreinforced Holes in Infinite Elastic Sheets,
Aeronautical Research Council (UK), R&M 3354, Oct. 1962.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1403

17.4 Yu, W.W. and Davis, C.S., Cold-Formed Steel Members with Perforated Elements, Journal of
the Structural Division, Proceedings ASCE, 99, No. ST 10, Oct. 1973, 2061- 2077.
17.5 Lekkerkerker, J.G., Stress Concentrations around Circular Holes in Cylindrical Shells, in: PIV-
ceedings oft he Eleventh lmernarional Congress ofApplied Mechanics, Munich, 1964, Springer-
Verlag, Berlin, New York, 1966, 283- 288.
17.6 Van Dyke, P., Stresses About a Circular Hole in a Cylindrical Shell, A!AA Journal, 3, 1965,
1733-1742.
Ebner, H. and Jung, 0., Stress Concentration around Holes in Plates and Shells, in: Contribu-

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
17.7
tions to the Theory ofAircraft Structures, Professor Arie van der Neut 65'h Anniversary Volume,
Delft University Press, 1972, 135-158.
17.8 Schlack, Jr. , A.L., Elastic Stability of Pierced Square Plates. Experimental Mechanics, 4, (2),
1964, 167- 172.
17.9 Fujita, Y., Yoshida, K. and Arai, H., Instability of Plates with Holes, 3rd Report, Journal of
the Society of Naval Architects, Japan, 127, 1970. 161-169.
17.10 Narayanan, R., Ultimate Shear Capacity of Plate Girders with Openings in Webs, in: Plated
Structures, Stability and Strength, Applied Science, London, New York, 1. 983, Chap. 2, 39-76.
17.11 Ritchie, D. and Rhodes, J., Buckling and Postbuckling Behavior of Plates with Holes, Aero-
nautical Quarterly, 26, (4), Nov. 1975, 281 - 296.
17.12 Narayanan, R. and Chow, F.Y., Experiments on Perforated Plates Subjec ted to Shear, Journal
of Strain Analysis (/MechE), 20, ( I), 1985, 23-34.
17.1 3 Rhodes, M.D., Mikulas, Jr., M.M., and McGowan, P.E., Effects of Orthotropy and Width on
the Compression Strength of Graphite-Epoxy Panels with Holes, AIAA Joumal, 22, (9), Sept.
1984, 1283-1292.
17.14 Nemeth. M.P., Stein, M. and Johnson, E.R., An Approximate Buckling Analysis for Rectangular
Orthotropic Plates with Centrally Located Cutouts, NASA Technical Paper 2528, Feb. 1986.
17.15 Larsson, P.-L., On Buckling of Orthotropic Compressed Plates with Circular Holes, Composite
Strucwres, 7, 1987, 103-121.
17.16 Marshall. I.H., Little., W., Tayeby, M.M. and Williams, J., Buckling of Perforated C omposite
Plates-An Approximate Solution, in: Proceedinxs of the Second Imemational Conference on
Fibre Reinforced Composites 1986, Institution of Mechanical Engineers. London, 1986, 29- 33.
17.17 Shuart, M.J. and Williams, J.G .. Compression Behavior of :t45°-0ominated Laminates with a
Circular Hole or impact Damage, AIAA Journal, 24, ( I), Jan. 1986, 115-122.
17.18 Marshall, l.H., Little, W. and EI-Tayeby, M.M., Membrane Stress Distributions in Post-Buckled
Composite Plates with Circular Holes. in: Proceedings of Sixth InternCitionaf Conference on
Composi1e Maleria/s, London, July 20-24, 1987, 5.57- 5.68.
17.19 Ravinger, J., Girders with Holes in Webs, in: Stability of Plate and Shell Structures, Proceedings
of an ECCS International Colloquium, April 6-8, 1987. Ghent University, Belgium, P. Dubas
and ID. Vandepitte, eds., Ghent University 1987, 101- 106.
17.20 Hom , W.J. and Rouhi, M., A Comparison of the Post-Buckling Behavior of Metallic and
Composite Plates with Centrally Located Cutouts, National Institute for Aviation Research ,
Wichita State University, Wichita, Kan., NIAR Report 89-11, July 1989.
17.21 Nemeth, M.P., Buckling Behavior of Compression-Loaded Symmetrically Laminated Angle-
Ply Plates with Holes, 1\/AA Journal, 26, (3), March 1988, 330- 336.
17.22 Nemeth, M.P., Buckling and Postbuckling Behavior of Square Compression-Loaded Graphite-
Epox.y Plates with Circular Cutouts, NASA Technical Paper 3007, August 1990.
17.23 Kawai. T. and Ohtsubo, H., A Method of Solution for the Complicated Buckling Problems of
Elastic Plates with Combined Use of Rayleigh-Ritz's Procedure in the Finite Element Method,
in: Proceedings of the 2nd Air Force Conference on Matrix Methods in Structural Mechanics,
AFFDL-TR 68-150, 1968, 967-994.
17.24 Yang, H. T. Y., A Finite Element Formulation for Stability Analysis of Doubly Curved Thin Shell
Structures, Ph.D. thesis, Cornell University, 1969.
17.25 Rouhi, M. and Horn, W.J., A Comparison of the Post-Buckling Behavior of Metallic and
Composite Plates with Centrally Located Cutouts, National Institute for Aviation Research ,
Wichita State University. Wichita, Kan., NIAR Report 90-14, May 1990.
17.26 Narayanan, R. and Chow, F.Y., Ultimate Capacity of Uniaxially Compressed Perforated Plates,
Thin-Walled Structures. 2 , 1984, 241-264.
17.27 Tennyson, R.C., The Effect of Unreinforced Circular Cutouts on the Buckling of Circular Cy-
lindrical Shells Under Axial Compression, Transactions ASME, Journal of Engineering for
lnduslly , 90, (4). Nov. 1968, 541-546.

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1464 Influence of Holes, Cutouts and Damaged Structures

17.28 Brogan, F. and Almroth, B., Buckling of Cylinders with Cutouts, AIAA Journal, 8, (2), Feb.
1970, 236-240.
17.29 Almroth, B.O. and Holmes, A.M.C., Buckling of Shells with Cutouts, Experiment and Analysis,
Intemational Journal of Solids and Structures, 8, (8), August 1972, 1057- 1071.
17.30 Stacnes, J.H., Jr., Effect of a Slot on the Buckling Load of a Cylindrical Shell with a Circular
Cutout, AIAA Journal. 10. (2). Feb. 1972, 227- 229.
17.31 Starnes, J.H., Jr., Effect of a Circular Hole on the Buckling of Cylindrical Shells Loaded by
Axial Compression, AIAA Joumal, 10, ( ll), Nov. 1972, 1466- 1472.
17.32 Starnes, J.H., Jr., The Effect of Cutouts on the Buckling of Thin Shells, in: 71lin Shell S!ruc!ures,
The01y, Experimem and Design. Y.C. Fung and E.E. Sechler, eds., Prentice-Hall, Englewood
Cliffs, N.J.. 1974, 289-304.
17.33 Almroth, B.O., Brogan, F. A. and Marlowe, M.B., Stability Analysis of Cylinders with Circular
Cutouts, AIAA Journal, ll, (I I), Nov. !973, 1582-1 584.
17.34 Schulz, U., Die Stabilitat axial belasteter Zylinderschalen mit Manteloffnungen, Der Bauingen-
ieur, 51, (I 0), 1976, 387- 396.
17.35 Toda, S., Experimental Investigation on the Effects of Elliptic Cutouts on the Buckling of
Cylindrical Shells Loaded by Axial Compression, Transactions, Japan Sociery for Aeronawical
and Space Sciences, 23, (59), May 1980, 57- 63.
17.36 Toda, S., Some Considerations on the BuckJing of the Thin Cylindrical Shells with Cutoi.Jlts,
Transactions, Japan Society fo r Aeronautical and !:,pace Sciences, 23, (6), August 1980, 104-
112.
17.37 Toda, S .. Buckling of Cylinders with Cutouts under Axial Compressio n, Experimental Me-
chanics, 23, (4), Dec. 1983,414-41 7.
17.38 Stru:nes, J.H., Jr., The Effect of a Circular Hole on the Buckling of Cylindrical Shells, Ph.D.
thesis, California Institute of Technology, Pasadena. Calif., I 970.
17.39 Tod!a, S., Buckling of Cylindrical Shells with Circular Cutouts under Axial Compression, Na-
tional Aerospace Laboratory NAL TR-560, Tokyo, 1979 (in Japanese).
17.40 Gavrish. VS., Shapovalov, A.P., Tamurov, N.G. and Tantsura, A.J., Inelastic Stability of Cylin-
drical Shells Weakened by Circular Cutouts, Prik/a.dnaya M eklumika, 7, (l), 1971, 105-109
(in Russian).
17.41 Antonenko, E.V., Geshtarovich, A.l. and Kuptsov, A.N., Stability of Cylindrical Shells with
Unreinforced Cutouts, Prikladnaya Mekhanika, 13, (7), 1977. 117- 121 (in Russian).
17.42 Stames, J.H., Jr., and Shuart, M.J., Composite Fuselage Shell Structures Research at NASA
Langley Research Center. in: Proceedings of the Second NASA Advan.ced Composites Tech-
nology Conference, Lake Tahoe, Nev., Nov. 4- 7, 1991. NASA Conference Publication 3 154.
June 1992, 57- 83.
17.43 Nemeth, M.P. and Starnes, J.H., Jr., The NASA Mo nographs on Shell Stability Design Rec-
ommendations- a Review and Suggested Improvements, AIAA Paper 97- I 302, in Collection
of Papers 38th AIAA/ ASME/ ASCE/ AHS/ ASC Structures, Structural Dynamics and Materials
Conference, Kissimmee, Fla., April 7·-10, 1997. Pan 4, 2639- 2652.
17.44 Babel, H.W., Christensen, R.H. and Dixon, H.H., Design, Fracture Control, Fabrication and
Testing of Pressurized Space Vehicle Structures, in: Thin Shell Strucwres, Theol")\ Experiment
and Design, Y.C. Fung and E.E. Sechler, eds., Prentice Hall, Englewood Cliffs, N.J., 1974,
549'- 603.
17.45 Palchevsk.ii, A.S. and Polyakov, P.S., Experimental Investigation of the Stress- Strain State and
Stability of Rib-Reinforced Cylindrical Shells with Sizable Rectangular Holes, Prikladnaya
Mekhanika, 12, Oct. 1976, ll8-122 (in Russian).
17.46 Jung, 0 ., Der Einfluss von Langsversteiften Ausschnitten auf das Beulverhalten von Zylinder-
schalen, in: Proceedings Hans Ebner Gediichlllis Kolloquium in Aachen, Oct. 1977, Mitteilung
aus dcm lnstitul fiir Leichtbau, RWTH Aachen, April 1978, 240- 267.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

17.47 Miller, C.D., Experimental Study of the Buckling of Cylindrical Shells with Reinforced Open-
ings, Presented at the ASME/ ANS Nuclear Engineering Conference. Portland, Oregon, July
25- 28, 1982.
17.48 Cervantes, J.A. and Palazollo, A.N., Cutout Reinforcement of Stiffened Cylindrical Shells,
Journal of Aircroji, 16. (3), March 1979, 203- 208.
17.49 Durelli, A.J., Stress Co ncentrations. in: Experimental Evaluation of Stress Concentration and
Intensify Factors, G.C. Sih, ed., Marti nus Nijhoff, The Hague, 198 1, 1-162.

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1465

17.50 Mansfield, E.H., Neutral Holes in Plane Sheet-Reinforced Holes Which are Elastically Equiv-
alent to the Uncut Sheet, Quarterly Journal of Mechanics and Applied Mathematics, 6, (3),
1953, 370-378.
17.5 1 Davies. G.A.O., Plate-Reinforced Holes, Aeronautical Quarterly, 18. Feb. 1967, 43-44.
17.52 Mansfield, E.H., Analysis of a Class of Variable Thickness Reinforcement around a Circular
Hole in a Flat Sheet, Aeronautical Quarterly, 21, Nov. 1970, 303- 312.
17.53 Report of Court of Inquil)' imo the Comet Accidems, H.M.S.O., 1955.
17.54 Pugsley, A., The Safety of Stmctures, Edward Arnold, London, 1966, 147-148.
17.55 Kumai, T., Elastic Stability of the Square Plate with a Central Circular Hole under Edge Thrust,
in: Proceedings of the 1st Japan National Congress for Applied Mechanics, 1951, 81-86.
17.56 Meissner. E., Ober des Knicken kreisfonniger Scheiben. Schweizerische Bauzeitung, JOt. 1933,
87-89
17.57 Levy, S., Woolley, R.M. and Kroll, W.D., Instability of Simply Supported Square Plate with
Reinforced Circular Hole in Edge Compression. Journal of Research, NaJional Bureau of Stan-
dards, 39, Dec. 1947.
17.58 Nemeth, M.P., Buckling and Postbuckling Behavior of Compression-Loaded Isotropic Plates
with Cutouts, AJAA paper AIAA-90-0965. in: Proceedings, AIAAI ASMEIASCEIAHSIASC
3/st Structures, Structural Dynamics and Materials Conference, Long Beach, Calif., April 2-
4, I 990, Part 2, 862-876.
17.59 Narayanan, R. and Rockey. K.C., Uhimate Load Capacity of Plate Girders with Webs Contain-

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
ing Circular Cut-Outs. Proceedings Institution of Civil Engineers, Pat1 2, 1, Sept. 1981, 845-
862.
17.60 Narayanan, R. , Der Avanessian, N.G.V. and Ghannam, M.M., Small-Scale Model Tests on
Perforated Webs, Strucwral Engineer, 618, (3), Sept. 1983. 47- 53.
17.6 1 Robens, T.M. and Azizian, Z.G., Strength of Perfora.ted Plates Subjected to In-Plane Loading,
Thill- Walled Stmctures, 2, 1984, 153-164.
17.62 Gilabert, A, Sibillot, P., Sornette, D., Vanneste, C., Maugis, D. and Muttin, F., Buckling Instta-
bility and Pattern Around Holes or Cracks in TI1in Plates under a Tensile Load, European
Journal of Medwnics-AI Solids, 11. ( I), 1992, 65- 89.
I 7.63 VandenBrink, D.J. and Kamat, M.P., Post buckling Response of Isotropic and Laminated Com-
posile Square Plates with Circular Holes, in: Fijih Intemational Conference on Composite
Materials, W.C. HarTigan, J. Strife and A. D. Dhingra, eds., MetallurgicaJ Society, Wanendale,
Pa., 1985. 1393-1409.
17.64 Whetstone, W.D., EAL Engineering Analysis Language Reference Manual, Engineering Infor-
mation Systems, Inc., Saratoga, Calif.. January 1979.
17.65 Lee, H.H. and Hyer, M. W., Postbuckling Failure of Composite Plates with Holes, AIAA Journal,
31, (7), July 1993, 1293-1298; also Collection of Papers, 33d AIAA/ ASME/ ASCE/ AHS/
ASC Structures, Structural Dynamics and Materials Conference, Dallas, Tex., April 13-15,
1992, Part I , 201- 221.
I 7.66 Gi.irdal, Z., Haftka, R.T. and Starnes, J.H. Jr.. The Effect of Slots on the Buckling and Post-
buckling Behavior of Laminated Plates, Journal of Composite.\' Technology and Research, 7,
(3), 1985, 82- 87.
17.67 Hyer, M.W., Rust, R.J. and Waters. W.A., Innovative Design of Composite Structures: Design,
Manufacturing, and Testing of Plates Utilizing Curvilinear Fiber Trajectories, Report VPI-E-
94-08, Virginia Polytechnic Institute and State University, Department of Engineering Science
and Mechanics, Nov. I 994.
17.68 Rouse, M., PostbuekJing and Failure Characteristics of Stiffened Graphite-Epoxy Shear Webs,
in: Collection of Papers, 28th AIAA / ASME/ ASCE/ AHS/ ASC Structures, Structural Dynam-
ics and Materials Conference, Momerey, Calif., April 6- 8, 1987, 181-193, AIAA Paper 87-
0733.
17.69 Rouse, M.. Effect of Cutouts or Low-Speed Impact Damage on the Postbuckling Behavior of
Composite Plates Loaded in Shear, in: Collection of Papers, 31st AIAA/ ASME/ASCE/AHS/
ASC Suuctures, Structural Dynamics and Materials Conference, Long Beach, Caljf., April 2-
4, 1990, 877-891, AIAA Paper 90-0966.
17.70 Saiz, J.M.B. and Cardaba, A.B .. Nonlinear Analysis of Composite Shear Webs with Holes and
Conelation with Tests, in: AGARD. Analyrica/ Qual(fication of Aircraft Structures, April 1991.
17.71 Nemeth. M.P.. Buckling and Postbuckling Behavior of Laminated Composite Plates with a
Cutout, NASA Technical Paper 3587. July 1996.

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1466 Influence of Holes, Cutouts and Damaged Structures

17.72 Duke, J.C., Post D., Czamek, R. and Asundi, A., Measurement of Displacement Around Holes
in Composite Plates Subjected to Quasi-Static Compression, NASA Contractor Report 3898,
June 1986.
17.73 Starnes, J.H. and Williams. J.G., Failure Characteristics of Graphite-Epoxy Structural Compo-
nent-s Loaded in Compression, NASA TM 84552, 1982.
17.74 Chai, H., Babcock, C.D. and Knauss. W.G., One Dimensional Modeling of Failure in Laminated
Plates by Delamination Buckling, lnlt'malional Joumal of Solids and Struclures, 17, 19&1,
1069- 1083.
17.75 Waas, A.M., Babcock, C.D. and Knauss, W.G., An Experimental Study of Compression Failllre
of Fibrous Laminated Composites in the Presence of Stress Gradients, lntemational Joumal of
Solids and Structures, 2.6, (9/10). 1990, 1071-1098.
17.76 Davidson, B.D., Delamination Buckling: Theory and Experiment. Journal of Composite Ma-
terials, 25, 1991, 1351-1378.
17.77 Com.iez, J.M., Waas. A.M. and Shahwan, K.W., Delamination Buckling; Experiment and Anal-
ysis, buernational Journal of Solids and Structures, 32, (617), I 995, 767-782.
17.78 Strenkowski, J.S. and Witmer, E.A., Bucking and Postbuckling Studies of Torsion-Loaded Thin-
Wall Cylindrical Shells with Cutouts, ASRL TR 171-1, Aeroelastic and Structures Research
Laboratory, Department of Aeronautics and Astronautics, MIT, Cambridge, Mass., Oct. 1973;
also AMMRC CTR 73-39, U.S. Army. Watertown, Mass.
17.79 Baker, W.E., Bennell, J.G. and Babcock, C.D., Experimental Buckling Investigation of Ring-
Stiffened Cylindrical Shells under Unsymmetrical Axial Loads, ASM£ Journal of Pressure
Vessel Technology, 105, Nov. 1983, 342-346.
17.80 Davis, R.C. and Carder, F., Buckling Tests of a 10-Foot Diameter Stiffened Cylinder wnth
Rectangular Cutouts, NASA TM 88996, Feb. 1987.
17.81 Jullien, J.P., Effects of Cut-Outs and Openings in Shells, Final Report, PartE of Enhancement
of ECCS Design Recommendations and Development of Eurocode 3, Parts Related to Shell
Buckling, INSA Lyon, URGC-Structures. March 1996.
17.82 Knight, N.F., Jr. And Starnes, J.H., Jr., Postbuckling Behavior of Axially Compressed Graphitte-
Epoxy Cylindrical Panels with Circular Holes, ASM£ Journal of Pressure Vessel Technology,
107, Nov. 1985, 394-402.
17.83 Janisse, T.C. and Pala:wtto, A.N., Collapse Analysis of Cylindrical Composite Panels with
Cutouts, AIM Journal of Aircraft, 21 , (9), Sept. 1984, 731-733.
17.84 Lee, C.E. and Palazolto. A.N., Collapse Analysis of Composite Cylindrical Panels with Small
Cutouts, Composite Structures. 4, 1985. 2 17-229.
17.85 Horban, B. and Palazotto, A.N., Expe1imental Buckling of Cylindrical Composite Panels wuth
Eccentrically Located Circular Delaminations, AIAA Joumal of Spacecraft and Rockets, 24, (4),
1987, 349-352.
17.86 Palazouo, A.N. and Tisler, T.W., Considerations of Cutouts in Composite Cyli ndrical Panels,
Computers and Sm.tclllres. 29, (6), 1988, 1101-1110.
17.87 Palazouo, A.N., An Experimental Study of a Curved Composite Panel with a Cutout, in: Com-
posile Materials: Tes£ing and Design (Eighth Conference), ASTM STP 972, J.D. Whitcomb,
ed., ASTM, Philadelphia, 1988, 191-202.
17.88 Hu, H.-T. and Wang, S.S., Optimization for Buckling Resistance of Fiber-Composite Laminate
Shel.ls with and without Cutouts, in: Collection of Papers, 31st AlAA/ ASME/ ASCE/ AHS/
ASC Structures, Structural Dynamics and Materials Conference, Long Beach, Calif., April 1990,
1300-1312, AIAA Paper 90-1069-CP.
17.89 Hilllburger, M.W., Starnes, J.H., Jr. and Waas, A.M., The Response of Composite CyLindrical
Shells with Cutouts andJ Subjected to Internal Pressure and Axial Compression Loads, in: Col-
lection of Papers, 39th AIAA/ ASME/ ASCE/ AHS/ ASC Structures, Structural Dynamics and
Material Conference, Long Deach, Calif., April 20-2.3, 1998, 576-584, AIAA Paper 98-1768.
17.90 Hillburger, M.W., Starnes, J.H .. Jr. and Waas, A.M., A Numerical and Experimental Study of
the Response of Selected Compression-Loaded Composite Shells with Cutouts, in: Collection
of Papers, 39th ATAA/ ASME/ ASCE/ AHS/ ASC Structures. Structural Dynamics and Material
Conference, Long Beach, Calif., April 20-23, 1998, 2338- 2351, AIAA Paper 98-1988.
17.9 1 Hillburger. M.W., Numerical and Experimental Study of the Compression Response of Com-
posite Cylindrical Shells with Cutouts. Ph.D. dissertation, University of Michigan. Ann Arbor,
Mich., 1998.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1~7

17.92 Taby, J., Moan, T. and Rashed, S.M.H., Theoretical and Experimental Study of the Behaviour
of Damaged Tubular Members in Offshore Su·uctures, Nonvegian Maritime Research, 9, (2),
198 1, 26-33.
17.93 Smith, C.S., Assessment of Damage in Offshore Steel Platforms, in: Proceedings of Marine
and Offshore Safety Conference, University of Glasgow, Sept. 1983, 279-293.
17.94 Ellinas, C.P. and Valsgard, S., Collisions and Damage of Offshore Structures: A State-of-the-
Art, Journal of Energy Resources Technology, 107, 1985, 297-3 14.
17.95 Ronalds, B.F., Mechanics of Dented Orthogonally Stiffened Cylinders, Ph.D. thesis, Imperial
College, University of London, 1985.
17.96 Ronalds, B.F. and Dowling, P.J., Damage of Onhogonally Stiffened Sluells, in: Behaviour of
Offshore Structures, Proceedings of Boss 85, Delft, July 1985, J.A. Battj es, ed., Elsevier, Am-
sterdam. 1985, 4 19-428.
I7.97 Harding, J.E. and Dowling, P.J., Recent Research on the Behaviour of Cylindrical Shells Used
in Offshore Structures, in: Steel Structures, Proceedings, International Conference on Steel
Structures, .Budva, Yugoslavia, Sepl. 1986, M. Pavlovic, ed., Elsevier Applied Science, London
1986, 317-338.
17.98 Ron.alds, B.F. and Dowling, P.J., Finite Deformations of Stringer Stiffened Plates and Shells
Under Knife Edge Loading. in: Proceedings of Fifth International Symposium on Offshore
Mechanics and Arctic Engineering (OMAE), Tokyo, vol. 3, 1986, ASME, New York, 323-331.
17.99 Walker, A.C. and Kwok, M.K., Process of Damage in Thin-Walled Cylindrical Shells, in: Ad-
vances in Marine Structures, C. S. Smith and J.D. Clarke, eds. , Elsevier Applied Science,
London, 1987, 111- 135.
17. 100 Ronalds, B.F. and Dowling, P.J .. Buckling of Intact and Damaged Offshore Shell Structures,
in: Advanced Marine Struc/1/res, C.S. Smith and J.D. Clarke, eds.. .Elsevier Applied Science,
London, 1987, 201-218.
17. 101 Ronalds . .B.F. and Dowling, P.J., A Denting Mechanism for Orthogonally Stiffened Cylinders,
International Journal of Mechanical Sciences, 29, (10/ II ), 1987. 743-759.
17.102 Ronalds, B.F. and Dowling, P.J., Residual Compressive Strength of Damaged Orthogonally
Stiffened Cylinders, in: Stability of Plate and Shell Structures, P. Dubas and D. Vandepitte,
eds., Proceedings of an International Colloquium, April 6-8 1987, Ghent, 1987. 503-512.
17.103 .Ronalds. B.F. and Dowling, P.J., Stiffening of Steel Cylindrical Shells against Accidental Lateral
Impact, Proceedings, Institution of Civil Engineers, Part 2, 83, December 1987, 799-814.
17.104 Kwok, M.K. , McCall, S. and Walker, A.C., The Behaviour of Damaged Cylindrical Shells
Subjected to Static andl Dynamic Loading, in: P1vceedings of Conference on Applied Solid
Mechanics- 2, A.S. Tooth and J. Spence, eds., Strathclyde University, Elsevier Applied Science,
London, 1987, 179-209.
17.105 Walker, A.C., McCall, S. and Thorpe, T.W., Strength of Damaged Ring and Orthogonally
Stiffened Shells- Par! 1: Plain Ring Stiffened Shells, Thin-Walled Strucwres, 5, 1987, 425-
453.
17.106 Walker, A.C., McCall, S. and Thorpe, T.W., Strength of Damaged Ring and Orthogonallly
Stiffened Shells-Part II: T-Ring and Orthogonally Stiffened Shells, Thin-Walled Structures, 6,
1988, 19-50.
17.1 07 Dowling, P.J. and Ronald s, B.F., On the Behaviour of Damaged and Intact Stiffened Cylindrical
Shells, .in: Proceedings of 1987 SSRC Annual Technical Session, Houston, Tex .. SSRC, Fritz
Engineering Laboratory, Lehigh University, Bethlehem, Pa., 1987, 265-274.
17.108 Ronalds, B.F. and Dowling, P.J., Compressive Strength o f Stiffened Cylindrical Shells wi.th
Large Imperfections, in: Buckling of Structures, Theo ry and Experiment, The Josef Singer An-
niver sary Volume, l. Elishakoff et al., eds., Elsevier. Amsterdam, 1988, 3 13-334.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

17.109 Dowling, P.J.• Strength and Reliabi.lity of Stringer-Stiffened Cylinders in Offshore Stmctures.
in: Buckling of Shell Structures, on Land, in the Sea, and in the Air, J.F. Jullien, ed., Elsevier
Applied Science, London, 1991. 242- 250.
17.I I0 Lancaster. E.R. and Palmer, S.C.. Model Testing of Mechanically Damaged Pipes Containing
Denls and Gouges, in: ASME PVP-vol. 235, ASME, New York, 1992, 143-148.
I7.11 1 Park, T.D. and Kyriakides, S., On the Collapse of Dented Cylinders under External Pressure,
International Journal of Mechanical Sciences. 38, (5) , 1996, 557-578.
17. 112 Demars, K.R., Nacci. V. A. and Wang, W.D., Pipeline Failure: a Need for Improved Analyses
and Site Surveys, in: Proceedings of Offshore Technology Conference, vol. 4, OTC 2966, 1977,
63- 70.

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1468 Influence of Holes, Cutouts and Damaged Structures

17. 113 Eiber, R.J., Causes of Pjpeline Failures Probed, Oil Gas Journal, 77, 1979, 80- 88.
17.114 Strating, J., A Survey of Pipelines in the North Sea, Incidents During Installation, Testing and
Operation, in: Proceedings of Offshore Technology Conference, vol. 3, OTC 4069, 1981, 25-
32.
17.115 KyrLakides, S., Babcock, C.D. and Elyada. D., lnitiation of Propagating Buckles from Local
Pipeline Damages, ASME Journal of Energy Resources Technology, 106, 1984, 79-87.
17.116 Estefen, S.F., Netto, T.A. and Alves, T M .. Residual Strength of Damaged Offshore Pipeli nes,
in: Proceedings, ASME Offshore Mechanics and Arctic Engineering Conference, Calgary. Alta.,
Canada, vol. 5, June 1992, 233-238.
17.1 17 Hambly, E.T. and Calladine, C.R. , Buckling Experiments on Damaged Cylindrical Shells, fn-
ternational Journal of Solids and Structures, 33, {24), 1996, 3539- 3548.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

17.ll8 Jones, R., Pa ul , J. and Broughton, W.. On the Effects of Delamination D amage in Fibre Com-
posite Laminates, Australian Department of Defence, Aeronautical Research Laboratories
(ARL), Melbourne, Victoria, Structures Report 403/ Material Report 116, June 1984.
17.1 19 Bass, M ., Gottesman, T. and Fingerhut, U., Criticality of Delaminations in Composite Mate rials
Structures, in: Collectio n of Papers of 28th Israel Annual Confere nce on Aviation and As tro-
nautics, Tel Aviv/Haifa, Israel, Feb. 19- 20, 1986, 186-190.
17.120 Simitses, G.J., Sallam, S. and Yin, W.-L., Effect of Delamination of Axially Loaded Ho•no-
geneous Laminated Plates, AIAA Journal, 23, (9), Sept. 1985, 1437-1444.
'17.12 1 Horban, B. and Palazono, A., Experimental Buckling of Cylindrical Composite Panels with
Eccentrically Located Circular Delaminations, AIAA l ou mal of Spacecraft and Rockets, 24, (4),
July-August 1987. 349- 352.
17. 122 Wilkins, D.J., Eismann, J.R., Camin, R.A., Margolis, W.S. and Be nson, R.A., Characterizing
Delamination Growth in Graphite-Epoxy, in: Danwge in Composite Materials, ASTM STP 775,
K.L. Reifsnider, ed., ASTM, Philadelphia, 1982, 168-183.
17.123 Simitses, G.J., Sal! am, S. and Yin, W.-L., Effects of Delamination o n Axia.lly-Loaded Laminated
Plates, in: Proceedings of 25th AIAA / t\SM£1ASCE/AHS Structures, Structural Dynamics and
Materials Conference, Part I , Palm Springs, Calif.. May I 984, 351-360.
17.124 Yin, W.-L., Sallam, S. and Simitses, G.J., Ultimate Axial Load Capacity of a Delaminated Plate,
in: Proceedings uf25th AIAAIASMEIASCEIAHS St rucwres, Structural Dynamics ami Mate-
rials Conference, Pan l , Palm Springs, Calif., May 1984. 159-166.
17. 125 Sallam, S. and Simitses, G.J., Delamination Buckling and Growth of Flat, Cross-Ply Laminates,
Composite Structures, 4 , 1985, 361-381.
17.126 Sallam, S. and Simitses, G .J., Delamination Buckling of Cylindrical Shells under Axial Corn-
pression, Composite Structures, 1. 1987, 83- 101.
17.127 Simitses, G.J., Buckling of Pressure-Loaded, Delaminated, Cylindrical Shells and Panels, Key
Engineering Materials, vols. 121-122, Trans Tech Otikon-Zurich, Switzerland, 1996, 407- 426.
17_1 28 Chai, H., Knauss, W.G. and Babcock, C.D., Observations of Damage Growth in Compressively
Loaded Laminates, Experimental Mechanics, 23, (3), Sept. 1983, 329-337.
I 7. I 29 Grady, J.E. and DePaola, K.J., Measurement of Impact-Induced Delaminat ion Buckling in Com-
posite Laminates, in: Proceedings, 1987 SEM (Society for Experimental Mechanics) Fall Con-
ference, Dynamic Failure, Savannah, Ga., Oct. 25- 28, 1987, 160- 168.
17.130 Grady, J.E., Chamis, C.C. and Aiello, R.A., Dynamic Delamination Buckling in Composite
Laminates under Impact Loading: Computational Simulation, NASA TM 100192, 1987.
17.13 1 Viot , A., Buckling of Delaminations in ARALL Laminates, TU Delft, Faculty of Aerospace
Engineering, Report LR 535, October 1987.
17.132 Horban, B., Palazotto, A. and Maddux, G .. The Use of Stereo X-Ray and Deply Techniques
for Evaluating Instability of Composite Cylindrical Panels with Delaminations, in: Proceedings
of SEM (Society for Experimental Mechanics) 1987 Spring Conference on Experimental Me-
chanics, Houston, Tex., June 14-19, 1987, pp. 341-352.
17.133 Kardomateas, G.A., Large Deformation Effects in the Postbuckling BeJ1avior of Composites
with Thi1\ Delamirtations. AJAA Jourmll, 27, (5}, May 1989, 624- 631.
17.134 Wang, J.T. and Whitcomb, J.D., The Effects of Base Laminate Damage On Instability-Related
Delamination Growth, in: Proceedings of 31st AIAA I ASM£1 ASCE IAHSI ASC Structures, Struc-
wral Dynamics and Material Confm,nce, Long Beach, Calif., April 2-4, 1990, Part 2. pp.
l20l-l207.
17.1 35 Krishnakumar, S. and Tennyson, R.C., Effect of Impact De!aminatio ns on Axial Buckling of
Composite Cylinders, in: Buckling of Shell Structures. on Land. in the Sea, and in the Air. J.F.
Jullien, ed., Elsevier Applied Science, London, 1991, 61-70.

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1469

17.136 Poon, C., Bellinger, N.C., Xiong, Y. and Gould, R.W., Edge Delaminations of Composite Lam-
inates, in: AGARD Debvnding I Delamination of Composites. Patras, Greece, Dec. 1992, 12-1-
12-12.
17.137 Guedra-DeGeorges, D., Maison, S., Trallero. D. and Petitniot, J.L., Buckling and Post-Buckling
Behavior of a Delamination in a Carbon-Epoxy Laminated Structure: Experiments and Mod-
elling. in: AGARD DebondingiDelamination of Composites, Patras, Greece, Dec. 1992,7-1-7-
11.
17.138 Tennyson, R.C. and Krishnakumar, S., Delamination Damage and Its Effect on Buckling of
Laminated Cylindrical Shells, in: AGARD Debonding I Delamination of Composites, Patras,

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Greece. Dec. 1992, 9-1-9-6.
17.139 Somers, M., Weller T. and Abramovich, H., Buckling and Postbuckling Behavior of Delami-
nated Sandwich Beams, Composite Structures. 21, 1992, 211- 232.
17.140 Labonte, S. and Wiggenraad, J.F.M., A Damage Tolerance Study Conducted with Structure
Relevant Specimens, National Aerospace Laboratory, The Netherlands, NLR Report TP
93067U, Feb. 1993.
17.141 Khamseh, A.R. and Waas, A.M., Compression Failure Mechanisms of Uni -Ply Composite Plates
with a Circular Cutout, in: Collection of Papers, 33d AIAA/ ASME/ ASCEI AHS/ ASC Struc-
tures, Structural Dynamics and Materials Conference, Dallas, Tex., April 13-15, 1992, Part I,
163- 173, AIAA Paper 92-2276-CP.
17.1 42 Kardomateas, G.A., A Solution for the Initial Postbuckling and Growth Behavior of Intemal
Delaminations, in: Collection of Papers, 34th AIAA/ ASMEI ASCE/ AHS/ ASC Structures,
Structural Dynamics and Materials Conference, La Jolla, Calif., April 19- 22, 1993, Part 5, 13-
39, AIAA Paper 93-1613.
17.143 Kardomateas, G.A., Pelegri, A.A. and Malik, B., Growth of lntemal Delaminations Under
Cyclic Compression in Composite Plates, in: Failure Mechanics in Advanced Polymeric Com-
posites, Proceedings of the 1994 lntemational Mechanical Engineering Congress and Exposi-
tion, Chicago, November 6-11, 1994, 13-29.
17.144 Klein, H., Geier, B., Goetting, H.C., Hilger, W., Pabsch, A. and Zimmermann, R., Budding
Tests with Curved, Stiffened CFRP Panels Damaged by Impact, in: Proceedings, ESA Confer-
ence· 011 Spacecraft Structures, Materials and Mechanical Testing, Noordwijk, The Netherlands,
March 27-29, 1996, ESA SP-386, June 1996, 787-793.
17. 145 Khamseh, A.R. and Waas, A.M., Failure Mechanisms of Composite Plates with a Circular Hole
under Remote Biaxial Planar Compressive Loads, ASME Journal of Engineering Materials and
Technology, 119, Jan. 1997, 56-64.

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
18
Buckling under Dynamic
Loads and Special Problems
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

18.1 Dynamic Buckling Phenomena


18.1.1 Background
The demand to design static buckling-resistant structures, particularly high-strength alloy
structures, has been recognized for over a century. On the other hand, the need to design
structures that have to withstand time-dependent dynamic loads, sometimes quite severe, and
thus may be susceptible to dynamic buckling (also referred to in the literature as dynamic
stability or dynamic instability), is relatively new. Introduction of the subject of dynamic
buckling can be traced to the investigations of Koning and Taub in the early thirties (see
[18.1]). However, buckling under dynamic loading has received serious attention only since
World War II, following the appearance of advanced technologies of large-scale, high-speed,
weight-efficient military and civilian vehicle structures, in which dynamic buckling could
have played a major role.
More recently, the problem of dynamic buckling was further emphasized with the intro-
duction of faster supersonic aircraft, ballistic missiles and launch and re-entry vehicles. The
growing demands for safety in transportation and industry have also had a strong impact on
the increasing interest in dynamic buckling.
Many forms of dynamic buckling of structures were reviewed in the proceedings of [18.2].
A few excellent monographs treating the various aspects and problems of dynamic instability,
published by Bolotin in the mid-sixties [18.3], Lindberg and Florence in the late eighties
[18.4] and Simitses in the beginning of the nineties [18.5] may assist the reader in becoming
acquainted with the topic. Detailed reviews on dynamic buckling were also presented in the
eighties by Ari-Gur, Weller and Singer (see [18.6]) and Singer and Ari-Gur (see [18.7]).
Though many classes of problems and many physical phenomena are encompassed by the
term dynamic stability (see [18.5]), the term dynamic buckling has in particular been assigned
in the literature to two essentially different phenomena (see [18.4] and [18.7]). One is as-
sociated with the response of structures to the action of oscillating loads (see Figure 18.la)
i.e. vibration buckling, where the transverse vibration becomes unacceptably large at critical
combinations of load amplitude, load frequency and structure damping. It can be shown for
the column of Figure 18.1 a that for sufficiently small values of the axial force, vibration
buckling (resonance) will occur when the loading frequency equals twice the natural bending
frequency of the column. Because of the similarity to vibration resonance this column be-
havior is called vibration buckling. Note that in ordinary forced vibration resonance, where
the pulsating load acts in the direction of motion and excites the motion directly as an applied
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
Buckling Experiments: Experimental Methods in Buckling of Thin-Walled
No reproduction or networking permitted without license from IHS
Structures: Shells, Built-Up Structures, Composites
Not for Resale, 02/13/2019 01:32:58 MST
and Additional Topics – Volume 2. J. Singer, J. Arbocz and T. Weller Copyright © 2002 John Wiley & Sons, Inc.
1472 Buckling under Dynamic Le~ads and Special Problems

~,
1\1\1\1\1\,,
v~~\Tv v
(a) (b)

Figure 18.1 Column subjected to dynamic load: (a) column supporting oscillating load, (b) a long
column subjected to a suddenly applied axial load (from H.E. Lindberg and A.L. Florence,
Dynamic Pulse Buckling, Figure 1.1, p. 2, copyright 1987, with kind permission from
Kluwer Academic Publishers)

force in the equation of motion, resonance takes place when the lateral loading frequency
coincides with the natural bending frequency of vibration. In vibration buckling, however,
the bending moment resulting from application of the axial pulsating force introduces the
force as a parameter multiplying the displacement in the equation of motion. Therefore,
vibration buckling is often more precisely termed parametric resonance. This phenomenon
has been extensively studied (see for example Bolotin's treatise in [18.3], Appendix A of
[18.5] and Evan-Iwanowski's survey in [18.8]) and is not treated in the present chapter or
elsewhere in this book.
The second phenomenon relates to the behavior of structures subjected to pulse loads. It
represents the loss of stability or the deformation of a structure to unbounded amplitudes as
the result of a transient response to an applied pulse load, i.e. dynamic buckling under impact
loads. A simple example of this phenomenon is represented by the long column subjected
to a suddenly applied axial load many times greater than its static Euler load (see Figure
18.1 b). In this case the column can survive a large axial load before reaching its critical
condition for buckling, i.e. the occurrence of unacceptably large deformations or stresses,
provided the load duration is short enough. It is apparent from Figure 18.lb that under intense
short-duration loading a very high-order deformation mode is experienced at buckling.
The conclusion that the permissible load intensity is load duration dependent reveals that
a special feature of pulse buckling is that the load determines the mode of buckling. Con-
sequently, whereas in static buckling analysis the buckling mode is known and the maximum
safe load is determined, in pulse buckling the load amplitude is prescribed and dictates the
buckling modes, thus determining the maximum safe duration of its application.
The phenomenon of dynamic pulse buckling was investigated by Koning and Taub in the
early thirties (see [18.1]), and since then has been studied experimentally and theoretically
by many researchers (see the detailed review of [18.5] and [18.7]). However, only within
the last three decades a basic understanding of buckling under impulse loads has been de-
veloped. The introduction of high-speed electronics and photographic instrumentation has
facilitated the experiments that made this development possible.
In the last decades dynamic buckling with emphasis on crashworthiness (see for example
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

[18.9]-[18.11]), energy absorption and nuclear containment structures (see [18.12] and
[18.13]), has been investigated. In these studies the emphasis has focused on the plastic
response (see for example the surveys by Jones in [18.14]-[18.16], and the analysis pertinent
to dynamic plastic buckling in Chapters 2.3, 3.4, 4.1, 4.2, 5.2 and 6.1-6.7 of [18.4]).
Despite the numerous investigations on dynamic buckling hardly any suitable experimen-
tally verified criteria for the determination of dynamic buckling loads or for presentation of
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Dynamic Buckling Phenomena 1473

design information for prediction of the dynamic buckling load of a given structure and
loading have appeared in the open literature. The need for such criteria and such a design
database motivated the extensive dynamic buckling tests carried out by Ari-Gur et al. at the
Aerospace Structures Laboratory, Technion, Israel, in the late seventies and early eighties
(see [18.17]-[18.21]).

18.1.2 Dynamic Buckling Criteria


For dynamic buckling of a perfect structure due to rapid pulse loading, there is no exact
counterpart of the classical static characteristic bifurcation load. Definition of dynamic buck-
ling is possible only in the presence of small lateral geometric imperfections in the structure
under consideration. Note that this requirement has already been recognized by Koning and
Taub in their early studies of the subject (see [18.1]). Instability in this case stems from the
growth of these imperfections. When the displacement profile growth is assessed with time
for different dynamic loads, buckling occurs when the dynamic load reaches a critical value
associated with a maximum acceptable deformation strain or stress, the magnitudes of which
are defined quite arbitrarily.
As mentioned above, the difference between the static and dynamic buckling cases should
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

be noted. In the classical static analysis we find the critical load at which a possible distur-
bance will result in an unbounded lateral deflection. In the dynamic case the disturbance has
to be specified a priori as initial conditions. Therefore, in a perfectly straight and symmetric
column, where the lateral initial conditions are all zero, a dynamic axial loading cannot
initiate an unbounded lateral motion. The buckling concept of the imperfection method,
which for static loading characterizes the buckling load as that for which the static displace-
ments of an imperfect system become excessive or unbounded (see for example of Ziegler,
[ 18.22]), therefore has to be resorted to in the case of dynamic loading. Simitses [ 18.5]
indicates that "for this phenomenon to happen, the configuration must possess two or more
static equilibrium positions, and 'escaping' motion occurs by having trajectories that can pass
near an unstable static equilibrium point." Consequently, Simitses indicates that the meth-
odologies developed by the various investigators to treat this phenomenon applied to struc-
tural configurations that exhibit snap-through buckling (violent) when loaded quasi-statically.
Dynamic buckling consideration, which consist of assessments of the response of imper-
fect structures, therefore introduces a certain arbitrariness, and there appear to be no perfect
criterion as yet for dynamic buckling and no general guidelines for design against dynamic
buckling, though it has been the subject of many studies. With the aid of simple imperfection-
sensitive models, which are modified versions of that employed by Killman, Dunn and Tsien
in [7.7] (see Figure 18.2), Budiansky and Hutchinson [18.23]-[18.25] proposed some the-
oretical criteria and estimates for dynamic buckling loads (see Figure 18.3). These criteria
suggest that as in static buckling, dynamic instability occurs when a very small increment
in the applied load given by
q(x, t) = A q0 (x, t) (t 2: 0) (18.1)
where q(x, t) represents an ensemble of loading histories generated by varying A in Eq.

Figure 18.2 Budiansky and Hutchinson's modified version of Karman, Dunn and Tsien's simple im-
perfection-sensitive model (from [18.24])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1474 Buckling under Dynamic Loads and Special Problems

Figure 18.3 Budiansky-Hutchinson plot for determination of the


dynamic buckling parameter (from [18.23])

(18.1), q0 (x, x
t) is a particular function of position and time t, and A is a parameter, results
in a large increase in the response strain, deflection, etc. of the structure. This event defines

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
the critical value of A, or the maximum load for which a bounded response exists. For
practical application of the Budiansky-Hutchinson criterion, the maximum response measure
of a structure Rmax versus a load parameter A is plotted, and the critical load parameter AJJ
is defined as the middle, more or less, of the narrow range in A over which Rmax rises steeply
(see [18.6] and [18.7]). One should note that because the loading and response are time
dependent, this analogy with static buckling is not quite complete.
Budiansky and Hutchinson applied their dynamic buckling criterion to the simple model
of Figure 18.2. Some of their results are depicted in Figures 18.4 and 18.5. Indeed, it is
apparent from Figure 18.4 that under relatively short-duration loadings the dynamic buckling
load parameter A0 is significantly larger than the critical static buckling load parameter A5 .
The effect of loading duration is more pronounced in Figure 18.5 (note that in this figure
T0 = 21TI w represents the vibration period corresponding to the natural frequency of the
unloaded structure w = (kl M) 112 and Ac is the classical static buckling load parameter).
Budiansky and Hutchinson [18.23] applied their criterion to a long cylindrical shell under
axial load. This application was further discussed in [18.25].
Budiansky and Hutchinson's proposed definition of the dynamic buckling load as the
maximum load for which a bounded response exists was found to be a possible starting
point in the search for a criterion that also lent itself to convenient experimental interpretation
(see the Technion Aerospace Structures Laboratory tests [18.17]-[18.21]). Hoff's definition
of dynamic stability (see [18.26]) "a structure is in a stable state if admissible finite distur-
bances of its initial state of static or dynamic equilibrium are followed by displacements
whose magnitude remains within allowable bounds during the required lifetime of the struc-
ture," may be another beginning. A somewhat different approach was taken by Lindberg and

Figure 18.4 Dynamic buckling amplification versus load duration under rectangular loading of the
modified Budiansky-Hutchinson imperfect models of Figure 18.2 (from [18.25])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Dynamic Buckling Phenomena 1475

0
o~---.~2----~~--~-~~.6~--~.8~--_J1.0
~,

Figure 18.5 Effect of load duration on the dynamic buckling under rectangular loading of the cubic
modified Budiansky-Hutchinson imperfect model of Figure 18.2 (from [18.24])

Florence [18.4]. They specified a load amplitude and sought its response. Knowing how the
buckling deformation grows with time, they then determined the maximum duration for
which the given load could be safely applied. Initiation of noticeable yielding is still another
criterion suggested by some investigators as suitable for dynamic buckling (for example
Mayman and Libai [18.27]).
Energy-based approaches provide an alternative for definition of dynamic buckling. In the
early sixties Humphreys and Bodner proposed an energy-based approach for determining the
minimum impulse necessary for dynamic snap-buckling of spherical caps and circular cylin-
drical panels. Their approach circumvents the solution of the nonlinear dynamical equations
for the response of the shell (see [18.28]). They stated that "because shallow shells dem-
onstrate snap buckling, the dynamic conditions for which snap buckling would occur are of
considerable interest." In contrast to the tedious straightforward approach of solving the
governing nonlinear dynamical equations for the response of the shell, and since the sole
interest was in determination of the impulse conditions or the energy input required for snap-
buckling to occur, Humphreys and Bodner indicated that the problem could be approached
from a general energy viewpoint, in which only static equilibrium states had to be examined.
In this case snap-buckling would occur when the energy input was sufficient for the shell to
pass through an unstable equilibrium position. Hence, for impulsive loading conditions the
lowest strain energy associated with an unstable equilibrium state for an unloaded structure
had to be found. This established the energy barrier, or the minimum kinetic energy input
required for snap-buckling to take place. Good correlation was reported between test results
and the analytical predictions of [18.28].
Note that the Budiansky-Hutchinson criteria [18.23]-[18.25]) also determine simplified
methods of finding the critical loading conditions without resorting to a direct solution of
the time-dependent nonlinear partial differential equations. In their methods, however, they
determine the relation between the static and dynamic buckling load.
Energy approaches, the total energy-phase plane and the total potential energy, were em-
phasized and discussed in detail by Simitses [ 18.5]. Furthermore, in the introductory part of
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

this book Simitses classified the totality of concepts and methodologies used by the various
investigators in estimating critical conditions for suddenly loaded elastic systems in accord-
ance with the following three groups:
1. Equations of motion approach (Budiansky-Hutchinson type), where the equations of
motion are solved for various values of the load parameter to obtain the response of the
system. When the motion of the system changes from small-amplitude oscillatory to
large-amplitude oscillatory or becomes associated with distinctly and finitely removed
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1476 Buckling under Dynamic Loads and Special Problems

positions from the undisturbed ones, then the corresponding load parameter (lowest
value) is called dynamically critical. Being computer oriented, this solution approach
has received wide acceptance. An important advantage of this approach is its capability
to determine the dynamic critical load. Its disadvantage is the extremely large amounts
of computer time required to solve the equations of motion at different levels of the
applied load.
2. Total energy-phase plane approach, where the critical conditions are related to the char-
acteristics of the system phase plane. Here the emphasis is on establishing lower bounds
(minimum possible critical loads MPCL's) and upper bounds conditions (minimum guar-
anteed critical loads MGCL's) for the critical dynamic loads.
3. Total potential energy approach, where critical conditions are related to the character-
istics of the systems total potential. In employing this approach, which is applicable to
conservative systems only, lower and upper bounds of critical conditions are established,
too. For design purposes one should estimate the lower bound obtained by approaches
2 and 3.
Simitses has shown that this approach may be used for any type of structure, and not only
for structures that exhibit snap-buckling.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

18.1.3 Theoretical Considerations in Dynamic Pulse Buckling


Lindberg and Florence [ 18.4] present the theory of dynamic pulse buckling (equations of
motion approach), elastic and plastic, of bars, rings, plates and shell elements. Some of the
material presented in this excellent monograph has already been summarized by Goodier
[18.29]. Lindberg and Florence emphasize that most of the features unique to dynamic pulse-
buckling are demonstrated by phenomena and responses associated with the elastic and
plastic dynamic buckling of bars. These include the many high-order modes, the buckling
modes that are excited, the amplitude of which grow exponentially with time as long as the
load is maintained, and the shape into which the bar deforms, which depends on the bar
imperfections, discrete or random, that initiate growth. It was also indicated by Lindberg
and Florence that it is convenient to examine in bar elements the change in buckling patterns
and critical load duration as the load amplitude is varied. For this purpose, experiments were
performed on bars with elastic impact stresses, induced by a nominally rectangular pulse
shape, ranging from low stresses to almost 80 percent of the yield stress. The pulse duration
was increased for each stress value until permanent buckles were experienced. This resulted
in a curve of critical pulse duration versus axial stress, which was compared with theoretical
curves for several assumed imperfections amplitudes. As shown in [18.4], similar critical
curves of peak pressure versus impulse were obtained for cylindrical shells under radial
pressure impulses.
Impact buckling of bars poses a more complex problem than buckling of shells under
radial pressure. In the case of the bar, the load is communicated by axial stress waves that
move up and down the bar. It is shown, however, in [18.4], that the amount of lateral buckling
motion that takes place during the time needed to traverse a buckle wavelength is small
compared with the total buckling motion. Thus, for practical purposes, the problem of buckle
motion as the wave passes can be neglected and it can be assumed that the bar is subjected
to a uniformly increasing load along its length in series of steps that add to the load with
each arrival of the wave at the ends of the bar. Furthermore, if the bar were very long, it
would buckle before any signal is received from the free end; thus, no bar length and hence
no physical Euler load exist. Nevertheless, it is shown in [18.4] that it is common practice
to formulate the theoretical problem as though the bar were a column supported at both
ends.
A detailed derivation of the theory of dynamic elastic and plastic buckling of a simply
supported bar was presented in [18.4] and [18.29]. Following these references some of the
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Dynamic Buckling Phenomena 1477

results obtained there will now be discussed to provide some insight into dynamic buckling
and present some of its pertinent features.

a. Elastic Buckling of a Simply Supported Bar

The dynamic problem is associated with loading a bar by an axial time-dependent stress, the
amplitude of which is much greater than its static Euler buckling stress. However, the solution
of this problem in [18.4] and [18.29] adopts the yield stress criterion for determination of
dynamic buckling, i.e. the applied stress is less than the yield stress of the bar material.
Under these circumstances buckling is elastic before yielding occurs at the crests of the
buckles.
The equation of motion associated with the bar shown in Figure 18.6 is given by:

a4 y a2 cPy
EI - 4 + p -2 (y + Yo) + pA -2 = 0 (18.2)
ax ax at
The solution of Eq. (18.2) subject to simple supports boundary conditions

at x = 0; L (18.3)

can be expressed by:

"' . wrrx
y(x, t) = ;~ q,( T)sm
1 L (18.4)

The initial geometrical imperfection can also be expressed in a similar series form:

~ . n7TX
Yo(X, t) = n~ A,( T)sm L (18.5)

where

---

Q + dQ

L- __j -- --- -
I
dx
-

(bl

Figure 18.6 Nomenclature used in the analysis of dynamic buckling of long bars in [18.4] (from H.E.
Lindberg and A.L. Florence, Dynamic Pulse Buckling, Figure 2.1, p. 14, copyright 1987,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
with kind permission from K1uwer Academic Publishers)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1478 Buckling under Dynamic Loads and Special Problems

A, =
2j'L y0 (x)sin Ln'TTX dx
L (18.6)
0

Substitution of Eqs. (18.4) and (18.5) into Eq. (18.2) and employing the relations P =
PI El; r 2 = I I A and c 2 = El p, where r is the radius of gyration of the cross-section and c
is the speed of wave propagation, yields:

(18.7)

Equation (18.7) presents one of the principal points of dynamic buckling. The coefficient
of q, in this equation dictates its solution. If (nTr!L) < k, this coefficient is negative and the
solutions of Eq. (18.7) are hyperbolic; if (nTrl L) > k, the coefficient is positive and the
solutions are trigonometric. It is therefore apparent that for large enough mode numbers n
the displacements are trigonometric and thus bounded. However, over the lower range of
mode numbers, n < (kLI Tr), and the hyperbolic solutions grow exponentially versus time,
implying that small initial imperfections can be greatly amplified. These amplified modes
are designated buckling modes. The above conditions imposed on n, further imply that.
immaterial how long the duration of load application, for any n < (kLITr) the motion di-
verges, whereas for n > (kLITr) it is bounded.
In dynamic pulse buckling problems we are interested in cases for which P > > Tr 2 EI IU.
Under such circumstances the mode numbers of the buckling modes are very high and the
wavelengths of the buckling become so short that the total length of the bar becomes rela-
tively unimportant. This implies that in experiments on dynamic buckling of bars one of its
ends has to be impacted, and because of the finite speed of axial wave propagation, buckling
occurs before any signal has been received from the opposite end. It was therefore suggested
in [18.4] that 11 k be used as a characteristic. Furthermore, normalizing the lateral deflection
of the bar with respect to the radius of gyration of the cross-section r and introducing
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

the ratio s 2 = r 2 k 2 = £ and the dimensionless variables: w = y I r; g = kx = sx I r and T =


s 2 ctlr, Eq. (18.2) becomes:
w'v + w" + w = -w;; (18.8)
with the simple supports boundary conditions given by:
sL
w = w" = 0 at ( = 0 ( = f =- (18.9)
r
the solution of Eq. (18.8) is expressed by:
% • nTrg
W( g; T) = '~' g,( T)Slll -{-
1 (18.10)

and the initial imperfections are:

(18.11)

where:

a, = e2 Jo( WoW . nTrg


Slll -{-~ dg (18.12)

Introducing the wave number YJ = nTrl ~ the equations of motion for the Fourier coefficients
g,( T) become:
g, + Y)
2
(Tl - l)g, = T7 2 a, (18.13)
and the transition from hyperbolic to trigonometric solutions occurs at YJ = 1.
For a bar initially at rest
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Dynamic Buckling Phenomena 1479

w(g, 0) = w(g, 0) = 0 (18.14)


the displacement is given by:

w(g, T) = ,~1 1 ~" TJ [ceo~~~ p,


2 T - J
1 sin n;g (18.15)

where
(18.16)
and it takes the hyperbolic diverging form for TJ < 1.
Equation (18.15) reveals that the buckling terms grow exponentially, yielding amplification
functions

G,( T) = -g,( T) = - -1 - [ COSh Pn T - 1 J (18.17)


a, 1 - TJ~ cos
7

A plot of this function is shown in Figure 18.7, where TJ is treated as a continuous function.
It is apparent from Figure 18.7 that there exists a wave number of the most amplified mode.
It was shown in [18.4] that for practical purposes this preferred mode corresponds to T/p =
1lv5_ and is associated with a wavelength equal to
xP = 8.88 riVe (18.18)
where e is the axial strain due to the compressive force P. The maximum amplification
versus time is depicted in Figure 18.8. It is observed in this figure that beyond T = 4 growth
is very rapid. These results suggest that bars under very high compression will buckle into
wave lengths near 8.88 riVe at nondimensional times, between 4-12. For better estimates
of critical buckling times, see further discussions in [18.4].

b. Elastic Buckling Under Eccentric Loads


In deriving the theory of dynamic elastic buckling under eccentric loading (see Figure 18.9)
for a bar of rectangular cross-section of height h and of initial shape

(18.19)

it was shown in Ref. 18.4 that the relation between the compressive impact stress cr, and
the time T"' at which yielding of the bar material occurs, is given by:

20

0
o 02 04 0.6 0.8 10
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

'1

Figure 18.7 Amplification versus wave-number functions indicating the wave number of the most
amplified preferred mode of buckling of a bar (from H.E. Lindberg and A.L. Florence,
Dynamic Pulse Buckling, Figure 2.3, p. 25, copyright 1987, with kind permission from
Kluwer Academic Publishers)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1480 Buckling under Dynamic Loads and Special Problems

400 , - - , - - , - - , - - , - - , - - - , . ,

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Figure 18.8 The effect of time on the maximum amplification of
a bar (from H.E. Lindberg and A.L. Florence, Dy-
namic Pulse Buckling, Figure 2.4, p. 27, copyright
1987, with kind permission from Kluwer Academic
NORMALIZED TIME, T Publishers).

(~:Yl = 1+ o.732~ [cosh(~·)- 1j (18.20)

In the development of this equation, the threshold of buckling was defined by the total stress
u 8 + uc reaching the material yield stress (note that u 8 is the bending stress induced by the
applied axial force in the inner fiber of the cross-section).
In Figure 18.10 a graph ofT" versus rrJu, obtained from Eq. (18.20), is presented for
various values of normalized eccentricity 3/ h. The figure also shows the amplification G,,,
obtained from Eq. (18.17) with 17 = 1/\/2, required to produce the first yielding for an
eccentricity of h = 0.01. It is apparent from Figure 18.10 that for small values of impact
stresses the amplification must be very large to induce yield. Under these conditions larger
buckling deformations may occur before the yield stress is reached, thus imposing an upper
limit on Tcr· It should, however, be noted that with the yield definition of dynamic buckling,
adopted in the derivation of Eq. (18.20), Tee reaches infinity as P approaches zero. As the
impact stress approaches the yield stress, the amplification needed to produce the first yield-
ing is quite small. Furthermore, since real materials do not exhibit a sharply defined yield
stress and, more important, the tangent modulus rapidly declines as the material yields, the
introduction of the elastic modulus in the buckling formulation is inappropriate. Hence, the
curves in Figure 18.10 become meaningless as uc reaches u,.
It was further shown in [18.4] that the curves of Fig. 18.10 can be interpreted physically.
The dimensionless time T corresponds to the impulse of the applied load, and therefore, from
its previous definition, this pulse is given by

AEr
Pt = - - T (18.21)
c

(here A, E, r, c T, P and t are the area of the cross-section, Young's modulus, stress propa-
gation velocity, nondimensional time, applied compressive force and duration of loading,
respectively) and the critical impulse that leads to first yield due to buckling is

p-.,~---------------6~~~p
t
Figure 18.9 Eccentrically loaded bar (from H.E. Lindberg and A.L. Florence, Dynamic Pulse Buckling,
Figure 2.5, p. 27, copyright 1987, with kind permission from Kluwer Academic Publish-
ers).
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Dynamic Buckling Phenomena 1481

Figure 18.10 Critical buckling times to first yield for bar loaded with different eccentricities (from
H.E. Lindberg and A.L. Florence, Dynamic Pulse Buckling, Figure 2.8, p. 32, copyright
1987, with kind permission from Kluwer Academic Publishers)

AEr
/cr = - - Tcr (18.22)
c
Furthermore, the applied load can be expressed by

P = A(T, = A(T, (;J (18.23)

The curves in Figure 18.10 represent, therefore, combinations of load amplitude P and load
pulse I that generate threshold buckling. Load points above the curves result in severe buck-
ling, whereas load points below the curves are not associated with permanent buckling de-
formations.
The problem of dynamic elastic buckling of bars with random imperfections was also
considered in [18.4]. Two types of imperfections were treated, imperfections proportional to
the buckle wavelength and imperfections proportional to the depth of the bar. In both cases,
results representing counterpart to Eq. (18.20), for buckling from eccentric impact, were
obtained. As in the case of the eccentric bar, these formulations yielded results similar to
those depicted in Figure 18.1 0, where T"', the threshold time of loading duration to first yield,
is plotted versus the ratio of the applied stress to the yield stress.

c. Dynamic Plastic Flow Buckling of Bars


Repeated plastic dynamic tests show reproducibility. On the basis of this observation Goodier
concluded that characteristic forms of plastic dynamic buckling overshadow accidental initial
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1482 Buckling under Dynamic L.oads and Special Problems

imperfections (see [18.29]). Some of these deformation forms could be well understood by
ignoring elastic deformations. For others, however, the final permanent deformation pattern
was set by the preceding elastic deformation phase, discussed above.
In thick (or short) bars under axial impact, in which the time required for significant
elastic buckling is significantly longer than the duration 2L/ c of the axially applied com-
pression load at the impact point, buckling will occur only when the impact stress and strain
will be well beyond the elastic range. Assuming that in such bars axial flow ceases at each
cross-section upon the arrival of the elastic unloading wave reflected from the end of the
bar, some part of it will be in sustained plastic flow during a total period of time required
for the elastic wave to travel from the impacted end to the free end and back. Furthermore,
in such bars the axial compressive strain at each cross-section is increasing during buckling,
therefore buckling is treated as a perturbation of the motion associated with the axial com-
pression. Supposing that the axial strain rate dominates the extensional strain rate introduced
by the bending motion, buckling has to be well developed before strain rate reversal occurs.
During this period the bending stiffness of the bar, being proportional to the tangent modulus
in the plastic range of the stress-strain curve, is very low and thus bending occurs relatively
easily. Once the elastic unloading wave returns from the free end of the bar, the bending
stiffness increases due to becoming partly or completely governed by the elastic modulus.
Since in tests on plastic dynamic buckling of bars no perceptible buckling is observed while
the wave front traverses a wavelength, the wave front is disregarded and the bar is assumed
to buckle under uniform pressure (see [18.4] and [18.29]).
The differential equation of motion of a bar suddenly subjected to a thrust P (exceeding
the yield value), using a simplified theory for dynamic plastic flow associated with the stress-
strain curve of Figure 18.11, is identical with Eq. (18.8) for elastic buckling. However, in
this equation the elastic modulus E is replaced by E" (see Figure 18.11). The solution of
this equation for an initially straight bar and a nearly straight one yields magnification func-
tions, the behavior of which are qualitatively similar to those obtained in the case of elastic
dynamic buckling (see Figure 18.7 and Figures 2.29a and b in [18.4]).

d. Technion Theoretical Model


Another theoretical study (which also represents an equations of motion approach) on the
dynamic response of thin bars and plates subject to pulse loading, in which the Budiansky-
Hutchinson dynamic buckling criterion, discussed in Subsection 18.1.2, was implemented to
determine the elastic dynamic buckling, was reported by Ari-Gur, Weller and Singer [18.6],
[18.7], [18.17]-[18.21] and [18.30]. The derivation of the theory for a bar struck by a mass
M0 at one end, which is guided by an axial bearing and clamped at its other end (see Figure
18.12), the details of which are provided Ln [18.18] and [18.19], represents an extension and
improvement of the analysis of Hayashi and Sano (see [18.31] and [18.32]). In contrast to
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

I
"' -
s·~
- SLOPE- Eh
~
...
a:

"'
UJ
>
v;
"'
~
~
0
u

STRAIN-E

Figure 18.11 Idealized stress-strain diagram (from H.E. Lindberg and A.L. Florence, Dynamic Pulse
Buckling, Figure 2.28, p. 62, copyright 1987, with kind permission from Kluwer Aca-
Copyright Wiley demic Publishers).
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Dynamic Buckling Phenomena 1483

,.
x,u

Figure 18.12 Ari-Gur et al.'s model of clamped column (from [18.19]).

the analysis of [18.4], which is based on Eqs. (18.2) and (18.8), in their derivations Ari-Gur
et a!., following Hayashi and Sano, included the effect of axial inertia. Consequently, their
theoretical model consists of two nonlinear coupled equations, the development of which is
presented next.
Assuming a thin-walled beam that complies with the Bernoulli-Navier hypothesis, the
equations of motion are given by:

(18.24)

where
h/2

(N,, M,) =
J _"
12
a.~ (1, z)dz

arc the force and moment per unit width, respectively, u and w are the axial and lateral
displacements, respectively, and p is the constant mass density of the column of uniform
depth h.
For a column made of a Hooke-type material, the forces and displacements are related
by:

(18.25)

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
where
h/2
(A, B, D) =
J
-h/2
(1, z, z2 )E(z)dz

and for which the strain-displacement relations are given by:

sx = u,, + 2
1 (w,~' - w0' .J

and (18.26)
K, = -(w - w 0 ),_,x
and the longitudinal wave-propagation velocity, the radius of gyration of the cross-section,
the slenderness ratio and the nondimensional displacements are c = (AI ph) 1 12 , r0 =
(DIA) 112 , ,\0 = Llr0 , u
= u/L and w = wiL, respectively, the nondimensional equations of
motion become
ff

u + ww
I ff

- ww
0
I ff

0 - (BIAL)(w- w =u 0)
Iff ··

I I I I ff I U I n
w(w w - 112w w)
I

(u w) + (312)w 2w - 0 0 0 + (BIAL) (18.27)

[u'' + w;;(w - W0 )], - (1 I .\6)(w - w0 ) , = ~ - 11 12(h!L?~,


Here ( )' and dot(·) represent derivatives with respect to the nondimensional axial coordinate
x = xiL and timet= ctiL.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1484 Buckling under Dynamic toads and Special Problems

Solution of Eqs. (18.27) has to comply with the dynamics of the striking mass and the
initial and boundary conditions.
Dynamics of the striking mass: At the initiation of contact between the mass and column,
the axial displacement of the mass is defined by
U(t = 0) = 0 (18.28)

the striking velocity is


V(t = 0) = U, 1 (t = 0) = Vo (18.29)
The mass is decelerated by the repulsive force bNJx = 0) at the surface of impact. Therefore,

V(t) = V0 + -b
Mo
lt N,(x = O)dt (18.30)

and

U(t) = I: V(t)dt (18.31)

Introducing the nondimensional terms:


-V=-v -
V0 =
V:
__Q
_ ct
and t = -
c' c L

Eq. (18.31) becomes:

V(t) = V0 + !!!_ (' eJx = O)dt (18.32)


M Jo

where m is the mass of the column.


Initial and boundary conditions: Assuming a rigid striking mass, the initial conditions are:

w(t = 0) = W0

W, 1(t = 0) = u(t = 0) = u, 1(x > 0, t = 0) = 0 (18.33)


U, 1(X = 0, t = 0) =c V0

and the boundary conditions are given by:


w = w,, = 0 at x = 0, L
u(x = L) =0 (18.34)
u(x = 0) = U when N,(x = 0) < 0, otherwise N,(x = 0) = 0
The last condition is time-dependent. It expresses the mutual interaction between the motion
of the specimen and the striking mass. Simultaneous solution is therefore required for both
bodies.
Analysis of the impact response of a thin bar using Eqs. (18.24 )-( 18.34) eventually yields
the Budiansky-Hutchinson type plots of Figure 18.3. Such plots are shown in Figure 18.13
for the maximum amplitude of the response of a bar versus the maximum amplitude of the
impulse loading, represented by the maximum amplitude of the axial strain e, that excited
the response. In this figure the higher the initial imperfection of the column w 0 , the lower
its dynamic buckling capacity. It should be noted that the location of the peak of the im-
perfection w 0 was found to have no influence on the flexure-compression behavior of the
column (see [18.6] and [18.19]). Furthermore, it was shown in these references that by
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

approximating the curves of Figure 18.13 by


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Dynamic Buckling Phenomena 1485

25 St••I,L:z.26m,ts2., ~ls-Ts.n::l
A.•H .~~246, tc:r:::: •163 ,uStra1n

20

15

Figure 18.13 Ari-Gur et al.'s nondimensional dynamic deflection-compression plots (Budiansky-


Hutchinson plots) for various magnitudes of initial geometrical imperfections (from
[18.19])

(18.35)

(where a is a constant common to all curves and f3 and sc, are constants that differ for each
w0 plot) and setting a = 0 and {3 = scr, Eq. (18.35) becomes the well-known equation of
the Southwell plot. Hence, Eq. (18.35) may be called a f?eneralized Southwell equation and
a plot of !lwls versus (!lw - aw0 s) provides scr (see Figure 18.14). Since the dynamic
Southwell plot requires a priori knowledge of w0 , a requirement that does not exist in the

Slt>t'l, L·.26m, l• 2.,6.X • -,i-. n:l W0 m<H :,OJmm

• ~ vs. aw
E
400
- ~ ..,,. b.W-a.Wot
E

300

E 200

oOL_____-L----~,L-----~----~---

f.W-ClWoE[mm)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 18.14 Dynamic Southwell plot proposed by Ari-Gur eta!. (from [18.19])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1486 Buckling under Dynamic Loads and Special Problems

case of static buckling, it was shown in [18.6] and [18.19] that when aw0 s is much smaller
than llw a plot of llele versus llw is a sufficiently close approximation of the exact dynamic
plot, which is represented by a straight line (dots in Figure 18.14 ).
Equations (18.24)-(18.34) were also employed to evaluate the effect of impulse duration
r = cT/2L (where Tis the duration of impulse) on the response of slender columns (see
Figure 18.15). It is apparent from this figure that the longer the impulse duration, the lower
the dynamic buckling capacity of the column.
The important role of the impulse duration on the dynamic buckling of imperfection-
sensitive structures was emphasized by Budiansky and Hutchinson in their studies (see
[18.23]-[18.25] and Subsection 18.1.2). In evaluating the influence of the duration of the
impulsive load in the studies of Ari-Gur et al., it was shown that an impulse can be given
by (see Figure 18.16):
1

I = [ fdt = YJFT (18.36)


.()

where F, T and YJ denote the maximum axial force, pulse duration and fraction of the impulse
within a rectangle bounded by F and T, respectively. The shape factor YJ depends on the
origin of the pulse load and varies, in practice, between YJ = 0.5 (a triangle) and YJ = 1 (a
rectangular loading history). For the load history of Figure 18.16, which was observed in
the tests of Ari-Gur et al., YJ = 0.6 (see [18.6] and [18.19]). It is apparent that the duration
of loading and pulse shape determine the dynamic character of the loading and therefore
have a decisive influence on the response of the structure.
Since for short-duration dynamic loading the impulse I is the dominating load parameter,
it was suggested in [18.6] that results obtained by one type of impulsive loading could be
readily applied to other types of dynamic loading. This can be achieved by relating the
specific impact 1/m to the dynamic compressive strains" in the equation developed in [18.6]:
( 18.37)
Due to the relation between the impulse and peak force associated with it, iso-response
curves should exist for a given structure, i.e. curves that represent the combinations of peak
force and impulse resulting in the same lateral response (see Figure 18.17).
For further reading on the theory of dynamic buckling of bar, plate, ring and shell elements
the reader may turn to [18.4] and [18.5] and the references cited in [18.7].

St••t ,L •. 20m, h~1.6mm, Wom 0 ~·.5Jmm

9 't•J dJ:a -fij.n•l


lwlma• ).•tt .189 tcrE • 276 j.J.Strain
[mm]
0 t~ ..u.d
2L

1000 2000 3000 4000 sooo 6000 7000 0000

(- t e )ma.11 [1JSirain]

Figure 18.15 Ari-Gur et al.'s prediction of the effect of impulse duration on the response of a column
(from [18.7]) --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Impact-Induced Buckling Experiments 1487

ff(t)clt=llFT
0
F - - - - - - - - ----------,

Q
<::;:_
I
w
0
0::
0
LL

Figure 18.16 A typical shape of applied compressive impulse (from


TIME-t [18.61).

18.2 Impact-Induced Buckling Experiments


18.2.1 Column Buckling
In the early fifties Gerard and Becker [18.33] studied theoretically and experimentally the
behavior of a column under short-duration impact. The objective of their investigation was
to demonstrate that under certain conditions of impact loading a column could momentarily
support dynamic compressive stresses of any magnitude. In contrast to earlier investigations,
e.g. Meier [18.34] and Hoff [18.35], where the period of lateral motion of the column was
assumed to be smaller than that required for the stress wave to propagate from the impacted
end of the column to the other end, thus suggesting the existence of constant stress through-
out the entire length of the bar, Gerard and Becker emphasized the propagation of an elastic

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
stress discontinuity in the column and consequently the instantaneous prevalence of noncon-
stant stress over the entire length of the bar. Under these conditions they adopted an approach

1
f·I0
ISO. R£SPONS£ CURY£ FOR w, 2 h
llrotPACl ON COLUMN
l• 266mm. h•l 6rnm. ;\ • S76,
m I aCilf, C I 5100 m,lf'(
Womax • 0.6mm

10 '

IMPULS.E
REALM

111

10
OYNA!oAIC
REALM

r QUASI. STATIC
1: "'lao REALM

4 6 8 10 4 s a 101

~·IOJ
me

Figure 18.17 Ari-Gur et al.'s prediction of iso-response curve (yielding a lateral deflection of w = 2h)
for impact of a column (from [18.7])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1488 Buckling under Dynamic Loads and Special Problems

aimed at determining the critical length of a column subject to a prescribed stress, rather
than finding the critical stress of a column of given length.
To verify their analysis, Gerard and Becker developed a simple and unique test technique,
which though published in the early fifties is still very attractive. In this test procedure, a
notched specimen is loaded in tension until fracture occurs. In essence this test condition
represents the impact of a long bar moving with velocity V against a massive (rigid) object
(see Figure 18.18). As shown in [18.33] and [18.4], the velocity at which the stress wave
induced by this impact travels equals:
V = ul pc ( 18.38)
Here u represents the average fracture stress at the notch of the specimen (or the magnitude
of the propagating induced stress wave in Figure 18.18).
In the tension tests, the velocity V is achieved by first applying with a tensile machine a
tensile stress u to pre-notched specimens (where notches of different depth are filed close
to the upper jaw to provide the desired stress u at fracture). After fracture, a compression
relief wave travels down the bar at velocity c, leaving the bar stress-free behind this wave
(see Figure 18.18) and traveling at a velocity given by Eq. (18.38). When the wave arrives
at the lower jaw, it reflects as a compressive wave. At the instant of this reflection Eq.
(18.38) holds and thus the compressive stress equals the tensile stress u.
Gerard and Becker assumed that tensioning of the bar eliminates its imperfection. How-
ever, they ignored the possibility that after fracture this might lead to introduction of lateral
vibrations and reintroduction of the imperfections. Furthermore, after fracture the cross-
section of the bar is not necessarily straight and perpendicular to the axis of the bar. Hence,
the stress wave front may not be straight.
The test procedure of Gerard and Becker was employed by Lindberg, who also photo-
graphed the development of the response of the bar (see [18.36] and [18.4]). From these
photographs (see Figure 2.15 of [18.4]) it appears that all the buckles remained nearly fixed
in position and merely grew in amplitude. Good correlation between the measurement of the
wavelength of the buckle with the calculated one was found.
The theoretical results presented in Subsections 18.l.3a-18.l.3c (which are based on
[18.4]) assume that impact of a bar induces a uniform thrust throughout its length. In impact,
however, the thrust is applied by a moving axial stress wave. Therefore, at each instant only
the length of the bar traversed by the wave is under compression and hence the theory is
not strictly applicable to the impact problem.
Lindberg [18.37] conducted tests to observe the possible effects of the moving wave and
validate the early exponential buckling growth predicted by the simple theory (see Subsection
18.l.3a). In these tests Lindberg employed a special experimental setup to amplify recording
of the tiny early motions of the bar, other than directly observing the buckling in an edge-
on view (see [18.36] and Figure 2.15 of [18.4]). In this setup the optical-lever method
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

together with a streak camera were used 1:0 record the wrinkles that formed near the support

Ill I II I I
-
1
v

II I I I I IiI I I 1 {a)

i-Lluli.J...:Il..~...lluluiiJ.JII...J..I l....:.l..~l....J..I...J..I.LILlLl lu_I....:.I....L!..LI


u.i ;_I

Figure 18.18
t==l-:-
Axial stress wave developed in a bar impacting a rigid wall at velocity V (from [18.4])
(b)

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Impact-Induced Buckling Experiments 1489

of a bar tested by the aforementioned Gerard and Becker testing technique (see [ 18.4] and
[ 18.37] for more details on the method).
In these tests Lindberg found that the observed buckling consisted of exponential growth,
which could be quite adequately predicted by the simple theory of Subsections 18.1.3a-
18.1.3c. In spite of the thrust being introduced by a moving wave, the simple assumption
of a uniform stress was adequate because the stress wave had propagated a long distance
along the bar before significant buckling displacements appeared.
Lindberg showed that an analytical expression could be derived for the distance the axial
stress wave had traveled, which was based on the preferred wavelength AI' = 21rv'2 and the
critical time Tcr at which the buckled form was unalterably determined. This distance could
be expressed in terms of the number N"' of preferred wavelengths LP, where

(18.39)

As indicated by Lindberg, this relation suggests that "the reasonableness of neglecting


axial wave effects depends only on the compressive strain of the axial thrust. In metals this
strain is very small within the elastic limit and, as we have observed, elastic buckling is
adequately represented by the constant thrust theory."
Lindberg [18.37] also reported that the magnification of the optical lever revealed that the
axial impact induced very high-frequency bending vibrations superimposed on the buckling
motion. It was observed that these oscillations had little effect, because of their very short
wavelength and short period compared with the buckling motion. These results also con-
firmed that effects dependent on the moving axial stress front had a negligible effect on the
buckling.
The test program of [18.37] also included experiments on pure gum rubber strips, which
could sustain large elastic compressive strains (up to about 15 percent in the tests), to provide
statistical information that would enhance the confidence in the theory of Subsections
18.1.3a-18.1.3c. These gum rubber tests strongly supported the previous conclusion, that

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
lateral motion behind the axial-stress wave front had a negligible influence on the wrinkle
deformation, and validated the constant-thrust theory. It was observed in the tests that the
main effect of the moving thrust was on its duration, which decreased with moving away
from the impacted end of the bar. Lindberg suggested that this could be accounted for by
assigning a different duration to each wrinkle. Furthermore, this conclusion could be applied
to more complicated structural elements such as cylindrical shells.
Lindberg and Florence [18.4] reported the results of tests estimating the thresholds of
pulse buckling. The Gerard and Becker test method of [18.33] was again used for running
experiments on thin 6061-T6 aluminum strips 12.7 and 6.3 mm wide and 0.31 and 0.63 mm
thick, subjected to initial nominal tensile stress (and reflected compressive stress) of 0.4 and
0. 7 times the yield stress of 290 MPa, obtained by appropriately sized fracture notches in
the strips. By varying the distance between the lower jaw and the notch, different trust
durations 2L/ c at the lower jaw were achieved. For each combination of strip width, thick-
ness and compressive stress, tests were performed at increasing lengths until plastic buckles
appeared. The dimensionless time (load duration) for each test is given by (see [18.4]):
2s,L
r= - - (18.40)
r
The test results, with open points representing tests in which no buckling was observed
and solid points corresponding to tests in which buckling took place, are shown in Figure
18.19, together with curves, corresponding to bars with different imperfections (Ace/h) pre-
dicted by the theory presented in Chapter 2 of [18.4] (here A"' is the imperfection in the
preferred mode of buckling). It is apparent from this figure that upper test points that are
associated with longer-impact duration are all solid (buckling), whereas the lower test points,
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1490 Buckling under Dynamic Loads and Special Problems

Figure 18.19 Comparison of buckling observed in Lindberg's tests on aluminum strips with predicted
critical curves for imperfections in preferred mode proportional to strip thickness (from
[18.4])

which correspond to short-impact duration, are nearly all open (no buckling). In the inter-
mediate range of durations, buckling and no-buckling points intermingle as a result of the
random nature of the imperfections. It is further observed in Figure 18.19 that the experi-
mental transition band of the intermingled points between buckling and no buckling follows
the trend of the calculated curves, with equivalent imperfections ranging from about 0.01-
0.03 times the thickness of the bar. On the basis of these observations, it was concluded in
Ref. 18.4 that the random imperfections in the tests were equivalent to single imperfections
in the preferred mode of one to three percent of the strip thickness. It was suggested that
this result could be used in calculations of critical pulse loads in a variety of dynamic pulse
buckling situations, including buckling of plates and shells. Lindberg and Florence indicated
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

that this was confirmed by tests on shells.


Extensive experimental studies on the dynamic buckling of columns were also included
in the aforementioned investigations of Ari-Gur, Weller and Singer (see [18.6], [18.7],
[18.17]-[18.21] and [18.38]). These studies aimed at defining suitable criteria for determi-
nation of dynamic buckling, identifying the dominant parameters responsible for the phe-
nomenon, evaluating their effects and finally providing designers with useful information
and charts.
In the test program of Ari-Gur et al. an impulsive axial compression of a column was
produced by direct collision with a striking mass. A special test rig (see Figure 18.20) was
designed and built in order to attain controlled speed and direction of the striking mass at
collision and repeatability of the response of the impacted column.
The test rig included a vertical tube along which a cylindrical mass could be dropped
from any desired height to strike a specimen positioned under the tube. With the aid of an
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Impact-Induced Buckling Experiments 1491

Figure 18.20 Ari-Gur et al.'s test setup for dynamic buckling of columns (from [1&.6))

accelerating spring, velocities of up to 15 m/ sec were obtained. Various strik ing masses
could be employed to achieve desired durations of impulse. Since rebound was not prevented,
the impulse transferred to the column was larger than the momentum of the striking mass.
Second impact was not prevented, but it was weaker and occurred after the maximum de-
flection was obtained.
The spec imens employed in the tests were of th in rectangul ar cross-sections with effective
slenderness ratios of about Acr1 " ' 80-410. The lower edge of the column was clamped by a
massive grip resting on the ground. The upper edge was inserted into a cyli nder made of
Akulon (trade name for type 6 Nylon). which served as a lightweight slide bearing inside
the vertical tube. The motion of the impacted edge was therefore restrai ned to the axial
di rection only and the boundary conditions approximated clamped-clamped conditions.
The response of a column in a typical impact test was measured by a pair of strain gages
located at the middle of the colunm length. The response of each gage was recorded sepa-
rately by a Biomation model 805 wavefonn high-frequency digital recorder. These records
were then ploued (see Figure 18.2 1) simu ltaneously with their sununation and subtraction
to yield records of the compression and bending responses, respectively. In a test a specimen
was impacted with increasing velocities and these responses were plotted for each impact.
Resu lts of direct impact tests conducted on 15 steel specimens ("Wardson" ground flat
oil hardeni ng AISI-0 1 carbon chrome alloy in untempered condition). 17 6061-T4 aluminum
alloy specimens and 14 glass-epoxy specimens with five different lay-ups. as wel l as of
seven steel specimens impacted through an intermediary to increase the duration of impulse
were reported in f 18.19]. Impact test resu lts on 8 1 composite specimens manufactured by
Domier GmbH, Germany, and 12 2024-T3 aluminum specimens were reported in [ 18.38].
T he composite specimens were fabricated from three types of fibers embedded in epoxy:
graphi te HT/ T300 (Torey), E-glass (Gevetex) and Kevlar (Dupont). For each material three
symmetrical lay-ups were manufactured: "U"-unidirectional (0°). "$"-shear (:!:45°) and
"M"-mixed (0°; :!:45°; 90°).
Typical plots of the maximum flexural response (e.) versus the peak magn itude of the
impu lsive compression (e..) are presented in Figure 18.22a and b. where each point corre-
--`,`,`````,`,````,,``,``

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1492 Buckling under Dynamic L.oads and Special Problems

VM •5
, 17m/~.~~
_,.l ·Goqe 1
VM 0 •6 ~mise<: r--
r\ IOOC ,._Sr.,.
! >7-;
r- '
I

I I

G(;!Qe2 Y\
I I~ 'i I

i 10001~ 51
I

I ~oobok
'{ _I'_
500 ~'- Sl I

I
I
Compren1cn

I
I 500 }! Sf\ I

-..L.. l msec
I 1m set Figure 18.21 Typical strain records of two separate impact tests
on a column: (a) V0 = 5.17 m/sec, (b) V0 = 6.19
m/sec, obtained in Ari-Gur et al.'s tests (from
[18.6])

sponds to one impact test. In this presentation of the overall behavior of a column, two
separate regions are observed; in the first region the slope of the curve is quite moderate,
while in the second region a small increase in the dynamic compression results in a much
larger increase in the bending response. Dynamic buckling occurs in the transition between
these two regions in accordance with the Budiansky-Hutchinson criterion (see [18.23]-
[18.25]), and it determines the dynamic buckling strain (scr) of each column, as well as the
Dynamic Load (Amplification) factor, DLF, given by

DLF=~ (18.41)
8 crL

where sccE is the static buckling strain

(18.42)

Here Ae~r is the effective slenderness of the column, defined by


(18.43)
where Le~r = Llr for clamped boundaries and r = htv?i.
A plot of the DLF versus slenderness ratio obtained in the tests of [18.19] is depicted in
Figure 18.23. It is apparent from the figure that the dynamic buckling capacity increases
with increase in the slenderness of the column.

I Eblmax
IEblmax
[,..strain] -Eccr[t.strai~ [i.strai~
3000 3000
3000
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

2000
2000

1000

0
~ 2000
- Ecmax (}' Strairi)
1000

~ 2000
-£cmax
4000
[tJ Strairi]

a STEEL ~. ALUMINUW

Figure 18.22 Typical maximum flexural response-maximum compression-amplitude plots (Budi-


Copyright Wiley ansky-Hutchinson plot) of metal columns obtained in Ari-Gur et al.'s tests (from [18.19])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Figure 18.23 Dynamic load (amplification) factor (DLF) versus slenderness for metal and composite
columns observed in Ari-Gur et al."s tests (from [18.7])

The effect of impact duration on the DLF's was also evaluated in the Technion investi-
gations of Ari-Gur et al. Results pertinent to this study for the different materials used in
the tests of [ 18.19] are shown in Figure 18.24. It is observed in this figure that the longer
the duration of loading T, the lower the DLF, whereas the type of material had no effect.
Additional experimental studies on elastic and plastic impact buckling of columns, all of
which were carried out between the beginning of the fifties and the early seventies, were
published by Davidson and Meier [15.17], Davidson [18.39], Erickson et al. [18.40], Hegglin
[18.41], Goodier [18.29], and Hayashi and Sano [18.31] and [18.32].
Erickson et al. (see [18.40]) designed and built a special testing machine for rapid con-
trolled loading to perform their tests at the Polytechnic Institute of Brooklyn in the mid-
fifties. This machine was later transferred to the Department of Aeronautics and Astronautics
at Stanford University and was modified and improved by Hegglin [18.41]. The test results
obtained with this machine demonstrated that the dynamic load amplification depended on
a dynamic similarity number proposed by Hoff et al. (see [18.42]):
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

DLF

8 "t•2
4
0
g. ""
~
8 "t•4
($>
0
0"
0

10'
8

6
o-Stul
4 o- Aluminum
ta.- Composi1eo

' 104

Figure 18.24 Effect of load duration and column slenderness on the DLF of metal and composite
Copyright Wiley
Provided by IHS Markit under license with WILEY columns tested by Ari-Gur et a!. (fromLicensee=McDermott
[18.7]) Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1494 Buckling under Dynamic Loads and Special Problems

( 18.44)

where a is the velocity of the load ing head of the test machine. Furthermore, good agreement
was found between the analytical predictions and test results when plouing the dynam ic load
amplification (designated as overload in [ 18.42)) versus the dynamic simi larity number (see
Figure 5 of [ 18.40) and Figure 18 of [ 18.41 ]).
In a recent experime/1/a/ investiJ(ation Zhang, Li and Zhcng a1 the Department of Naval
Arch itecture and Ocean Engineering, Huazhong Uni versi ty of Science and Technology, Wu -
han, P.R.C .. conducted tesL~ on the dynamic response, buckling and collapsing of elastic-
plastic straight columns subjected to relatively long-duration impacts induced by axial solid-
fluid slamm ing compression (sec f 18.43]). Using this technique they showed that, unlike in
columns that are exposed to short-duration loading and thus do not experience plastic failure.
compression-bending coupling developed in the long-duration slammed columns cou ld result
in large deformations, making plastic collapse possible. Furthermore. Zhang et al. concluded
that because of this behavior. the colum ns could experience different critical conditions
duri ng slamm ing. Indeed, in the tests three such critical conditions were observed: plastic
incipience, buckling and plastic coll apse.
The test rig used by Zhang et al. is depicted in Figure 18.25. It consists of an electric
motor (1), a 6-m high slamming tOwer (2), sl iding vehicle guides (3), an electro-magnetic
releaser (4), a 450 kg weight sliding vehicle hooked by the releaser and moving along the
guides (5). a 500 kg weight axial loadi ng device a\lached to the slid ing vehicle (6), the
details of which arc shown in Figure 18.26 (note that in this figure the axial load is applied
10 the specimen by the slammed bollom plate), and a 4- m deep water pool (7).
The Huazhong University tests consisted of 15 clamped-clamped columns of rectangular
cross-section (see Table I of Ref. 18.43) with lengths varying from 400-600 mm. T he
speci mens were classified into two categories: (I) large-slenderness columns for which the
static Eu ler strain e«c was smaller than the material yield strain e,. (these columns were found
to buckle elastically) and (2) middle-slenderness col umns for which the static Euler strain
e.,L was close to e, (these columns experienced plastic buckling). To detennine the deformed
profi les and plastic regions of the tested columns. pairs of strain gages were bonded face to
face along their sides (see Figure 5 of (18.43 ]).
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`

Copyright Wiley Figure 18.25 Huazhong University selllp for slamming buckl ing ( from [I 8.431)
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Impact-Induced Buckling Experiments 1495

o uter casing
reinforcement supp·o rt . ....;;~....::::;;;.,;:;:
upper sensors
4 upper boundary
Hank boundary
guiding rods
S(.X..::imcn
sliOin& guides
lower boundary
lower sensors
bottom pl::.h:
slam:niug load

Figure 18.26 The axial loading frame used in the slamming buckling testS of Huamong University
(from [ 18.43])

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
A typical ftuid-stn.cture slamming pulse obtained in the tests of [18.43) is given in Figure
18.27.

18.2.2 Plate Buckling


Static buckling of plates under in-plane compression loading and their response to lateral
dynamics have been extensively sntdicd. However, the dynamic response of plates subjected
to high rates of in-plane loading and similar loadings of short duration has not been thor-
oughly investigated, either theoretically nor experimentall y. Several of the few theoretical
studies o n plate dynamic buckling that have been published were reviewed by Ari-Gur et al.
in [ 18.301 and further discussed in [ 18.7]. These studies. together with those of [ 18.30), have
indicated that the main characteristics of dynamic buckling of plates were similar to those
obtained in the investigations on buckl ing of columns (see Subsections 18.1. 1-18.1.3 and
18.2. 1).
The 1981 Technion report and paper [ 18.30) included experiments on dynamic elastic
buckling of plates obtained in tests that were carried out at the Technion Aerospace Structures
Laboratory (ASL}. Further experimental studies on dynamic buckl ing of plates, metallic and
composite. that were pursued at the Technion were published by Ari-Gur and Weller in
[ 18.44) and Weller et al. [ 18.45] in the mid-eighties and by Abramovich and Grunwald
11 8.46] in the mid-nineties.
Except for these studies, no experimental results of elastic dynamic buckling of plates
were found in the literature, though the results of a theoretical and experimental investigation
on dynamic buckling of rectangular plates in sustained plastic compressive flow were re-
ported by Goodier in the late sixties (see [18.47) and [1 8.4)).
The specimens used in the tests of Goodier consisted of 6063-TS aluminum square tubes,
the sides of which were regarded as rectang ular plates, simply supported along their edges
undergoing axial compression (sec Figure 18.28).

Figure 18.27 A typical fluid-stn•cture s lamming pulse obtained in


Copyright Wiley
Provided by IHS Markit under license with WILEY
Huazhong University loading frame (from ] t8.43))
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1496 Buckling under Dynamic Loads and Special Problems

Figure 18.28 Buckling patterns observed in Goodier's tests of square tubes impacted at different ve-
locities (from [ I8.46])

Each tube was mounted on a push rod (see on the right in Figure 18.28), which was fired
from a standard rifle. Titc tubes were projected end-on at a machined and pol ished flat face
massive target steel slab. The slab and rifle were supported on an "I"-beam, and the rifle
was aligned optically to be perpendicular to the flat face of the slab. After firing, the tube-
plug-rod assemblage traveled along Teflon tracks, in which electrical contact pins were ac-
curately spaced for measuring the tube velocity.
It can be seen in Figure 18.28 that the sides of the tubes buckled in a mode consisting of
a single transverse half-wave and several half-waves lengthwise. These experimentally ob-
served buckling modes were compared with those predicted by the theory developed in
[18.47]. Fairly good agreement between the predicted modes and the experimental ones was
reported in (18.47) and [18.4]. Furthermore, it was indicated that the agreement improved
for tubes with well-developed buckling patlerns.
Some of the tubes tested by Goodier were sli t along all four corners to provide free
boundary cond itions. The dynamic buckl ing patlerns observed in these tubes are depicted in
Figure I 8.29. Reasonable agreement between theory and experiments was reported for these
tubes as well.
In the elastic dynamic plate buckling tests of Ari-Gur. Singer and Weller [ 18.30)the same
drop-weight test rig used in the tests of columns (sec Figure 18.20). as well as the measuring
and recordi ng devices used in the column experiments, were employed for impacting the
plates. Two test series of plates (the dimensions attd masses of which are given in Table I
of [18.30)) with different boundary condi tions were carried out within the framework of the
expe1imental program . The first series included plates with clamped boundary conditions
along the loaded edges x = 0, a and was designated by "PC." The unloaded edges y = 0,
b were symmetrical ly stiffened by angle stiffeners bolted to the plate. A "PC" -type specimen
positioned for testing is depicted in Figure 18.30. The second series consisted of simpl y
supported plates, designated by "PS," that were held and tested in a rigid fixture. In both
series there was no di rect collision between the striking mass and the edge x = 0 of the
plate, because an intemtediate mass (M,) was required to maintain the appropriate boundary
conditions along this edge (see Figure 18.30). In the "PC" specimens this intermediary (of
70 g) was auached to the impacted edge of the specimens, whereas in the "PS" plates the
intermediate mass (of 550 g) rested on the plate. Hence. the condition for no tensile force
along x = 0 was satisfied for the "PS" plates, whereas the condition of a straight impacted
edge could be fulfilled only in the case of the "PC" specimens, and even there on ly ap-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
proximately, due to the bending flexibility of the intermediary.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Impact-Induced Buckling Experiments 1497

Figure 18.29 Buckling panems observed in Goodier's tests of square wbes slit along the comers (to
simulate free boundary conditions) impacted at different velocities (from [ 18.46])

In 1he tests of [ 18.30] all lhe specimens yielded DLF's g reater than unily. T hese DLF's
were de1ermined from the Budiansky-Hutchinson type plots (see Figure 18.3), which were
obtained from reducing the data, recorded hy pairs of strain gages bonded face to face at
the center of lhe plate, into compression (load) and bending (response) strains. Furl hermore,
the DLF's correspondi ng to the " PS " specimens were fo und to be larger than those obtained
for the " PC" plates (see Figure 18.31). A reasonable correlation was found between test and
numerical results obtained by lhe analysis developed in [18.30]. This, however, was achieved
for an assumed shape of geometrical imperfection, which rough ly approximated the actual
o ne.
Ari-Gur and Wel ler [18.44] presented the results of further experimental stud ies with
alu minum and sleel plates of various geometries, aspect ratios and initial shape imperfections
subjected 10 shon- and long-duration impacts. The prime objective of these tests was to
determine whether a range of loading duralion exists. in the vici nity of lhe fi rst lateral
bending frequency of the plate for wh ich DLF < I , to identify and isolate most of the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 18.30 PC-type pla1e positioned for testing in the Tech-


nion drop weight teSI rig of Figure t8.20 (from
( 18.30])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1498 Buckling under Dynamic toads and Special Problems

4000
PC
j,
PS I
3000
I I sta~k
Ecr d
~;/ ol
[!'Strain]
I
2000
I
ol
I
I

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
1000

0~------"------~------~-----
0 1000 2000 3000

E cr,1 [fl. Strain]

Figure 18.31 Comparison of static and dynamic buckling strains of plates observed in Ari-Gur et al.'s
tests (see [18.30])

physical parameters, or their combination, that induce this phenomenon and find their role
in this undesirable behavior. Another objective of the study of [18.44] was to assess the
effects of various materials on the DLF's experienced by the plates.
In these tests the same drop-weight test rig (see Figure 18.20) was used for loading the
specimens, but a new loading frame was introduced (see Figure 18.32). This frame provided
a wide range of specimen geometries, any desired width and up to 400 mm in length, and
various combinations of boundary conditions (either simple support or clamping along each
edge of the plate). Also, the in-plane boundary conditions along the unloaded side edges
could be adjusted to either free or restrained in-plane motion. In the case of long-duration
loading tests (l0- 1-l0- 2 sec), the loading frame was placed in a universal MTS loading
machine and controlled frequency of head motion was used to load the plate.
To obtain the Budiansky-Hutchinson-type plots for the determination of the plate dynamic
buckling load, strain gages were bonded face to face at the center of the plate, and as in the
previous tests (see [18.30]) they provided the compression and bending strains necessary to
construct these plots.
In the two tests of [18.44] conducted m1der load durations corresponding to periods in the
vicinity of the first natural lateral frequerrcy (T"), DLF < 1 was observed. The DLF's ob-
served for all of the plates in this study are shown in Figure 18.33. It is apparent from this
figure that the results for the different aspect ratios and materials closely fit a single plot of
the DLF versus the duration of loading, normalized by the period of free lateral vibration.
The loading frame of Figure 18.32 was also used in the Technion extensive experimental
studies on dynamic buckling of composite plates carried out by Weller et al. (see [18.45]).
The objectives of this test program were to evaluate the effects of various parameters: ma-
terials (HT-T300 [Torey] graphite-epoxy. E-Glass [Gevetex] glass-epoxy and Kevlar [Du-
pont]); laminate layups (9-ply 1.25-mm thick mixed [0; +45; 0; -45; 90] 5 , 12-ply 1.50-mm
thick shear [ ± 45] 35 and 8-ply 1.0-mm thick quasi-isotropic [0; +45; 90; -451s); plate di-
mensions, length X width (150 X 150 mm 2 , 300 X 150 mm 2 and 225 X 225 mm 2 ), and
impact duration on the DLF's. (Note that the quasi-isotropic plates were only fabricated from
graphite-epoxy.)
The test results of this study (a total of 48 plates, 2 plates of each material, layup and
aspect ratio) were summarized in Table 7 of [18.45] and are depicted in Figure 18.34. It is
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Impact-Induced Buckling Experiments 1499

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Oeti il A: Side Vi e• of Upper 6 l ..,.r O.U i l B: Upper Vltw of Side Supports
Supports

Figure 18.32 Ari·Gur and Wellers loading fmmc (from [18.441)

apparent from lhis figure that a~ in the tests of [ 18.44 ]. the magnitudes of the DLF's were
mainly affected by the impact duration. Thus, as in [18.44 ). all the test results could again
be fitted into a single curve. DLF versus nondimensionaltime duration 2TITB (where TB is
the period corresponding to the first natural lateral frequency). As anticipated. high values
of DLF's were observed for short-duration impacts. In the vici nity of load durati ons corre-
spondi ng to the period of the nrst eigenvalue, the DLF's were close to unity.
In fU11her pursuing the studies at the Technion on the dynamic stability of impacted plates,
additional tests on laminated composite plates of various aspect ratios and boundary con<li-
tion~ subjected to axial impact were conducted by Abramovich and Grunwald (~cc [ 18.46)).
Fhe composite plates were included in this study: two 12 layers HT-T300 (Torey) graphite-
epoxy ~ymmetric shear type. ± 45. with an aspect ratio of 2 (300 x 150 mm'). one 9 layers
graphite epoxy of mixed lay-up (0:+45:0:-45:90),. with an aspect ratio of I (225 x 225
mm 2 ). one square (225 X 225 mm 2) 9 layers E-Giass (Gcvctcx) glass-epoxy of mixed lay-
up, and o ne square (225 x 225 mrn ' ) 9 layers Kcvlar (Dupont) of mixed Jay-up (fo r more
details sec Tables I and 2 of f 18.46)).
In the tests the specimens were posi tioned in the loading frame of [18.44] (sec Fig ure
I !!.32) and masses, the sizes of which were chosen in compliance with the required duration
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1500 Buckling under Dynamic Loads and Special Problems

~
I
I
I
._II
a
II

. ,__
----------
1 ------------------------

..
•,r--~----~.2~--.~l--~.4:---~.s~---.~s----~.7~---.s~---.~9--~
•.•
[2T/Tbl

Figure 18.33 DLF versus load duration observed in Ari-Gur and Weller's tests on dynamic buckling
of plates (from [18.44])

of impact, were dropped along a guiding rod onto an intermediate mass (see Figure 18.35).
As in the previous tests, the response of the plate, bending strain versus compression strain,
was obtained by reducing the data recorded by a pair of gages bonded face to face at the
center of the plate. Note that in order to determine the duration of loading relative to T" and
accordingly the size of the required mass, each plate was vibrated prior to its testing to
obtain its natural frequency.
Since determination of the dynamic buckling load of plates according to the Budiansky-
Hutchinson dynamic buckling criterion does not provide unique results, because the exper-
imental response plot does not exhibit a well-defined knee (see Figure 18.3), the use of the
Donnell approach for this purpose was recommended by Weller et al. (see [8.85] and [8.86]
of Chapter 8). Using this approach for the determination of both the static buckling load
and the dynamic one corresponding to a given plate provided the DLF's corresponding to
each of the tested plates. The variation of the DLF's versus the loading duration is shown

~28 ,
" 18 +
16 t
14 -+
I+
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

12

u It

8
~ :t ~T
++
---- -
+ ~
---+---±..- -+-
8 .2 .4 .6 .8 1.8 1.2 1.4 l.B 1.8 2.1 2.2 2.4 2.6 2.8 3.1
2T/T8

Figure 18.34 DLF versus load duration observed in Weller et al.'s tests on dynamic buckling of com-
posite plates (from [18.45])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Impact-Induced Buckling Experiments 1501

•· boundary condition

b- loading frame

c. specimen

d- droping ma,ss

c- guiding rod

r- impacted buis
1- straia-xagc

a-~~~~
Types of Boundary Conditions:
c
:r- Nominal S.S.

f-- ~aminal Clamped

Figure 18.35 Abramovich and Grunwald's loading frame (from [18.46])

in Figure 18.36. It is apparent from this figure that as in the previous studies with metal and
composite plates, large DLF's are associated with short load durations, whereas in the vicinity
of loading durations corresponding to the period of the first lateral natural frequency, they
become smaller than 1 (see also Table 3 of [ 18.46]).

18.2.3 Buckling of Arches and Spherical Shells


a. Arches
As indicated in Chapter 7, the shallow arch inherently possesses all of the nonlinear char-
acteristics of the much more complicated shell structure. Therefore, as in static buckling,
this rather simple model has been chosen for theoretical and experimental investigations on
the dynamic buckling behavior of stmctures. Altering the boundary conditions of the arch,
its geometry and loading, allows a wide variety of nonlinear behavior to be demonstrated
and evaluated (see also [7.9] and [7.10]).

• ~
II: I +csne

.
i
• '"""'
• ~
u..
5 1
\
., .
i
I

I•
• !
I
0 5
I
i

6 10 12
Timp/Tb

Figure 18.36 DLF versus load duration observed in Abramovich and Grunwald's tests on dynamic
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
buckling of composite plates (from [18.46])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1502 Buckling under Dynamic L.oads and Special Problems

Simitses [18.5, chap. 7] reviews all the relatively small number of investigations on dy-
namic stability of arches, most of which were carried out during the sixties. He evaluates
various concepts and methodologies introduced for estimating the critical conditions for
suddenly loaded arches since the first study of this problem by Hoff and Bruce in 1954 (see
[18.48]).
Humphreys [18.49] presented the results of an experimental and theoretical investigation
on buckling of clamped hard steel (130 kpsi) arches under impulsive loading, carried out at
Aeronautical Research Associates of Princeton, Inc. In the tests impulsive load was intro-
duced by detonating sheet explosive on the convex surface of the arch, and a modified split-
frame Fastax camera, with motion pictures taken at about 10,000 frames/ sec, was used for
recording the behavior of the arches (for details on the explosives, geometry and measure-
ments of each specimen tested see [18.49]). A pendulum swing provided a measure of initial
impulse. Each model was subjected to one blast only, and a qualitative evaluation was made
from the Fastax film record, whether or not the arch buckled dynamically during the test.
According to this evaluation, buckling occurred when the average deflection exceeded the
initial average arch rise.
In summarizing the test results, Humphreys indicated that the Fastax motion pictures
clearly demonstrated a critical range of loading above which the arches snapped through,
followed by an elastic return snap to their original position. Using the test results, which
exhibited noticeable scatter, it was possible to specify a critical range of impulse, above
which buckling definitely occurred and below which it definitely did not occur.
The test results were compared with predictions obtained by the analysis developed in
[18.49], where the nonlinear equations of motion were numerically solved using shallow

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
thin-shell assumptions and an approximate Galerkin technique (for details see [18.49]). The
arches tested appeared to buckle at about 60-70 percent of the impulse level predicted.
Experimental studies on dynamic buckling of arches were also included in the investiga-
tions of Cheung and Babcock, carried out at the GALCIT Caltech, which were discussed in
Chapter 7 (see [7.9] and [7.10]), and in [18.50]. In an attempt to clarify the dynamic behavior
of clamped arches subjected to very short-duration uniform pressure, the problem was re-
examined in [18.50] by tests conducted with 2024 T3 aluminum circular arches. Short du-
ration pressure load was applied in the tests of this study by a sheet explosive, and the
motion of the arch was recorded by a high-speed camera (HyCam, Red Lake Laboratories).
The fabrication procedures of the arches and their initial imperfection measurement were
identical with those described in Chapter 7 in connection with the static tests of [7.9] and
[7.10]. For details on the explosives used and their preparation, on the impulsive load itself,
the camera setup, the response measurement and test procedure, see Chapters 2, 4, 5 and 6,
respectively, of [18.50].
In the GALCIT experiments [18.50] no evidence was found that the clamped arch under
impulse loading could be rigorously categorized as a dynamic buckling problem. However,
it was observed that when the maximum response of the arch versus the applied impulse
(Budiansky-Hutchinson-type plot) was plotted, there were small ranges of impulse where the
arch underwent a significant increase in the response. From a practical viewpoint it is there-
fore important to determine this range of impulse. Furthermore, it was experimentally de-
termined that no stable equilibrium position other than the undeformed position existed for
the clamped arch free from lateral load. Also, no unstable equilibrium position was detected.
As indicated in [18.50], this confirmed the claim of Vahidi of non-existence of snap-through
for clamped shallow elastic arches subjected to impulsive load [18.51].
In [7 .9] and [7 .1 0] of Chapter 7 the dynamic stability of clamped arches under step loading
was studied. Tests were conducted on specimens identical to those used in the static tests
reported in these references and in the dynamic tests of [18.50]. The experimental studies
of [7.9] and [7.10] were complemented by a theoretical development of an energy-based
approach for determining upper and lower bounds for dynamic buckling of the arches.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Impact-Induced Buckling Experiments 1503

The loading apparatus used in the static tests of [7.91 and [7.10) wa~ used in the dynamic
te~ts reported there. ln the dynamic tests. however, the knife-edge device used to load the
arch was restrained with a pin from loading the arch (sec Figure 18.37). When the pin was
rapidly retracted, a step load was transferred to the arch. To mini mize the eJTcct of inertia
of the weight, it was auacheclto a soft spring. This re~u hcd in less than 10 percent variation
in the magnitude of the applied total load. Also. due to this arrangement the large variation
in load occurred after the arch buckled and iL' effect on the determination of the dynamic
buckling load was small.
TI1c step-loading tests of Cheung and Babcock ~ho" ed :1 very distinct dynamic buckling
phenomenon. They reported that the critical load wa~ detennined from the test results to
within half a percent for all cases. A typical test re~uh of these tests is shown in Figure
18.38. In this figure deHection-time traces of an arch for various loading levels (the numbers
altached to the curves) are presen ted. The closest pair of cu rves in this figure, wh ich ex hi bit
a pronounced change in behavior, defi nes the critica l load. The dolled straight lines in the
figure represent the equilibri um positions obtained in the static tests at the corresponding
lo:1cl levels. It is apparent from Figure 18.38 that the sutx:ritical responses were damped to
the nenr equilibrium position\ while the supercrilical response vibrated around the far equi-
librium position.
A comparison between the dynamic critical loads observed in the tests of [7.9] and [7.10]
and predictions by the energy approach developed in these references, and by a numerical
solution developed in 118.521. is presented in Figure 18.39. Good agreement between the
test results and the lower-bound solution of the energy approach and the predictio ns of
118.521 is observed. Note thnt the classical static buck ling loads that no rmalize the dynamic
crit ica l loads in Figure 18.39 were those of H.L. Schreyer and E. F. Masur (see 12.321).
As in the static tesL~ of [7.91 and 17.10]. the effect of load otTset on the dynamic critical
load wa~ examined. But in contrast to the static case. the influence of loading location was
found to be less important. Furthemtorc. the cusp-type behavior near the center of Figure
7.13 (of Chapter 7) was not observed in the dynamic te~ts.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure IH.37 Knife-edge assembly "'"din Cheung and Babcock's tests on dynamic stnbil ity of arches
(from [7.10))
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1504 Buckling under Dynamic Loads and Special Problems

• 06 IH
c

~
g OB
~
6
1.0

2H
1.2

Figure 18.38 Dynamic response of an arch at different loading values observed in Cheung and Bab-
cock's tests (see [7.10])

b. Spherical Shells
A spherical shell, like the arch, represents an ideal comparison example, though of a higher
order, for the assessment of various analysis methods developed for studying the complicated
dynamic response and instability of various shell-type structures, which pose formidable
difficulties to the analyst even today. Simitses [18.5, chap. 8] presents a historical perspective
and a detailed review of the problem, which was first reported by Suhara (see [18.53]), in
1960, and then in 1962 by Budiansky and Roth [18.54]. In the later study the frequently
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

used Budiansky-Roth dynamic buckling criterion was suggested and successfully applied for
determination of critical pressures as a function of duration time. This criterion was later
analytically verified by Budiansky and Hutchinson and proposed as a generic tool for de-
termination of dynamic buckling (see discussion in Subsection 18.1.2 of this chapter and
[18.23]-[18.25]). Further investigations that were not included in Simitses's survey were
presented by Simitses and Tabiei in a complementary article [18.55]).
Simitses [18.5 chap. 8] also presented a comparison between the results obtained by the
analysis developed by him and by those reported in his literature survey. Another thorough
evaluation and comparison of various methods for the analysis of dynamic buckling of shells

.... 1.2
e:"
§
.E
:ii
g ---~-
"'
"
.'1
"'8 0.

·=
0

~
-g 0.4
3 ---Upper Bound
"
~ - - - Eneroy Lower Bound
g 0.2 --Numerical Solution (Ref.18.52)
"'" -----0--- Pra!•ent Experiments

l" 0 4 12 16 20
Gel)metnc ~aTameter. r

Figure 18.39 Comparison of Cheung and Babcock's dynamic buckling test results with predictions
(from [7.10])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Impact-Induced Buckling Experiments 1505

was reported by Svalbonas and Kalnins in 1977 [18.56]. A shallow spherical cap was em-
ployed in this investigation for application of the various methods.
Hardly any experimental studies on dynamic buckling of spherical shells were reported in
the open literature. Humphreys, Roth and Zatlers [18.57] presented the results of shock-tube
experiments on the dynamic buckling threshold of four heavy steel clamped shallow spherical
shells under shock loading, carried out at Avco Corporation, Wilmington, Massachusetts,
before the mid-sixties. The aim of these tests was the examination of the applicability of the
Budiansky-Roth approach [18.54]. The experiments were conducted in the 6-ft ( ~ 1.83-m)
diameter shock tube at the Air Force shock tube facility, New Mexico. The pressure pulse
in the tests was generated by detonation of explosives at the closed end of the shock tube.
The length of the tube to the test section was about 150 ft ( ~ 45.7 m). Various instruments
were used in these tests: pressure transducers, a specially designed displacement gage for
measuring the central (maximum) deflection of the spherical shells, strain gages on the
concave side of the shells to determine qualitatively the occurrence of dynamic buckling,
and readout equipment. (For details see [18.57]. In particular, note the emphasis on pressure
calibration in Section III.)
The experimental results (Budiansky-Roth-type plots) obtained in the tests for a shell with
A = 5 are depicted in Figure 18.40 together with predictions obtained by numerical solution
of the Budiansky-Roth nonlinear differential equations (of [18.54]). Note in this figure that
A = [ 48( 1 - v 2 )(HI h)2] 1 14 represents the geometric parameter of transition between the
axisymmetric and nonsymmetric types of static buckling (where H and v are the rise of
the shell and Poisson ratio, respectively), T = ct/ R is the nondimensional time (where t is
the time, R the radius of the shells, c = (EI p) 112 is the stress propagation velocity, E is
Young's modulus and p is the material density), while Pa, is an average pressure over the
area of the model (see [18.57] for its determination). The central deformations in Figure
18.40 clearly indicate a sudden rise corresponding to a snapping phenomenon (below p =
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

6). However, unlike in static buckling, this change in response with increasing nondimen-
sional load parameter p (which represents the peak pressure divided by the critical static
pressure of a complete sphere of the same thickness and radius under hydrostatic load, and
as indicated by Budiansky and Roth is the appropriate load parameter to use with shells of
the same A) does not exhibit a clear jump, nor does it occur in exactly the same way at the

"r----,-----,-----.-----.-----r-----r-----r----,---~

\v
c
~W-(Cl

;;I W{Ol•ZH
·~,"
;•
t~:. I
L.Ll -
~
~

~ I
:.::
SHU S 3,4
- /-SM!Ll.. 2
tDor:~:_/
r------t--~'-d----t.,,r'<?-f-".__+---+-----H ~ :::~~ ~
0 '""'' - X• • '

= ~: :. :
0 :\,"' 4 I

• ro• X. •
: -HUM[ItiCAI.. SOLUTION SHfLL _
4
o

~ n+--Jf--t--l---t----1H ~ COM~UT£0 POINTS


t----1f----+--l+ r---
: ~1L 1 1. ~--,_--_,~
.......,. _I ~~ ~

i
:1 - WITH INITIAl PA[SSUAE; p 0 ., -
f l czr-~ ,.- SPIKE TO 3/Z P,., U:l..OW ~ PULSI:

~ T•l'""'"' - -1-t+-----T--+-~1-&---1
0
o~--~o~z----o~.~-~o~.~-~o~.~-~.~o---~.~.---~----~,.~---7
LOAD PAIIIAMET[.It i ~ P
0
"
~)
lE
.!!,
hoe:

Figure 18.40 Comparison of the maximum central deflections of the spherical shells measured in the
blast tests of Humphreys et al. with theory (from [ 18.57])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1506 Buckling under Dynamic L.oads and Special Problems

same values of the load parameter. The reason for this was explained by Huang in [18.58],
where it was shown that while a structure must possess more than one stable equilibrium
position under quasi-static loading for meaningful dynamic buckling, a clamped shallow
spherical shell does not comply with this requirement and its equilibrium position is its
undeformed configuration. Hence, if definition of the dynamic snap-through is based on the
finite jump behavior of deflection, then there is no possibility for dynamic snap-through
under impulsive loading (see also previous discussions on the GALCIT arch tests in [18.50],
as well as [ 18.51 ]).
It is also apparent from Figure 18.40 that dynamic buckling occurs experimentally at
slightly lower peak pressures than predicted (for A = 4.8). Figure 18.40 describes as well
the approximate shape of the impulse load observed in the tests that was used in the cal-
culations. Calculations for a modified impulse shape (the peak of the load at T = 0.1 as in
the above calculations and decaying back to the pulse shape at T = 0.2, rather than at T =
3) yielded the dashed curve in Figure 18.40. It can be seen in the figure that this curve
agrees better with the experimental results.
Lock, Okubo and Whittier reported the results of experiments on the snapping of six 7075
T6 aluminum shallow domes under a step pressure load, carried out in the low-density shock
tube of Aerospace Corporation, El Segundo, California, in the late sixties (see [18.59]). This
study was undertaken in order to determine dynamic weakening exhibited by typical snap-
ping systems.
Lock has shown [18.60] that the snapping process was governed by two distinct mecha-
nisms: a "direct" mechanism, associated with snapping occurring when the symmetric mode
deformations attain an unstable equilibrium state and the nonsymmetric mode deformations
are unimportant, and an "indirect" mechanism, where a complicated interaction between the
symmetric and nonsymmetric modes precipitates snapping. Furthermore, it was indicated
that the dynamic weakening effect was worst when the snapping was controlled by the
"indirect" mechanism. In this case the ratio of the natural frequencies corresponding to
the nonsymmetric and symmetric modes was greater than 0.7. In view of this, and with the
knowledge that in static buckling the transition between the direct and indirect mechanisms
of snapping is determined by the geometric parameter A = [48(1 - v 2 )(Hih) 2 ] 114 , the test
specimens of [18.59] were chosen such that A = 7.5. This value was about 50 percent greater
than the static buckling transition value and well within the range of the indirect mechanism.
Also, the aforementioned frequency ratio was about 1.2. It was therefore concluded that this
combination will demonstrate significant dynamic buckling weakening.
In light of these arguments, the tests consisted of three phases: nondestructive static buck-
ling tests, vibration tests and dynamic load tests (for details see [18.59]). In the dynamic
tests uniform pressure was generated in the shock tube by reflecting a normal shock wave
off the surface of the dome, and thus the load was instantaneously applied. To achieve the
conditions of a step loading of infinite duration, the domes were designed to collapse during
the time interval immediately after the anival and reflection of the shock wave, the available
"test time", during which the pressure load remained essentially constant.
Lock et al. reported that the critical step pressures were between 55-64 percent of the
experimentally observed static buckling pressures. Also, the measured supercritical response
of the shells was similar to the predictions of theory.
Extensive shock-tube experiments on the nonlinear response of polyvinylchloride (PVC)
plastic shells subjected to blast wave, carried out in the 6-ft ( ~ 1.83-m) diameter, 150-ft
( ~45.7-m) long explosive-driven shock tube of the Defence Research Establishment Suffield
(D.R.E.S.) Canada, were reported by Tawadros and Glockner in the early seventies (see
[18.61]). In this test program 32 spherical and paraboloidal shells, hinged or clamped along
their boundary, were tested (for details on material properties and geometries see [ 18.61 ]).
High-speed and television cameras were employed to monitor and record the shock wave
pressure and velocity as well as the defmmation of the shell models.
The lay-out of this blast simulator is shown in Figure 18.41. The shells were mounted in
the work section of the simulator on a specially designed mounting structure. In the design
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Impact-Induced Buckling Experiments 1507

I"
TUTI KCTION A
TEST SE:CTlON 8

I TEST SECTION C
lMOOELS WERE MOUNTED HERE 1
TEST SECTION 0
85'

'J
..
~ 6'

SEC A-A

Figure 18.41 Layout of six-ft diameter blast simulator at the Defence Research Establishment Suffield
(D.R.E.S.) Canada (from K.Z. Tawadros and P.G. Glockner, "Experiments on the Non-
Linear Dynamic Response of Shells under Blast Waves," Journal of Sound and Vibration,
26, (4 ), copyright 1973 Academic Press Ltd.)

of the mount particular requirements of rigidity and aerodynamic fairing were considered.
Details of mounting of the specimens in the shock tube are shown in Figure 18.42. The
special rings designed to support the models on the mounting structure are depicted in Figure
18.43. Particular care was given in their design to minimize aerodynamic disturbances and
to simulate, as closely as possible, nominal clamped (see Figure 18.43a) and hinged (see
Figure 18.43b) boundary conditions. To simulate the hinged-edge effect, a discrete hinge
mechanism, consisting of 100 0.5-in. (12.7-mm) diameter hard nylon ball bearings moving
in a toroidal groove of the same base diameter as the shell model, was designed. Each ball
bearing was slotted along its diametrical plane to accommodate the edge base of the shell

Figure 18.42 Details of mounting models in the D.R.E.S. blast simulator (from K.Z. Tawadros and
P.G. Glockner, "Experiments on the Non-Linear Dynamic Response of Shells under Blast
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Waves," Journal of Sound and Vibration, 26, (4), copyright 1973 Academic Press Ltd.)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1508 Buckling under Dynamic Loads and Special Problems

(a)

SECTION A-A

\ ~~~-~~r~~5=·a.="=I==U=P=PE=R=R=IN=G========3•1
~·~1~F:~ ~:...... -~~
"'t,; •t~
1-~" N.F. Threads
tor ~MU~~ttlnt
ttoltt

J
(b)

Figure 18.43 Ring supports of the shells tested in the D.R.E.S. blast tube: (a) rings providing clamped
boundary, (b) rings providing hmged boundary (from K.Z. Tawadros and P.G. Glockner,
"Experiments on the Non-Linear Dynamic Response of Shells under Blast Waves," Jour-
nal of Sound and Vibration, 26, (4), copyright 1973 Academic Press Ltd.)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(in compliance with the thickest shell in the series). Screws were used to apply the required
tightening pressure in the bearings.
Details on the instrumentation and test procedures were provided in Section 4 of [18.61].
Note that the behavior of the shells was monitored by the high-speed and TV cameras
through top and side windows (see Figure 18.42).
The test results of [18.61] revealed that the critical overpressure was sensitive to the radius-
to-thickness ratio (R/h) of the shell (see Figure 14 of [18.61]). A comparison of the test
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Impact-Induced Buckling Experiments 1509

results with those obtained in the shock-tube experiments of Humphreys et al. [18.57] and
Witmer et al. [18.62] is depicted in Figure 18.44. It appears from this figure that the test
results of [18.61] agree quite well with those obtained in the other investigations. The latter
tests [18.62] were performed on 6061-0 aluminum alloy spherical shells in the MIT-ASD
8 x 24-in. 2 X 98-ft (~20.3 X ~6l-cm 2 X ~29.9-m) shock tube, which provided step-
function blast waves.
Another experimental program on shock-tube impulsive tests on the static and dynamic
buckling strength of 2014-0 and 2014-T6 closed thin shells of revolution, carried out in the
Martin Company pulse generator, was reported by Bums in the mid-sixties [18.63].
In the early seventies Liu and Babcock performed an experimental study at GALACIT
Caltech to determine the dynamic buckling load of four 7075-T651 aluminum spherical caps
under impulsive loading [18.64]. In the tests the impulsive load was generated by use of a
spray-deposited explosive. Since, however, it was shown in [18.58] that the spherical cap
with no loading has only one equilibrium position, which corresponds to the undeformed
position, the caps were first statically preloaded and then subjected to the impulsive load. In
this case at least three equilibrium positions existed for a preload above a certain bound.
An energy approach for calculating the dynamic buckling loads was proposed in the study.
Its application required knowledge of the potential energy barrier that separated the buckled
and unbuckled states of the cap. Knowledge of the size of this barrier determined the amount
of kinetic energy that had to be introduced to the cap to make it buckle. Since the governing
equations of the cap are highly nonlinear, unlike in the case of an arch, analytical calculation
of the potential energy of the cap at each equilibrium position was a formidable task. Liu
and Babcock proposed that the potential energy levels would be determined directly from
static experiments. These experiments consisted of measuring the pressure-versus-volume
curve for each cap and numerically integrating the pressure volume relation to determine the
potential energy at each equilibrium position (see the special apparatus developed for these
measurements, the plots of pressure-volume and the potential energy-pressure plots in Figures
1-4 of [18.64]). Once the experimentally determined potential energy plots were given, the
impulse required to add a sufficient amount of kinetic energy to buckle the preloaded cap
could be calculated. Comparison of the experimental results with those predicted by the
energy criterion revealed that the test results were consistently considerably lower than the
calculated ones.

21 0

2·4 8
2·2
0
2·0

1·1
... B
0 0

--------------
1·6 0
.;

_r;::~--~J
1·4
'.; 1·2
1·0

0·8

06

0·4
0·2

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

10 3!1 40

Figure 18.44 Comparison of the shock tube test results observed by Tawadros and Glockner with those
obtained in the tests of Humphreys eta!. [18.57] and Witmer et al. [18.62]: O-Tawadros
and Glockner-clamped; D-Tawadros and Glockner-hinged; •-Humphreys et al.; 6 -
Witmer et al. (from K.Z. Tawadros and P.O. Glockner, "Experiments on the Non-Linear
Dynamic Response of Shells under Blast Waves," Journal of Sound and Vibration, 26,
(4 ), copyright 1973 Academic Press Ltd.)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1510 Buckling under Dynamic Loads and Special Problems

It was argued in [18.64] that the differences (25-50 percent) could not be attributed to
experimental errors alone. It was suggested that the static experimentally determined energy
barrier might not be the smallest one. Many postbuckled equilibrium positions could exist,
but the one found experimentally depended upon the static and dynamic characteristics of
the specimen and testing machine.

18.2.4 Buckling of Cylindrical Shells


The first investigation on dynamic buckling of a cylindrical shell subjected to a suddenly
applied load (compressive) was reported by Volmir [18.65]. He showed that in a shallow
cylindrical panel escaping motion, which determines dynamic buckling, occurred under a

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
load that was about 1.7 times greater than the static buckling load corresponding to this
panel. (It is worth noting that in this investigation Volmir showed that similar results were
obtained in the limiting case, when the panel became fiat.) A comprehensive historical review
of dynamic buckling of cylindrical shells, as well as a detailed mathematical treatise on the
problem (for thin imperfect, laminated, eccentrically closely stiffened, loaded by transverse
and eccentric in-plane compression and shear and with any kind of lateral and in-plane
boundary conditions), were presented by Simitses [18.5, chap. 9] and Simitses and Tabiei
[18.55], who surveyed investigations prior to 1990 that were not reported in Simitses's book,
as well as those published since 1990. Furthermore, this survey also addressed issues related
to the dynamic behavior of cylindrical shells, such as imperfection sensitivity, nonlinear
material behavior, orthotropic and laminated construction, seismic loading and explosion
conditions.
A general review on dynamic elastic and inelastic buckling of shells was published by
Jones [18.66]. This paper emphasized the various difficulties associated with dynamic buck-
ling, the elimination of which was required to enhance our understanding of the phenomenon.
Extensive formulations of the theory of elastic and plastic dynamic pulse loading of cylin-
drical shells, relevant experimental studies and a broad literature survey were also presented
by Lindberg and Florence [18.4, chaps. 3-5].
Simitses presented a comprehensive mathematical formulation for the analysis of dynamic
buckling of stiffened shells. Though these structures are extensively used as primary com-
ponents in complex structural systems because of their weight efficiency, and in spite of
their being subjected to destabilizing load, often of dynamic nature, relatively few studies
pertinent to their dynamic buckling have appeared in the literature. Among these are the
investigations of Dietz [18.67], Mayman and Singer [18.68], Lakshmikantham and Tsui
[18.69] and [18.70] and Jones and Papageorgiou [18.71]. The last deals with plastic buckling.
It should be noted that the theoretical results of both Simitses [18.5] and Lakshmikantham
[18.69] and [18.70] indicate that externally stiffened shells, though sustaining higher buckling
loads, are more sensitive to the presence of geometrical imperfections. This sensitivity be-
comes more pronounced with increase in stiffener cross-section area and is similar to that
occurring in static buckling.
Dynamic buckling of cylindrical shells can be divided into buckling under radial (lateral)
loading and under in-plane (axial and shear) loading. This distinction is also made in the
discussion that follows on the experimental programs, in particular on buckling under axial
loading, that have been reported in the literature.

a. Buckling of Rings and Cylindrical Shells Subjected to Radial Loading


It was shown (see [18.4]) that, depending upon the radius-to-thickness ratio of a ring or
shell, threshold buckling under impulsive and nearly impulsive loads can be either elastic or
plastic. As mentioned in Subsection 18.1.3, this buckling behavior is quite analogous to the
dynamic elastic and plastic flow buckling of long bars. In the case of long-duration loads,
buckling will be elastic and the critical load will be slightly higher than the static buckling
load
Copyright Wiley corresponding to the shell.
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Impact-Induced Buckling Experiments 1511

Because wavelengths associated with impulsive buckling are short relative to the shell
length, it can be assumed, for simplicity, that the impulsive buckling is independent of the
axial coordinate, i.e. the end supports of the shell become immaterial. Consequently, the
complexities of plastic flow buckling can be treated with a reasonably simple theory. When
buckling under long-duration loads is considered, the pressures sustained by the shell are
significantly lower and the stresses at buckling are elastic. In this case elastic shell theory
can be employed and the boundary conditions and variation of buckle amplitude with length
can easily be incorporated.
Experimental and complementary theoretical studies on dynamic buckling of cylindrical
shells fabricated from Type A Dupont Mylar polyester films and subjected to simultaneously
dead-weight axial compression with either axisymmetric transient hydrostatic pressures or
asymmetric lateral transient pressure were carried out by Wood and Koval at the Space
Technology Laboratories, Inc., Los Angeles, California, in the early sixties (see [18.72]).
The following types of transient pressure loading were involved in the program: ramp-step
of various magnitudes and rise time with axisymmetric pressure distribution; impulsive ax-
isymmetric hydrostatic pressure of variable magnitude; local ramp-step lateral pressure and
local impulsive lateral pressure.
One of the problems associated with the tests of [18.72] was how to produce rapid external
pressure engulfment of the cylindrical shell. The technique that was finally selected produced
a pressure differential between the interior of the shell and the outside atmospheric pressure.
Motion of an electromagnetically driven piston connected to the interior of the specimen
generated a negative pressure in the interior of the specimen, while the atmospheric pressure
acting on the exterior produced the required uniform inward force over the surface of the
specimen (see Figure 2 of [18.72]).
The experimental results of this program indicated that under the hydrostatic impulsive
loading the load-carrying ability of the shell was significantly greater than in the static
loading case (about four times). For the other types of loading, no significant differences
from with the static loading case were found. Qualitatively, the test results agreed quite well
with the simplified analyses based on Donnell-type linear shell theory, which was developed
to supplement the test program, but what was more important, they provided a better insight
into the mode of collapse of the shells under various loads.
Tests carried out at the Poulter Laboratories, Stanford Research Institute, Menlo Park,
California, were reported by Abrahamson and Goodier [18.73], [18.29] and [18.4], Lindberg
[18.36] and [18.4] and Anderson and Lindberg [18.74] and [18.4]. Tests of brief explosive
loading all around the cylindrical shell were conducted in [18.73]. Details on the procedures
involved in performing these tests are presented in [18.73] as well as in Subsection 3.2.7 of
Chapter 3 in [18.4]. Their aim was validation of the simplified rigid plastic model that was
derived in [18.73] for determining the dynamic plastic flow buckling of moderately thick
cylindrical shells subjected to uniform radial loading (details on the derivation of this model
can be also found in Section II of [18.29] and Subsection 3.2 of Chapter 3 in [18.4]). As
indicated by Lindberg (see Subsection 3.2.8 of Chapter 3 in [18.4]), the nonuniformities in
the experimental velocities did not allow a precise comparison of experiment with theory.
Alternatively, the circumferential number of crests in the buckled shape observed in the tests
was compared with a theoretical limit that depends upon the dimensions and properties of
the cylinder while it is independent on the nonuniformities. Such a comparison with the
observed results was presented in Figure. 10 of [18.29] and Figure 3.16 of [18.4]. It is
apparent from these figures that all observed data points agree with the theory (theoretical
limit), in that all the points lie below the limit line. (For additional discussions of the com-
parison between the test results and the theoretical model see Subsection 3.2.8 of Chapter 3
in [18.4]).
The test procedures of [18.73] were also employed in the tests performed by Lindberg
within the framework of an investigation on dynamic buckling of viscoelastic cylindrical
shells (see [18.75]). The theory developed in this study predicted buckling wavelengths in
general agreement with experiments.
--`,`,`````,`,````

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1512 Buckling under Dynamic Loads and Special Problems

Tile test resuiL~ of (18.73) were used to validate the theoretical model~ developed by
Florence and Vaughan for analy7ing the dynamic pla;tic buckling of shon cylindrical ~hells
subjected to uniform mdially inward impulses (see )18.76)), and the plastic now buckl ing
behavior of cylindrical shells caused by uniform rad ial ly inward impulses (sec )18.71]).
Lindberg [18.36] presented a theoretical and ex reriment investigation on buckl ing of a
very thin cylindrical shell due to a uniform radial impulsive pressure (see ah.o Section IV
of )18.29)). The re~uh~ of this study showed that when very thin cylindrical shells are
subjected to an external irnpuJ.,ive pressure. the intcr:tction between the radial. purely exten-
sional mode and the nexural mode was sufficient to precipilate permanent "rini..Jes. These
pla\tic large-amplitude wrinkle;, correspond 10 wa' clcngths. which depend upon the mag-
nilude of the pressure pulse. as well as on the cylinder parameters. The experiments con-
ducted within the framework of the program confim1 lhe hypothesis. which wa;, 1he driver
for lhe investigation. that "relatively small elaslic waves could precipitate plaslic waves of
signilicam ampl i1ude at a wavclenglh dictated by the claslic motion."
In 1hc ahove tests the cy linders were made of 606 1-T6 alum inum shee1 stock of 0. 125 -in.
(- 0.32-mm) thickness rolled into 12-in. (-304.8-mm) diameler, resulting in a radius to
thickness ratio of R /lr = 480. This material demons1ra1ed a very sharp knee at a yield stress
of 44.000 psi (-3W MPa). lmpul~e was applied by mean., of an explosive gas mixture (for
detail;, on this mixture. its detOnation and the experimenlal arrangement. sec )18.36)). Typical
buckling panems obtained in 1he lcsl~ under threshold impact and under an excessively large
impu lse (double of the lhreshold) arc depicted in Figure 18.45. Though the ~hell in Figure
18.45a was subjected to a pu lse sufficient 1o drive it beyond its elastic limi t, the number of
wrinkles (waves) cou ld still be predicted quite well by the clast ic theory. This was possible
bccau ~e 1he dominance of the prc fcrTed mode was well established. while the deformati ons
were still clastic (see Sub;ection 18. 1.3a). So far as the shell of Figure 18.45b i ~ concerned.
the number of wrinkles could be predicted by both the purely elastic and purely plas1ic
1hcories (see also Section IV of [ 18.29)).
To observe the transient formation of the type of" rinkle in the cylinders of Figure 18.-15,
a unique approach employrng the technique of Gerard and Becker ( 18.33] wa~ u~ed. The
wrinkles developed in the ruplllred buckled strips. 1he development of which was recorded
hy a hi gh-speed camera, were compared to I hose produced in the cylinder (for fun her details
sec )18.361).
Anderson and Lindberg )18.74) presented an el a~t i c model for the analysis of dynamic
pulse buckling of cyl indrical shel ls due to synunetric 1ransient radial pressure of duration
ranging from an ideal impu lse to one so long that the pul~e was a step load (see also
Subsection 3.5 of Chapter 3 rn ( 18.4 )). To cover thi;, range, three theoretical models were
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

u;,ed: a 1angen1 model for impul~e loads that produce plastic now buckling. a ;,train reversal

••

t • ..
t

·I
0

Fij!ure 18.45 Typical wrinkle; ob,crvcd in impulsively loaded cyli nders (from [18.36]): (a) thrc>hold
impulse. (b) doubled threshold pubc
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Impact-Induced Buckling Experiments 1513

model for loads of intermediate duration that lead 10 complicated elastic-plastic buckl ing and
an clastic model for long loads that induce elastic buckling. The experimental program
included in this investigation focused on mid- and long-duration loadings in order to verify
the above theory. Exten~ive experiments were run on aluminum and magnesium shells of a
length-to-diameter ratio Ll D = I and radius-to-thickne~~ ratios from 24-250. Explosive
techniques (explosive spheres and an explosive shock tube) 10 blast load the cylinders radially
were also used in this test program. The ~hells were loaded both symmetrically and asym-
metrically. Details on the test specimens. application of the explosives, the test techniques
and procedures and the test results arc given in [ 18.74] and Subsection 3.5 of 118.4]. It was
rcpot1ed that the experimentally observed critical loads. as well as the buckling patterns
(wave numbers), agreed well with theory. The buckling pressures under asymmetric loading
were found to be larger than for symmetric loading. More extensive damage was observed
in the shells which buckled under symmetric loading. Typical buckling panems associated
with various load durations are shown in Figure I 8.46.
Additional experimental programs that were included in investigations on dynamic buck-
ling of radially loaded cy lindrical shells carried o ut at the Poulter Laboratory were reported
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

by Lindberg [ 18.78 1 (stiffening and signi ficantly increasing the resistance of a cylindrical
shell to buckling due 10 an external irnpul~ive pressure by means of an elastic core), Vaughan
and Lindberg ( 18.8 11 (dynannic plastic buckling of sandwich shells) and Lindberg [ 18.80]
(Mre~s amplification and fracture in a ring caused by dynamic instability).
The effecrs of inirial imperfecrions on the dynamic buckling of 6061-T6 aluminum rings
and shells, subjected to a lateral impul ~e of a circumferential half-cosine distribution, were
studied by Kirkpatrick and Holmes at SRI International, Men lo Park, California. in the late
eighties [ 18.81] and 118.82]. Details of the test program. in which explosives were used to
introduce the impulse loading. are provided in these references. Prior to testing of the shells,
their initial imperfections were measured. The data obtained from these mea~urements were
transformed into idealiLed shape imperfections (see Eq. ( I) of I I 8.8 I]). as well as used to
obtain the measured di~tribution using fa~t Fourier transform~. The finite element DYNA 3D
computer code [ 18.83 1 was used to ca lculote the dynamic buckling loads in presence of the
measured shape im perfections (see deta ils in 118.83]). Good cotTelation between the ohserved
test results and the DYNA 3D predictio ns was found. Thus. the importance of incorporating
imperfection into dynamic buckling calculations was empha~ized.
Note that the influence of initial geometrical imperfection~ was also studied by Simitses
(~ee Chapter 9 of I I 8.5)). Additional investigations on dynamic pulse buckling of imperfec-
tion sensitive shells were reported by Lindberg in the late eighties [18.84]. In this study the
theoretical bases of two related but distinctly different dynamic buckling criteria were sum-
marized. The object ive was to demonstrate the range of applicabi lity of each so that together
they wou ld cover the entire range, from nearly impu lsive loads to step loads of infinite
duration. of dynamic pulse loading.

11"-tlll • ltl

Figure 18.46 tnnucncc of transient latcr:tl pressure magnitude and duration on dynamic pu lse buckling
pattern' in cylindrical ~hells (from H.E. Lindil<:rg und A.L. Florence. Vytwmic Pulse
Bud.fill/1. figure 1. 1. p. 2. copyright 1987. with kind permission from Kluwcr Academic
Publt shc")
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1514 Buckling under Dynamic Loads and Special Problems

b. Buckling of Cylindrical Shells Under Axial Impact


The advent of missiles and space vehi cles has created a great deal of interest in the dynamic
stability of cy lindrical shells subjected to dynamic axial loads. Their structural behavior
d uring impact is of prime importance because they have to sustain the sudden application
of thrust from rocket engi ne ig niti on, as well as enable a survivable landing of their payload
at their final destination (terrestr ial, lunar and other planetary surfaces). This latter require-
ment and the introduction of collapsible structures capable of absorbing impact loads, as
well as the enhancement of crashwonhiness, have further promoted the swdy of dynamic
buckling under axial loads. In these cases the investigations focused on postbuckling and
dynamic plastic progressive buckl ing, which can become a usefu l mechanism for energy
absorption, provided the failure modes and characteristics arc predictable.
Analytical and experimental studies on dynam ic buckling of cylindrical shells under axial
impact are summarized in Chapter 5 of [ 18.4]. A comprehensive literature survey was pre-
sented by S imitses in Chapter 9 of [18.5] and in the complementary review of ( 18.56].
Chapter 9 of [18.5 ] also contains detai led analyses of the problem.
In the sixties theoretical and experimental studies were carried o ut at the Space Science
Laboratory, Missile and Space Division, General Electric Company. Pennsylvania, by Coppa
and his co-workers [ 18.85]-[ 18.7)). Based on Coppa 's earlier experimental observatio ns. a
simplified hypothesi s was proposed in (18.85) and ( 18.86] in an attempt to explain the
mechanisms of buckling and postbuckl ing collapse of axially impacted thin-walled metal
cylinders. These earlier tests were performed by impacting the free edge of the shell vertically
with a Rat surface, either by dropping the specimen (with a weight attached at the end
opposite to the impact) against a flat base or by dropping a flat plate on the edge of a
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

stationary shell. These tests demonstrated that a thin-walled cylindiical or conical shell sub-
ject to a very short-duration impact buckled at the impacted edge in the no tably o rderly
fashion of the familiar diamond-shaped buckles. These buckles allowed an almost total axial
shortening of the shell surface with very little membrane (extensional) strain (see Figure
I 8.47). It was found that for shells of simi lar large radi us/ thickness ratios, the buckles were
of essentially similar form (see Figure 18.48). The patterns consisted of an array of folded
triangular planes, wh ich in the fully collapsed configuration of the cyl inder lay in the cross-
section plane. 1\vo such triangular planes, sharing a common base in the horizontal direction,
formed the fami liar diamond patterns.
It became evident from these observations that buckling occurred at the impacted end and
progressed discretely row after row at the veloci ty of the impacting head. Hence, during the
entire process there was a region of collapsed folded material adjacent to the impacting head

:.2s:
-
-
~~

lr
:l:i:

our

:S~

Copyright Wiley
.=-1~
~ Figure 18.47 Shortening of a cylindrical shell due to axial impact
observed in Coppa·s tests (from (18.86])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Impact-Induced Buckling Experiments 1515

(l>) .Uumlnuo (r/h = 125) .


Figure 18.48 Collapse pancrns in thin shells observed in Coppa's tests (from [ I8.86])

and an undisturbed region of the shell extending from there to the other end. Furthermore.
this local ization of the buckl ing and collapse occurred even when the unimpacted end was
laterally unrestrained on the base of the test apparatus.
During the collapse process. internal pressure was built up inside the shell. This pressure,
which was limited by an air vent, was sufficient to maintain roundness in the unbuckled
region of the shell during the collapse process. and thus contribu ted to the o rderli ness of the
buckle deformation, especiall y for very thin shells.
From these tests, in which the collapse configurat ion of the shell consisted of folded plane
surfaces, it was suggested that this configuration was an incxtensio nal deformation process.
Conseq uently. Coppa proposed a buckling mode l for a very thin cylinder (Rih - oo), in
which the cross-sections of the collapsed cylinder consists of polygons (see details in
[ 18.86)).
Furthermore. the model based o n this observation showed that buckling would develop
step by step with a propagatio n velocity (V) equal to the initial impact veloci ty at the shell
edge. Instabilities would form wi thi n each step. at the speed of the stress wave propagation
(c) in the undisturbed part of the cy linder. Under these condi tions Coppa (see I 18.86]) defined
buckli ng as "the change from cylindrical form to an inextensional form having a particular
wave number (n) and aspect ratio (wavelengths) which provide. by inextensional deforma-
tion, the same unit axial shortening, as the ax ial compression strain which existed in this
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

form."' In this case, the critical buckl ing fonn is that configuration. whose axial leng th is the
sho rtest to satisfy the condition e" s VI c. It was further shown that the shorter the axial
wavelength. the hig her the number of ci rcumferential waves, the larger the critical stresses
and the shorter the buckled segment of the shell.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1516 Buckling under Dynamic Loads and Special Problems

Coppa funhcr argued that because the impacted edges of the cylinders were assumed to
be free to move laterally, they could suddenly change from a circle to a regular polygon
when buckling had been initiated. Therefore. the initial buckle consisted of a single row of
axial half-waves, i.e. a row of triangular planes. However. under high impact velocities the
edges could not move freely in the radial direction and therefore axisymmetric buckling ncar
the impacted edge was anticipated. In thin-walled cyli nders an asymmetric diamond buckling
configuration would develop with further propagation of the local ized buckles (sec Figure
18.49).
Add itional tests were performed to validate these experimentally based models, including
rigid impact unpressurizcd tests. The tests studied the effects of internal pressurizati on and
nonrigid impact on the dynamic stability of axia lly impacted cylinders [ I 8.851. Later an
experimental st udy was carried out on the e ffects of edge conditions on the buckling behavior
of cylindrical shells under axial compression impact [ 18.87]. In this investigation the effects
due to (I) inward radial displacement restraint of the shell wall at the impacted edge. (2)
the asymmetry of the applied loading caused by the obliquity between t.he impact plate and
the impacted end cross-section plane of the shell. and (3) impact velocity were evaluated.
This investigation included rigid impact unprcssuritcd and pressurized experiments.
In the rigid unpressurized tests the General Electric Company Space Sciences Laboratory
Drop Tester (shown in Figure 18.50a) was used. This apparatus consisted of a 40-ft (- 13.7-
m) long hardened steel shaft, which was auached under tension between two vertically
aligned points. Along this shaft a carriage rode concentrically by means of two ball bushings.
The impact head was a hardened and ground steel disc, 10 in. ( - 254 mm) in diameter.
anachcd to the carriage. The drop head (carriage) assembly was hoisted to the requ ired drop
height by means of an electric winch fastened to the carriage by an electromagnet. A hard-
ened and ground steel base, 8 in. ( -203 111111) thick and 20 in. ( -508 mm) in diameter, was
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

mounted at the bollom end of the shaft and located concentrically with it. For accurme
positioning at the base in the horizontal plane. the hase was mounted on three adjustable
screws. For final accurate alignment of the head with the impact end of the shell, the carriage
could also be adjusted.

FiAurc 18.49 Axisymmetric buckling followed by asymmetric buckling observed by Coppa in axial
impact tests of thin cylindrical shells (from [18.86])

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Impact-Induced Buckling Experiments 1517

G UIOE WIRE_..- ~ELEASE


MECHANISM

HEAD
COUNTERWEIGHT
FOAM PAD - - l - --ift>
SHOCX
ABSORBER

{a) {b)

IMPACT
1-EOIUM

{c)
(d)
Figure 18.50 General Electric Company Space Sciences Laboratory Drop Tester {from [ 18.85]) (a)
apparJtus used in rigid unpressurized impact tests of shells, (b) apparatus for rigid impact
tests of pressurized shells, (c) device fo r internal pressurizing of cylind1ical shells. (d)
apparatus for impacting shells into nonrigid media

T he drop head assembly contajned potts for rapid venting of the air from the cyli nder
upon collapse. A buildu p of internal pressure was thus prevented . The vent area could be
adjusted in compliance wi th the specific requirements of the test.
T he specimen and its retainer were mounted on the base concentrically with the shaft. To
meet this requi rement the shaft had to be raised clear of the base. This was done by a
manually operated winch.
The apparatus on which the experiments of pressurized shells under rigid impact were
performed, which is a modi fication of the device of Figure 18.50a, is shown in Figure 18.50b.
T he vent chamber of Figure 18.50a was replaced by a cross-member, which could be relea.~ed
by two mechanically actuated release mechanisms mounted on the beam extremi ties. A drop
head was altached to one of the mechani sms and a counterweight to the other. The shell
specimen, in its retainer. could also be suspended from the beam for tests in whi ch the
specimen itsel f was dropped. In the device of Figure 18.50b the base was moved to be
located ven ically under the center of the drop head. The specimen. in its retai ner, was secured
10 the base via the central hole of the base.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1518 Buckling under Dynamic V:>ads and Special Problems

The cross-beam assembly, including the drop head and the counterweight, were hoisted
to the drop height and released (the drop head could also be dropped alone). When the cross-
beam assembly was dropped, it was decelerated by a shock absorber located at the lower
end of the shaft. The initial deceleration force on the beam assembly was imposed via a
foam pad to reduce the shock associated with the release of the drop head.
In order to preserve the unsupported edge condition of the impact end of the shell (as in
the unpressurized tests) and eliminate the different types of constraint imposed on the edge
by an end closure, a special pressurization technique was devised for the experiments with
pressurized cylinders. The pressurization rig (see Figure 18.50c) utilized a thin 0.0008-in.
( ~0.20-mm) thick Mylar membrane to form an internal cylindrical pressure-tight bag. At
the impact end the membrane was supported by a plate, which was attached to the cylinder
mount via a transmission chain. This arrangement relieved the membrane from a large portion
of the axial pressure loading by transmitting it to the base fixture, thereby greatly reducing
the deflection of the membrane. Upon collapse of the cylinder, the chain readily folded into
a cavity in the bottom of the mount and consequently did not interfere with the axial short-
ening of the cylinder. Apparently, this pressurization technique generated radial pressure on
the shell because the axial force was neutralized by the tension in the chain and in the
membrane. Axial force could only be introduced by friction between the cylinder and the
membrane (for further details see [18.85]).
The device used for performing the nonrigid impact loading tests is depicted in Figure
18.50d. It is essentially similar to that used for the pressurized tests (Figure 18.50b and c),
except that the pressurizing apparatus was replaced by a cylindrical tank 4 ft ( ~ 1.22 m) in
diameter and 50 in. ( ~ 1.27 m) long, which allowed free dropping of specimens into fluids
and other deformable or penetrable media. For viewing purposes the tank was equipped with
a window, the dimension of which provided an undestructed view of the specimen before
striking the impacted medium and subsequently during the impact process.
The experiments were conducted with unstiffened cylindrical shells fabricated from alu-
minum alloys 2024 F, 2024 T3 and 5052 H-38 and I /2 H 301 stainless steel. The shells
had the following dimensions: 5.70 in. ( ~ 145 mm) inner diameter, 22.8 in. ( ~580 mm)
length and 0.004 in. (~0.1 mm), 0.008 in. (~0.2 mm), 0.016 in. (~0.4 mm) and 0.019 in.
( ~0.5 mm) wall thickness. (These correspond to a length-to-diameter ratio (LID) = 4 and
radius-to-thickness ratios of 712, 356, 188 and 150.) Details about fabrication of the speci-
mens are given in [18.84] and [18.85].
Instrumentation was provided for the following measurements (see [18.85] for details):
strain measurement, acceleration measurement, transient pressure and high-speed photogra-
phy. All recording components were initiated by a common trigger, which consisted of three
pivoting rods. The triggering circuit was activated upon contact of the drop head with the
triggering rods. The triggering rods were positioned such that the oscilloscopes were acti-
vated prior to impact on the shell. The test results of [ 18.85] and [ 18.86] were compared
with the magnitudes and trends of the predictions that resulted from Coppa's hypothesis for
the mechanism of buckling due to axial impact. The test results of [18.87] demonstrated
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

that:
"l. Restrain of radially inward displacements at the impacted edge is effective in increasing
the initial buckling load. Such restraint also tends to produce symmetrical buckling
modes local to the restraint.
2. Although the initial response can be strongly influenced by the end restraint, the sub-
sequent response can be virtually independent of the end conditions.
3. Loading asymmetry results in a relatively low response depending of the degree of
asymmetry.
4. Higher impact velocities tend to produce higher buckling stresses, but the range of
velocity over which this occurs may depend on the particular mechanical, thermal, and
geometrical properties of the structure.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Impact-Induced Buckling Experiments 1519

5. The preferred buckling mode for sufficiently high velocities is symmetrical even for
thin, unpressurized cylindrical shells."

An axial impact loading apparatus employing a different approach to impact the cylinders
than that used by Coppa (Figure 18.50), was used by Lindberg et al. in their experimental
investigations, is depicted in Figure 18.51 (see Lindberg and Herbert [9.210], Subsection
5.15 of Chapter 5 in Lindberg and Florence [18.4] and Lindberg, Rubin and Schwer [18.88]).
This apparatus was used to impact the shells simultaneously around their circumference at
an accurately known time. This objective was achieved by attaching the shells to a massive
internal end ring (12 times the weight of the shell), the impact ring, with epoxy (see details
on application of the epoxy in [18.88]), and a band clamped around the outside of the shell
to avoid severe crimping at the end (which could rapidly reduce the applied thrust). The
other side of the shell was free. The unsupported length of the shell, opposite the impact,
was determined in compliance with the objective of the tests; a length such that buckling
would take place before the stress wave was reflected from the free end, i.e. compressive
impact stress duration, at most, equal to 2 Ll c of the longitudinal stress wave up and down
the shell (see [9.210] and [18.4]); and also shorter shells such that the reverberations of the
induced axial stress wave would result in several axial oscillations from compression to
tension [18.88].
The shell-impact ring unit was placed on top of a massive anvil bar, which had its upper
surface accurately ground and lapped to mate with the impact ring, which was similarly
ground and lapped (to obtain higher impact velocities, the solid anvil bar was replaced by a
long, hollow thick shell). Then a sheet of explosive was detonated at the base of the anvil
bar, sending a step-fronted shock into the anvil. (This shock entered the ring and bounced
off it in much the same way as end pellets are bounced from a Hopkinson bar.) The pressure
gradient behind was shallow enough for the reverberating stresses in the impact ring to be
small. Hence, to a good approximation the ring was stress free upon impacting the cylinder
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

8" 0-11' Tilt Cylinder

I mPKt Ring. Steel


0.4" Thick x 1.5" Height
Dece-.ting Tabs - to
Retain Cylinder after
TIJI. 3 Placos

I
Explosive Pad (0.030" to 0.200" Thick)

I
Figure 18.51 Lindberg et al.'s impact test apparatus (from [18.88])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1520 Buckling under Dynamic Loads and Special Problems

(in practice, the ring had a small residual stress, which resulted in a low-amplitude, high-
frequency oscillation in the stress wave imparted to the shell).
The remainder of the apparatus shown in Figure 18.51 (which was not included in the
tests of [9.210] and [18.4]) was required to decelerate the anvil bar and deflect the blast
wave from the test cylinder. Decelerating tabs with crushable Hexcell pads at three locations
around the impact ring were employed to bring the specimen to rest without damage. High-
speed photography was used to determine the ring velocity, resulting from various thick-
nesses of sheet explosives in a series of experiments, with the ring free to fly away from the
anvil bar. Ring velocities were found to vary linearly with explosive thickness up to velocities
as high as 15.3 m/sec.
High-speed photographs were taken with a Backman-Whitley framing camera running at
240,000 frames/sec in the tests of [9.210] of Chapter 9 and [18.4] (which was fast enough
to monitor details of early buckling wave formation) and with a Hycam motion picture
camera running at 40,000 frames/sec in the tests of [18.88].
The objective of the experimental study of [9.210] of Chapter 9 (also discussed in Sub-
section 5 .1.5 of Chapter 5 in Lindberg and Florence [ 18.4]) was to demonstrate that in
contrast to static buckling, the simple linear theory derived in this study should adequately
predict the large-deflection dynamic buckling of a thin cylindrical shell, with initial imper-
fections approximated by white noise, subject to axial impact. It was argued there that
because dynamic buckling started with a linear theory pattern, the shell would continue to
deform in this pattern because it would not have time to convert to another pattern. This test
objective dictated that the behavior of the shell should be monitored, particularly during the
early motion of the shell. Consequently, this imposed the aforementioned requirements on
the shell length and required the use of a very fast camera. The experiments have confirmed
this hypothesis. Furthermore, they demonstrated that the shell could sustain dynamic buckling
loads higher than the static critical load.
Comparison of the high-speed photographs taken during impact of the shells with the
permanent buckles remaining after impact revealed that most of the observed buckling was
elastic. Most of the permanent buckles, which were of small size and large aspect ratio, were
confined to the area close to the impacted end. A comparison of these dynamic buckling
patterns with those of a statically buckled shell of the same material and dimensions is shown
in Figure 18.52. Note that in the static case the aspect ratio of the buckles is near unity,
whereas after impact the aspect ratio is much larger. The wavelengths observed in the dy-
namic tests compared quite well with the predicted ones.
The experimental investigations of Lindberg et al. [ 18.88] dealt with dynamic buckling of
thin cylindrical shells under oscillating stress waves following axial impact. This type of
buckling behavior imposed the requirement that the shell be short enough to allow for the
reverberating axial stress waves to result in several axial stress oscillations from compression
to tension. The carefully conducted experiments in this study, compared with calculations,
demonstrated that essential features of the buckling were predicted by the theory. In cases
of pulses comparable in duration to the hoop breathing mode of the shell response induced
by the Poisson effect, circumferential resultant forces and symmetric deformations played a
major role in initiating buckling. Furthennore, the stress oscillation allowed the buckling,
initially localized near the impacted end, to propagate up the shell towards the free end. The
radial deformation of the breathing motion provided a deterministic mechanism for initiation
of symmetric buckling that dominated the early response under supercritical impact loads.
This is apparent in the high-speed photographs (Figures 9 and 11 of [18.88]). This was
observed more clearly with increase in impact velocity, which corresponded to higher applied
load relative to the static buckling load. The circumferential stress resultants excited hyper-
bolic growth with time of a large group of asymmetrical modes for subcritical impact loads.
Hence, there existed a continuous range of axial impact loads extending to values below the
static buckling load for which dynamic buckling occurred within the scope of classical
buckling theory. --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Impact-Induced Buckling Experiments 1521

f •) n/oc1 • 1.5&

(b)

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
141 •foe.~ • 1.» (a)

fi~u~ 18.52 Compari;on of d) narnic buckling panems afler 3Xinl impact at lower end' of a ~hell (a)
with buckling pancm under gradual 3Xial cornpre;;ioo (b) (from [18.29])

Furthennore. it was observed that. as in a si ngle-pu lse buckling, dynamic buckl ing under
finite-duration loading was fundamentally a response pheno menon and that the corresponding
criterion fo r critical loads was o ne of allowable growth. Hence. an important feature of thi s
growt h in the case of multiple pu lses was that the region of buckling spread over the shell
during the intervals between compression pulses. Therefore. because of the result ing in-
creased energy absorption capacity of the shell within an allowable stress or dcfonnation.
the integrated impulse of the multi pulse loading could be >Ubstantially larger than the critical
impul\e for a single pulse.
Clo•er inspection of the high->peed photographs revealed that in the case of mu ltipulse
buckling, because the motion changed from hyperbolic to trigonometric. the buckles grew
duri ng the fi rst and second axial compression compressive pu lses and propagated up the shell
as bendi ng waves during the interveni ng axial tensile pu lses. After the tests the shells re-
turned to thei r o rigi nal shape. indicating that the observed buck ling behavior was clastic.
£tperimemal and till'oreliwl swdies on dynamic buckling of circulat cyli ndrical shells
were carried out at the Institute for Aerospace Studies. University of Toronto. during the
~iJttics (see Tulk [18.89) and Zimcik and Tennyson (9.209)). These studies were summarized
by Tennyson [18.90(. The ba~ic objective of these ~tudies was to obtain reliable experimental
data on the behavior of ncar-perfect and imperfect cylinclrical shells subjected to dynamic
loading. The test programs. therefore. included three important steps: (I) manu facturing of
qhell &peci mens to very precise tolerances; (2) means fo r measuring the shell surface accu-
rately and chatacterizing its ini tial shape imperfections (those arising from the manufacturing
technique and those deliberately machined into the speci mens for experimental purposes);
(3) a method of testing the shells under controlled and reproducible conditions.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1522 Buckling under Dynamic Lc1ads and Special Problems

Here, very carefully and precisely manufactured photo-elastic plastic test models were
employed in both test programs (see [18.89] and [9.209] for details of their manufacturing
and initial shape measurements). These specimens were tested in specially designed dynamic
testing machines.
A schematic view of the testing machine used in Tulk's tests [18.89] is shown in Figure
18.53. It was capable of generating controlled ramp-type loads at rates ranging from quasi-
static to higher than 200 kpound/ sec (-90,720 kg/ sec) and accurately measuring the axial
end-deflection of the shell during the loading-buckling cycle.
To operate the machine, both the upper and lower chambers of its loading cylinder (see
Figure 18.53) were filled with compressed gas. By venting of the gas in the upper chamber,
the pressure differential across the piston caused a compressive force on the shell. The rate
of buildup of this compressive force was controlled by the size of the venting orifice and
the initial pressure in the chamber. High rates of loading were obtained by allowing the
Mylar diaphragm in the main vent to rupture.
The instrumentation employed in these tests is described in Subsection 3.4 of [18.89]. It
also included a high-speed movie camera. With the aid of a photo-elastic technique and high-
speed photography, the development of internal prebuckling strains in the shell wall could
be observed at impact.
The test results were compared with predictions obtained by a nonlinear Donnell-von
Karman-based theory, developed in [18.89], and close agreement was found. A typical com-
parison of dynamic buckling loads versus axial deflection rate is presented in Figure 18.54.
The aim of the University of Toronto experiments was the determination of dynamic
buckling response of circular cylindrical shells to square-pulse transient loading. The gas
loading machine of Figure 18.53 could not provide the finite load time duration. Therefore,
a different experimental technique was employed to load the cylinders. In this method a
free-flying sandwich-type of shell impactor, whose length governed the impact duration, was

l....oling Scrowt

Figure 18.53 University of Toronto controlled ram loading testing machine (from [18.89])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Impact-Induced Buckling Experiments 1523

1.5

14
1.3 0

1.2 "
1.1

1.0
"0
0
"
0.9

'"·"
ll. 0.8
~ 0.7
.3 06 Near Perfect Shells

..a
<t:
0.5
04
- - Theoretical Model

Experimental Results
0.3 0 Shell I
02 IJ. Shell 2

01

0
Ax.lol Deflection Rote v.

Figure 18.54 Effect of axial deflection rate on the dynamic buckling loads observed in Tulk's tests
and comparison with theory (from [18.89])

used (see Figure 18.55). This dynamic impact apparatus was used in the tests of Zimcik and
Tennyson (see [9.209]) and Zimcik [18.103]. In this testing machine a gas gun was used to
propel a hardened steel projectile down a 2.5-m barrel to impact on the striker shell (for
details on its fabrication see [9.209]). The inside motion of the projectile was transferred to
the striker shell by means of a steel piston, attached by an aluminum cross-beam to an
adjustable outside ring, which guided the striker ring (see Figure 18.55). A restrain on the
displacement of the piston-ring assembly limited the contact time with the striker. The striker
moved axially on bearings to impact the test specimen, which was also free to travel axially.
To ensure uniform impact over the shell cross-section, the machine permitted alignment of
the striker and specimen.
In the test apparatus of Figure 18.55 the end conditions of the specimen approximated
simple edge conditions. The radial and circumferential displacements were restricted by
friction due to loading, but the shell was relatively free to rotate. Circularity of the specimen
was maintained by the supporting bearings.

?~-~-"""'--'
_;:_"~ -
-~~-----~"~-,~~~~~---.•-•It
-;t------------------H- -
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Ptlron-c,~on~• Asnmbly

COMPLETE ASSEMBLY

Ahg'11TM!nl AdjUSI!Tief'll
Ac.c.eleroti:M' RmQI

"<"6orrtl

Vol~e
Aclr~OilnQ
Cyhnder
"''"
Aeservorr Cross·Deom
C"ou- beam Glildl

PIS TON - CYLINDER ASSEMBLY

Figure 18.55 University of Toronto dynamic square-wave axial loading apparatus (from [18.103])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1524 Buckling under Dynamic Loads and Special Problems

The response of the shells was recorded by strain gages bonded back to back on their
surface. A 16 mm Hycam high-speed camera, filming at rates of 1500-6000 frames I sec,
was used to supplement the strain gage data. As in the tests of Tulk [ 18.89], the high-speed
photography allowed observation of the development of deformation buckles in the shells at
impact.
A Budiansky-Hutchinson-type criterion, combined with Southwell plots, was used for the
determination of both the experimental and theoretical dynamic buckling loads of the shells
employed in the test program of [9.209]. Experimental and theoretical plots of maximum
vibrational amplitudes versus input stresses were constructed (see Figure 18.56a). These data
were used to construct Southwell plots from which the buckling loads were obtained (Figure
18.56b). The shells were subjected to impacts of various load duration to evaluate the effect
of load duration on the dynamic load amplification. A typical comparison of test results with
analytical predictions of variation of dynamic buckling stress with load duration presented
in Figure 18.56c (note that Figure 18.56a-c relate to the same shell, ASl).
The good agreement between the test results and analysis apparent in these figures was
typical of all of the shells included in the investigation of [9.209]. Hence, it was concluded
that the simplified analytical model chosen in this study adequately accounted for the prin-
cipal mechanisms governing the behavior of the cylinders. Figure 18.56c reveals that for
short duration of loading, the buckling capacity of the shell increased dramatically. It was
found in [9.209] of Chapter 9, that even for relatively imperfect shells the experienced
dynamic buckling stresses were higher than the static ones, provided the duration of loading
was short, comparable to the free vibration in the static buckling mode. However, the pres-
ence of initial geometric shape imperfections in the wall of the shell, which were geomet-
rically close to the classical static buckling mode, could substantially reduce the dynamic
buckling stress.

18.3 Propagating Buckles

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
In designing buckling resistant structure>., application of the classical approach, which is
concerned with their onset of instability, i.e. the determination of the smallest load or de-
formation at which alternate states of equilibria become possible, is commonly accepted as
adequate. Usually this is established through linearized bifurcation analysis. In practice,
however, the presence of initial geometrical shape imperfections, material imperfections and
nonlinearities and other factors influences the critical conditions. Incorporating them into the
problem analysis requires solution of the more complicated nonlinear models.
In the last two decades a class of structural instabilities, the analysis of which by the
classical approach is insufficient, has emerged. In general, these instabilities affect structures
of larger size that, following their onset of instability, exhibit localized collapse. Under
prevailing conditions this initial collapse can propagate or spread, often in a dynamic fashion,
over the rest of the structure. It has been observed that a substantially lower load was required
to propagate such instabilities than to initiate them.
Propagating instabilities in structures have been extensively investigated, experimentally
and theoretically, by Kyriakides et al., who contributed the majority of publications related
to this phenomenon to the literature. Recently, Kyriakides published a state-of-the-art article
in which, via thorough experimental description and analytical-numerical simulations of four
different structural examples that exhibit this type of behavior, the various aspects associated
with this phenomenon and its analytical treatise were clarified [18.91]. Kyriakides stated that
"In practice, propagating instabilities are initiated from local imperfections which locally
weaken the structure and cause local changes in its geometry (collapse). Very much like a
set of dominoes once the geometric integrity of such structures is compromised locally, the
instability has the potential of spreading over the whole structure." He also demonstrated
that "the underlying common characteristic of structures exhibiting this class of instabilities
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Propagating Buckles 1525

500 11XX) 1!;11)


STRAIN
(b)

!1000 lN95~~\.
\
$,

i
' '',,,~'
I!IQI:II

''•,,,..,_ _______ ~--


------o;
~, EXP
A.''•~"ALYSIS. 1 ~ I I() o Exp
- Allalysrs !Firite Decoy Time)
Al'JALYSIS. T = 225 pSJ:.'C ---Analysis, A• ua

L___--=-----o::;:--,c;;;,.;----=-- ·~.~~---~~~----~-­
5TMSS (po• 1
LOAD DURATION (T) ps

(a) (c)

Figure 18.56 Typical test results obtained in Zimcik and Tennyson's tests (from [9.209]) (a) growth
of vibrational amplitude with increasing stress, (b) Southwell plot, (c) effect of load
duration on dynamic buckling stress

is a nonlinear, local load-deformation response with two branches with positive slope joined
by an intermediate one with negative slope, as shown in Figure 18.57. Thus, for some range
of loads, such structures have three possible equilibria for each value of load. The one on
the first ascending branch will represent prebuckling deformations. The one on the second
ascending branch will represent buckled configurations. Under some conditions, the two
states can co-exist in the same structure."
Following Kyriakides' paper [18.91], the examples presented there will now be summa-
rized to provide insight into the phenomenon of propagating instabilities and familiarize the
reader with its breadth of relevance. The problems studied by Kyriakides, as well as in other
pertinent investigations, were treated as being time independent. The discussion there is
limited to the conditions of initiation and quasi-static propagation of the instabilities. In
essence, however, these present dynamic phenomena, and hence it is most appropriate to
address them within the framework of this chapter.

18.3.1 Propagation of Bulges in Inflated Elastic Tubes


Kyriakides [18.91] indicated that this example of the inflated tube vividly illustrates the
initiation and subsequent propagation of propagating instabilities. Consequently, he used it

Figure 18.57 Local load-deformation response characteristic of


Deformation structures exhibiting propagating instabilities
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1526 Buckling under Dynamic Loads and Special Problems

10 identify the char:tcteristics of this class of problem~. describe the experimental and ana-
lytical demands to study this phenomena and establish the rc lcvam terminology.
The events that take place during infl ating a section of a tube made from a rubber com-
pound that can undergo finite elastic deformations arc depicted in Figure 18.58. The tube
first expands uniformly. in both the radial and axial directions. The effect of this unifom1
increase in deformation is limited and occurs under increa~ing pressure. At time 11 (see Figure
18.58) the pressure reaches a local maximum and the cylindrical configuration becomes
unstable. The pressure drops and somewhere along the leng th of the tube a bu lge. a few
tube diameters long, appears. The following events strongly depend o n the method used to
infl ate the tube. Consideri ng a fluid volume-controll ed pressurization and assuming that the
internal volume is increased at a slow, steady state rate. fol lowing an initial transient the
bulge grows both radially and axially in a controlled fashion while the pre~sure in the tube
continues to decrease. When the diameter of the bulge reache~ a critical value at time 12 • its
radial expansion stops and it start~ spreading axially. By ~election of the rate at which fluid
i~ supplied into the tube, the ax ial growth can occur under Meady-state conditions in a quasi-
static fashion, at a propagation pressure P" much lower than that required to initiate the
bulge. P,. This pro pagation phase is characterized by the coexistence of inllatcd and unin-
flated configurations of the tube. The length of the inflated ~cction grows under the pressure
Pr until the whole tube is inflated at r,. Beyond this point inflation results again in uniform.
cylindrical expansion of the tube in a Mahle fashion under increasing pres~ure.
Under prescribed pressure, the response of the tube between r, and '~ is un~table. Thus.
when the tube is connected to a constalll pressure source, the bulge will dynamically spread

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
axial -wise. Under the conditio ns described. the interna l fl ow of the Auid supplied wil l
strongly affect the rate of expansion of the hulge.
The above characteristics of the propagating bu lge were observed in two test programs
conducted by Kyriakides and Chang: ( I ) tests on tubes inflated under volume-controlled
conditions to capture the details of the initiation of the proces~ [ 18.92) and (2) tests in which
the tubes were inflated with compressed air 10 mea5ure tltc propagation pressure of the bulge
under steady-state quasi-static conditions r18.93].

a. Volume-Controlled Inflation Experiments


The setup used in these experiments is depicted in Figure 18.59. ln this apparatus one end
of the tube was clamped to a rigid manifold while the other wa~ free. Axial load was
introduced by connecting the free end to a weight. The tube and all connected to it were

Figure 18.58 Schemat ic of pressure history in an elastic tube subj~ctcd to volu me conti'OIIcd inflation
(from I I 8.93])

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Propagating Buckles 1527

upport Frome

..
Figure 18.59 Ky1iakides and Chang's apparatus for inflaling lubes under volume conlrol (from [ 18.92])

filled with water. The effect of the weight of the water inside the tube was negated by
immersing the tube in a transparent bat h of water. (For further detai ls on the test setup see
[1 8.91) and [ 18.92)).
A typical result obtained in the tests of tubes infl ated in presence of an axial load Fl JJA
(where JL = in iti al shear modulus and A = initial cross-sectional area of tube) is shown in
Fig. 18.60. This figure consists of a sequence of tube configurations obtained from video
recordi ng and a corresponding plot of the recorded pressure-time histOry (the pressure cor-
responding to each tube configuration is marked on the pressure-time record). It is apparent
from this figure that at the early stages of the inflation process the ini tial stiffness of the
tube was high and the pressure rose sharply wi th ti me. The stiffness of the tube decreased
with increase in its deformation. and eventually the pressure reached a maximum value. Very
close to reaching the maximum pressure a bulge developed. The appearance of the bulge

(<1]>) -----
0~-----------------------------------------,
0

---
<i> _ _ __

®
--·---

Figure 18.60 Pressure-lime his10ry recorded in Kyriakidcs and Chang's 1es1s on propaga1ion of insla-
bililies in an inflated elaslic tube. and sequence of corresponding tube configurations
(from [ 18.92])

Copyright Wiley
--`,`,`````,`,`

Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1528 Buckling under Dynamic Loads and Special Problems

was associated with a drop in pressure in the fonn of a jump (indicated by a dashed line in
Figure 18.60), but then the deformation process returned to a controlled quasi-static rate of
growth.
On the basis of the numerical results obtai ned by the analysis presented within the frame-
work of the investigation (in [ 18.92]). Kyriakides and Chang explained the occurrence of
the jump as follow~:
The downturn in pressure caused unloading in most of the tube. except in the bulged ;cction. which
continued to experience an increase in deformation. As a result of the unloading. the volume of
the cylindrical secti ons was reduced. Ry contrast, the volume in the bulging ;cction increased.
Since the instamaneous volume of nuid in,ide the test specimen was prescribed, for a shorter tube
the volume increaw in the bulge could be larger than the volume decrease in the cylindrical ends.
and the postlirnit load response could progress in a controlled fashion; i.e. without "jumps" in
deformation. If the tube were longer. w that the opposite "'ould be true. the initial unloadmg would
occur dynanucally a\ the structure "snaps" to a configuratoon that has the same internal volume.
Consequently, they concluded that the longer the initial length of the tube. the greater the
jump.
It appears from Figure 18.60 that with further pumping of water into the tube. the bulge
grew and at the same time the pressure dropped gradually and asymptotically :tpproached
the steady-state value. At this stage the bulge stopped expanding diametrically and starred
spreading axially ~o that the shape of the transition between the inflated and uninflated
sections was maintained and the pressure remained con;,tant (configuration I 0 in Figure
18.60) at the value of the bulge propagation pressure P,,. Once this steady-state condition
was achieved, the pumping rate was increased by a factor of 3 to accelerate the ex periment.
The bulge gradual ly spread over the whole length of the tube.
In the tests the length of the test speci mens and the applied axial load were varied. It was
observed that an increase of the axial load reduced the initiation pressure P, as well as the
propagation pressure P,. and significantly altered the shape of the bulge formed (see Figures
7 and 8 in [18.92)). The analytical predictions (in [18.92]) compared well with these obser-
vations.

b. Determination of Propagation Pressure of Bulges


In these tests (see [ 18.93]) the tube was fixed at one end to an outlet of a rigid manifold
and allowed to hang freely. The lower end was sealed and a calibrated weight was suspended
from it (see Figure 18.61). Constant pressure-compressed air was supplied to inflate the tube
at a controlled rate. Ini tiation of the bulge, the form of which depended on the total volume
of air available in the system (i.e. on the length of the tube and the internal volume of the
manifold and accessories), occurred dynamical ly. Once the bulge was initiated, air was sup·
plied at a rate that resulted in axial expansion of tlte bu lge by approximately one-half a

Figure 18.61 Kyriakidcs and Chong's setup for mcusuring prop-


agating prc.-urc bulges in e lastic tubes subjected
to prescribed axial tension (frorn [18.931)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Propagating Buckles 1529

bu lge diameter per minute. A typical sequence of pictures from initiation and steady-state
propagation of a bu lge in one of these experiments is shown in Figure 18.62.
ln the experiments the propagation pressure was measured for different values of axial
tension (see Figure 18.63). This figure reveals very good agreement with predictions, as we ll
as that the propagation pressure decreases with tension. Profi les of propagating bulges for
different values of ax ial tension shown in Figure 18.64 demonstrated that the tension caused
elongation of the profiles and reduction of the diameters of the in flated and uninA ated sec-
tions. The analysis [18.93] predicted this behavior very well (see Figures 9-11 in [18.93]).

18.3.2 Propagation of Buckles in Long Tubes and Pipes under External


Pressure
Induced dents or geometric imperfections in long tubes and pipes designed to withstand high
external pressure can lead to local collapse. in many cases at a pressure drastically lower
than that corresponding to their buckling (elastic, see Eq. (2.1 23) in Chapter 2, or plastic,
depending on their dimensions and the material they are made from). At first such a collapse
would be local in nature. However, as indicated by Kyriakides [18.91], " such a local collapse
can initiate a more global instability where, driven by external pressure, the buckled (col-

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
lapsed) section spreads (propagates) along the long structure, often at a high velocity, and
flattens it. The buckle stops only if it encounters a physical obstacle that resists the flattening
or when it reaches an area of low pressure where propagation cannot be sustained. The
phenomenon is known as a propagating buckle." The lowest pressure under whi ch buck le
propagation can be sustained is designated as the propagation pressure P, .. (In the cases
discussed in Kyriakides' article 0.1 5 < P,JP, < 0.33, where P, is the aforementioned buck-
ling load.)
Kyriakides [ 18.9 I] demonstrated the propagating buckle phenomenon and its sources by
going through some possible scenarios of its occurrence in offshore pipelines (see Figures
2 1-23 in [ 18.91]). Based on this presentation, he concluded that propagating buckles could
occur at any pressure P,. :s P :s P,., provided that: (I) the pipes or tubes were under the
action of sufficiently high external pressure and (2) the buckle ini tiated somewhere along
the pipe or tube. This implied that a buckle initiated in a constant pressure environment,

-~----------~--~----------------

Figure 18.62 A typical sequence of patterns observed in Kyriakides and Chang's tests from bulge
initiation in an elastic tube to its steady state propagat ion (from [18.93])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1530 Buckling under Dynamic Loads and Special Problems

02~
P, -PrediCtions Mo;..·• etl Construct•on
;r • E:t.Ptrlm@tlfS

I 020

0·5
,..__
00

005

0~~--~--~--~~--~
o o~ 10 ~~ zo z~ 30
- - F.joA

f iJ:Ure 18.63 lnOuence of applied tens1on oo propagation prc\,un" obsened in Kyrialude\ and Chang's
tests. and comparison "'lh prediction (from [18.9311

where P > P, .. wou ld propagate in a dynamic fashion. whereas if initiated at the propagation
pressure P,, it would propagate in a quasi -static manner. The catastrophic nature of the
induced collapse is apparent from the two examples of buc~lcs propagated at the propagation
pressure in the te~t speci mens shown in Figure 18.65.
Kyriak.ides (in Sub~ection IliA of (18.91 () indicated that a comprehensive treatise of the
propagating buckles phenomenon imolve~ the issue of inttiation of a propagating buckle,
the detennination of the lowest pressure that can sustain a propagating buckle (propagation
pressure P,.), dynamics of buckles initiated at pressure~ above P,, and method' for arresting
such a buckle (sec for example (18.94]).

a. Propagation Buckle Experiments


A standard test method (sec footnote on p. 99 of [ 18.91 J) u~ing volume- controlled pressur-
ization of a long pipe specimen was developed in the test program carried out by Kyriakides
anti Babcock at Caltcch [18.91 1 and 118.94(- (18.96]10 provide condition' under which the
buckle could be ini tiated by any pressure. but was al lowed to propagate on ly under con-
trolled, quasi-static conditions (sec ex peri mental apparatus in Figure 18.66). The test pro-
cedure in using thi~ loading device. which was used for Mudying both the dynamics and
arrest of the propag:uing buckle. wa~ a' follows. A ~ection of a pipe. more than 50 pipe
diameters long. was ~ealed at both end~ and placed in a "stiff" sealed prcs~ure vessel com-
pletely filled with water. The vessel was pressurized u~ing :1 positive displacement hydraulic
pump to capture details of the initiut ion process [ 18.9 1(. 118.96) anti [ 18.\17]. To ensure
constant internal pressure of the pipe during the collapse process. it was vented to the

-· - ----
747

Figure 18.64 PI'Or.lcs of propagating hutgc• in an inflated eta,tic tube \objected to diiTcrcnt values of
axialtcn•ion (from [18.931J
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Propagating Buckles 1531

(a)

(b)
Figure ll!.65 l>ropugat ion buckles in long tube' nnd pipes under extemal pressure: (a) X·42 line grade
>~eel. Df1- 27.7. (b) AI-6061 -T6. Off 35.7 (Figure 24 from " Propagming lnswbilities
in StruciUres" by S. Kyriakidcs in 1\d\'tlllt'e.v i11 Applied M echanics, volume 30. edited
by J.W. HUichingon and T.V. Wu . copyright 1984 by Academic Press. reproduced by
pcnnission of the publisher)

atmosphere through an umbilical. Collapse wa> initiated at one end of the specimen by
inductng a dent one diameter long. three to four diameters away from one of the end plugs.
The events observed in such an experiment arc depicted in Figure 18.67 and were as follows.
Pressuritation stan ed at time 10 by pumping water into the system at a relatively slow,
constant rate. Initially the system defomtcd elastically and was relatively stiff. As the pressure
increased, the dented region of the tube became plastic, resulting in a "softer'' response of
the tube. Eventually. at 11, a pressure maximum was achieved and the dented section col-
lapsed, lend ing to a significant drop in pressure. This often occurred in an uncontrolled
"ju mp" fashion . The extent of this pressure jump and the length of the tube that collapsed
dynami ca lly depended on the severity of the induced dent and the postbuckling sti ffness of

Figure ll!.66 Kyriakides and Babcock's expel'imental nppar:llus for detennination of the propng:uion
prc>Stll'e of long tubes (Figlll'c 25 fmm " Propagating Instabilities in Structures" by S.
Kyriakidcs in Adva11ces i11 Applied M e/'ll(lllif'<, volume 30, edited by J.W. Hutchingon
and T.V. Wu, copyright 1984 by Academic Pres~. reproduced by permission or the pub-
li,hcr)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1532 Buckling under Dynamic L.oads and Special Problems

/"""Initial Dent
EC':~:!:.:·~~~=========::·-Vent
~=======:=;
~-

~ ~
Cross Sections
p
"
~

"'
""'
a:

'· '· Time {I)


Figure 18.67 Typical pressure history observed in Kyriakides and Babcock's tests on propagating
buckles in long tubes (Figure 26 from "Propagating Instabilities in Structures" by S.
Kyriakides in Advances in Applied Mechanics, volume 30, edited by J.W. Hutchingon
and T.V. Wu, copyright 1984 by Academic Press, reproduced by permission of the pub-
lisher)

the pipe, as well as of the test apparatus. Following the pressure decrease, energy stored in
the test rig became available and contributed to the extent of the local collapse. As more
water was pumped into the vessel, the collapsed section and its neighborhood gradually
acquired the configuration of the propagating buckle front shown in Figure 18.65, and at
time t2 the pressure stabilized at the propagation pressure of the pipe. Retaining the pumping
rate at a relatively low, constant rate, the buckle gradually spread axial-wise quasi-statically,
while the pressure in the vessel remained at P1,. At comprehensive collapse (at time ! 1 ), the
deformation process became stable again and the pressure in the tank started rising.
Note that the transient part of the response between the events ! 1 and t2 corresponds to
the initiation of the propagating buckle. The local pressure maximum is the initiation pressure
P1, the value of which is strongly dependent on the shape and amplitude of the initial dent.
The extent of the transient, however, was affected as well by the stiffness of the test system
as a whole.
A set of rings cut through the profile of a buckle propagated quasi-statically at PP• illus-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

trating the deformation history of a cross .. section of the tube as it was collapsed by a prop-
agating buckle, is depicted in Figure 18.68 (tube same as in Figure 18.65). The severity of
the induced collapse is apparent from the figure. In Figure 18.68 "dog bone''-shaped cross-
sections from similar tubes that were collapsed by buckles propagated at pressures higher
than PP are shown. The severity of the induced collapse appears to increase with pressure.

b. Propagation Pressure in Presence ol Axial Tension


Babcock and Madhavan showed [18.98] that introduction of tension can lead to a reduction
of the buckling and collapse pressures of long pipes. This stems from a reduction of the
yielding of the material as a result of the induced biaxial stress state. Kyriakides and Chang
further demonstrated experimentally [ 18.99] that the presence of tension significantly reduced
the propagation pressure of such pipes. A simple model for numerical simulation of the
propagation pressure was proposed as well in this reference (see also [18.91] and [18.96]).
To perform the combined pressure-tension experiments, Kyriakides and Chang modified
Babcock and Kyriakides' loading apparatus (see Figure 18.66) by adding a specially designed
tensioning device. The combined pressure-tensioning test facility and the tensioning unit are
depicted in Figure 18.69a and b, respectively. The tensioning device consisted of a long and
narrow reaction frame that could slide into the pressure vessel. Axial tension was applied
through a single-ended hydraulic actuator located outside the pressure vessel and connected
to the pressure vessel flange. A rod that penetrated the vessel flange through low-friction
Copyright Wiley
dynamic seals transferred the load to the specimen.
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Propagating Buckles 1533

(a)

(b)
Figure 18.68 Defonnatioo panems ()fa propagaung buckle (a) deformation history of a cros<o-socuon
of a tube. (b) cross-section' of lube' collap;ed by propagating buckle> under tncreasing
pressures-P, on the lefl and progre""cly h1gher (Figure 28 from .. Propagating Ins Ia·
btblles in Structures .. by S. KyriaLide' 111 Atlwmct's i11 Applied illec!Jnnicr. volume 30.
edited by J .W. Hutcbingon and T.V. Wu. cop)ngbt 19~ by Academic Pre''· reproduced
by permission of the publisher)

Figure 18.69 Kyri:okides and Chang's prc«urc-ocnsion oest facility (from [ 18.991): (a) schemat ic of
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley combined pressure-tension lest faci lity. (b) ocnsioning device
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1534 Buckling under Dynamic Loads and Special Problems

The specimen was connected at one end to the cross-head of the load frame and at its
other end to the loading rod through a universal joint. A load cell was placed in the load
loop and a LVDT (linear voltage differential transformer) was used to monitor the overall
axial displacement (see Figure 18.69). The system had a load capacity of 10,000 lb (~45
kN) and maximum displacement of 6 in. ( ~0.15 m). Axial tension was prescribed by a
closed-loop servo-controlled system. The loading frame had a stiffness such that the energy
stored in it was at least one order of magnitude less than that in the specimen. The complete
axial loading system rested on wheels to allow easy rolling of the test section into the
pressure chamber. A dent induced at one end of the specimen initiated buckles.
A schematic of a typical set of results observed in Kyriakides and Chang's tests is shown
in Figure 18.70. Since no tension was applied until the time t 1, the events up to that time
were identical to those shown in Figure 18.67. At time t3 , and without interrupting the
pumping of water into the system, tension was first applied. This caused an immediate drop
in the pressure in the vessel. The lowest pressure recorded was found to be associated with
the completion of the tension load ramp at t4 • Again, a transient t 3-t4 in pressure response
was observed, beyond which the propagation buckle reached a new steady state at a pressure
Pn. In addition to the drop in the propagation pressure, it appears from Figure 18.70 that
application of tension led to a change in geometry of the buckle profile and resulted in an
elongation of the profile of the transition region between the collapsed and circular sections.
It was indicated by Kyriakides and Chang that the extent of the transient t3 -t5 depended on
geometry and material of the pipe as well as on the magnitude of the applied tension load.
Furthermore, they concluded that steady-state propagation could occur only after the load
moved far enough away from its initiation point (5-10 tube diameters). A numerical simu-
lation based on the simple model proposed [18.99] of initiation and propagation of a buckle
in the presence of tension is presented in Figure 18.70. It is apparent from this figure that
the numerical analysis could capture the effect of tension on the steady-state behavior. Fur-
thermore, Kyriakides and Chang reported that the simulation agreed quite well, qualitatively
and quantitatively, with the experimental observations in Figure 18.70. On the basis of the
propagation results, they drew interaction, propagation-tension, curves. These plots also com-
pared well with predictions obtained for the simplified model (see Figure 54 in [18.91] and
Figures 6-8 in [18.99]).
Dyau and Kyriakides indicated [18.97] that the stiffness of the test facility could have
significantly affected the measured response of the pipes. Furthermore, based on their in-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

vestigation [18.97] they stated that "even for geometrically intact tubes under external pres-

c::c~~----·~ --·-------
!,~"

~===
~.c:=~-T

I,G ~T

{a) { bl

Figure 18.70 Pressure history observed in Kyriakides and Chang's tests on the effect of axial tension
on the propagation pressure of long cylindrical shells (from [ 18.99]): (a) schematic of
typical history observed in experiments, (b) numerical simulation of initiation and prop-
agation of buckle in the presence of tension
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Propagating Buckles 1535

sure, collapse localizes very quickly after the onset of instability." Since, as shown before,
a propagating buckle can be initiated at any pressure P 1, < P :S P,, they concluded, "Under
prevailing conditions, the local collapse initiates a propagating buckle." To examine this
stiffness effect they developed a stiff pressure testing facility (see Figure 13 of [18.97]).

18.3.3 Propagating Buckles in Long, Confined Cylindrical Shells


In long cylindrical structures that are lined with thin liner shells to protect their inner surfaces,
conditions can develop that may lead to buckling of the liner, followed by its separation
from the structure and its collapse. Initially this collapse is local. However, if between the
collapsed layer and the outer structure hydrostatic pressure develops, this collapse can prop-
agate at a relatively low pressure and destroy large sections of the structure. In 1986 this
phenomenon was first demonstrated by Kyriakides [18.100] and named "confined propa-
gation buckle."
Kyriakides indicated [18.91] and [18.100] that the diameter-to-thickness ratio D/t of the
liner varies extremely from application to application (from 500 to 10). Therefore, "the cause
of the initial collapse and the potential of developing a confined propagating buckle also
vary with application and must be addressed individuality." Of course, this affects the as-
sumptions and idealizations made in both the experimental and analytical treatise of each
application (see detailed discussion on individual applications in [18.91] and [18.100]).
Apparently, as in the previous cases of propagating instabilities (Subsec:tions 18.3.1 and
18.3.2), the present problem consists of an initiation phase and propagation phase. Kyriakides
[18.91] and [18.100] reviewed the literature pertinent to the initiation of a buckle (see also
Subsection 7.3.1 of Chapter 7). Following Kyriakides [18.91] and [18.100], herein the pre-
sentation will be limited to the propagation phase. Methods for determining the lowest pres-
sure at which the phenomenon of confined buckle propagation occur will be discussed.
Furthermore, Kyriakides indicated that unlike the propagating buckle discussed in Subsec-
tions 18.3.1 and 18.3.2, confined propagation can also occur for strictly elastic linear defor-
mation. Therefore, he employed two different sets of experiments, one involving linearly
elastic Mylar shells and the second using aluminum alloy and steel tubes.

a. Linearly Elastic Shells

The experimental apparatus used in the tests with the Mylar shells is shown in Figure 18.71.
In this test the Mylar tube was fitted inside a thick, transparent acrylic tube. The inside
diameter of this tube was the same as the outer diameter of the Mylar tubes. The ends of
the Mylar tubes were sealed, arranging for access to a vacuum pump and pressure-measuring
transducers (see Figure 18.71). The dimensions of the Mylar tubes were selected such that
atmospheric pressure was enough to cause collapse and propagation. By reduction of the
internal pressure in the tube with the vacuum pump external pressure was applied.
A mechanical plunger was used to initiate a buckle at one end of the tube and to separate
the liner from the cavity wall. This reduced the local collapse pressure and resulted in local
buckling of the liner, which was 8-10 diameters long (see Figure 18.72). Due to the "flex-
ible" nature of the whole system, the initial collapse occurred in a dynamic fashion and was

Needle Valve
Mylar Shell

Aery lie Tube

Figure 18.71 Kyriakides' experimental apparatus for propagation of confined buckles in linearly elastic
Mylar shells (Figure 60 from "Propagating Instabilities in Structures" by S. Kyriakides
in Advances in Applied Mechanics, volume 30, edited by J.W. Hutchinson and T.V. Wu,
copyright 1984 by Academic Press, reproduced by permission of the publisher)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1536 Buckling under Dynamic Loads and Special Problems

(1). ll'liiMtlon pi"OOftU.

Fi~'Urc 18.72 Initiation of buckle in a thin-walled Mylar shell and ste<tdy-state propagation of a confined
buckle in the shell observed in Kyriakides' tests: (a) initiation process. (b) steady-state
propagation (Figure 61 from ''Propagating Instabi lities in Structures·· by S. Kyriakides
in Advances in Applied Mechanics. volume 30, edited by J.W. Hutchinson and T.V. Wu.
copytight 1984 by Academic Press. reproduced by permission of the publisher)

associated with a s ignificant pressure transient. Steady-state quasi-static propagation was


achieved within 10 tube diameters from the initiation point. The rate of propagation was
controlled by using a need le valve (see Figure 18. 72) to select the rate at which the tube
was evacuated. A picture depic ti ng the propagation stage is shown in Figure 18.72.

b. Elastic-Plastic Shells
To perform these experiments, the pressures required to ini tiate and propagate confined prop-
agating buckles were relatively high (see d iscussion in [ 18.91] and (18.100]). Hence. the
apparatus of Figure 18.71 could not be employed and an alternative experime ntal scheme
was proposed. The ci rc ular confinement was obtained by molding a thick. concentric cylinder

Steel Motd Plug


r Plaster of Pons

~~~~Venr Test Spectmen


Concenrnc Spocer

Figure 18.73 Assembly of metal shell specimen and confinement in Kyriakides' tests on propagating
buckles in long confined elastic-plastic cylindrical shells (from [ 18.100])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Propagating B uckles 1537

? : i
L! !! ? § 1 ::m:uu:d
j;f,£2 :::;:
I
= ~
A
!£ I~
• ftc~
j I I
£III\ I::
~:LI fP, I I I

.. '·'· '• Time {l}

Figure 18.74 Typical pressure-time history observed in Kyriakidcs· quasi-static confined propagation
experiments (Figure 63 from "Propagating Instabilities in Structures·· by S. Kyriakides
in Admncts in AiJ/Jii~d M~clwnics. volume 30. cdiled by J.W. Hutchinson and T.V. Wu.
cop)"right 1984 by Academic Press. reproduced b) permission of the publi; her)

of plaster of paris around the tube. To reduce friction between the two surfaces, the ex terior
of the tubes were well lubricated prior to casting the pla>ter around them. A split outer steel
tube was used as the mold . Ring spacers inserted at each end of the steel tube maintained
the concentricity of the assembly (sec Figure 18.73).
The objective of the investigation was to demonstrate the phenomenon and establish the
confined propagation pressure. No attempt was made to simulate the initiation process.
Hence. a local buckle was initiated by physically denting it on a section of tube left outside
the confinement (see Figure 18.73).
To ensure quasi-static propagation conditions the experiments were perfonned under vol-
ume-controlled pressurization (690 har) using a pres>ure vessel and the experimental setup
shown in Figure 18.66. The inside of the test specimen was vented to the atmosphere to
ensure that the same reference level of the interna l pressure was maintained during the
collapse process.
A typical schematic of pressure history recorded by the pressure transducer during such
an experiment is presented in Figure 18.74. The collapse process initiated from the dent on
the unconfined section under pressure P, at time 11 • The process was completed at time t 2•
Sub<;equently. the buckle propagated at the propagation pressure PP of the aforementioned
dog bone collapse mode.
Once the buckle reached the confined section at 1, it stopped propagating. Continued
pumping of water into the vessel resulted in a relatively sharp rise in pressure. Apparently
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 11!.75 Typical profile or a confined propagating buckle observed in Kyriakides' confined prop-
agation experiment' (from [ 18.100))
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1538 Buckling under Dynamic Loads and Special Problems

Fi~urc 1!1.76 Sections through profile of confined propagation buckle observed in Kyriakides ' tests
(from r18.1001)

the rise was not instantaneous (because a finite volume of water was required to lla1ten
funher the already collapsed section of the tube). The confined tube remained undisturbed
until time '•· when the llauened section of the tube adjacent to the edge of the confinement
snapped in a ··u" shape. This allowed the huckle to ~tan penetrating the confinement. The
pressure at which this occurred. P1c. was the highest experienced during the experiment. By
time 1,, the profile of confined propagation was fully developed (typically five tube diameters
from the edge of the confinement). The pressure stabililed at a new plateau as the buckle
propagated at a steady state. The propagation of the buckle continued at this pressure plateau,
P,.,. until the whole lengt h of the shel l was collapsed.
The collapse process was visually observed in a separate experiment [ 18.9 11 and ( 18.100].
A typical profile of a confined propagating buckle is shown in Figure 18.75. A sequence of
configurations that a cross section underwent during the collapse process is shown in Figure
18.76.

18.3.4 Buckle Propagation in Long Shallow Panels


The response of long clastic panels of shallow arch cross-sections to uniform pres~ure loading
has the nonli neari ty and instabil ity characteristics of shallow arches (see Subsecti on 2. 1.10
and 7.2 of Chapters 2 and 7. respectively). Similar 10 the arches, these long panel structures
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

show locally the up-down-up load deformation response (sec Figure 18.77) common to struc-
lures exhibiting propagating instabilities (discussed in the preceding Subsections 18.3.1-
18.3.3). thus meeting the charactcrislics necessary for propagation instability to develop.
Following the occurrence of the first instability. the deformation ceases to be axially uni-
form and localizes to a region on the panel surface. Thi; localized collapse is accompanied
by a drop in pressure (as would occur in volume-controlled pressurization) and causes un-
loading in the remainder of the panel. Subsequem deformation is confi ned then to the above
loca l region unt il the local collapse is arrested by membrane tension. Following this and

{
~;

"("-...... .,~
• l:t
-- .
.....

'· ·-·-·-·-· d

'" "'
Figure 18.77 Pressure volume rc;ponse of arches ([rom 118.101]): (a) symmetric buckling behavior,
(b) aoymmctric buckling behavior ror part or the postbuckliug history
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Propagating Buckles 1539

..
figure 18.78 Kyriakides and Ar<.eculerame·, expenmemal <c1Up for dcmon\trating propagating hucklc'
m long <hallow panel< (from [18.101]1: (a) <ebemauc of experimental apparaiU<, !hi
cro<~·<.eelion of experimental apparatu'

under prcscril>ed constant loading. the collapse can spread along the length of 1he panel al
1he propagation pre~surc and collapse 1hc ''hole slruciUrc. very much like a propagation
bucl..lc m a long ptpehne, discussed in Subsec1ion IIU.2 of this chap1er. This propagation
pressure i~ sufficiently lower than 1he clas-.cal cntical one associated wi1h 1he pcrfec1 struc·
lure.
Kyriakidcs et al.(l8.91), [18.101] and [1 8. 102( developed an analysis for numerical sim-
ulalion of buckle propagation in long shallow panels. Resuhs of an experimenlal program
carried oul to demonstrate the initiation and slcady· \ laic propagation of such buckles were
also prcsemed in lhese studies. !1. schemat ic of lhc simple te!>ls apparatus used by Kyriakidc;
el al. is prcsemed in Figure 18.78. In this device long shallow panels were formed by bent
strips of 1hin Mylar into the cross-scclion shown m Figure 17.78. The dimem.ions of 1he
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

panels were chosen such that snap-through bucl.hng would occur at external pre"ure' only
a small fraction of an atmosphere. The edge' of 1hc M) lar \lrips were bonded imo obhquc
slils cui imo 1wo long metal end-suppon beams lthe radtu~ and enclosed angle of the arch
could be 'aricd b) u~ing slits with different oricnt.ation.,). thu~ providing nominal clamped-
end condtllon... The.,e end beams were mounted on a U-shaped aluminum suppon \tntcturc.
The ends of the 10-ft ( - 3-m) long 1ube-hkc Slntclurc were closed with flanges and provi~ton.,

--...---...,..,): I I
•• I

~.r·----r-- .- c, - 1
~.c----®· --- .... -...........~
~e------x--- ·-- - -- --- ------ ·---·-- ·a
Fi~:urc t8.79 Schcmauc of typical pressure-lime hi,hlry ol u pmpagaling buckle in a long ' halk>w
panel ob>er"ed in Kyriakides and Power\ cxperimenls (from [ 18. 102))
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1540 Buckling under Dynamic Loads and Special Problems

(a) (b) (C)

Figure 18.80 Def01med panel observed in Kyriakides and Power's tests (from [ 18.1 02]): (a) onset of
asymmetric buckling. (b) symmeltic buckling associated with dynamic collapse after
onset of buckling. (c) transition region between collapsed and uncollapsed port ions

were made for the inside to be hermetically closed. Grid lines were drawn on the panel
surface before installing it to augment visualization of the buckling process. External pressure
loading was applied by evacuating the air inside the closed tube with a vacuum pump.
A typical pressure-time history recorded during a test of a panel exhibiting unsymmetric
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

buckling (see Figure 18.77), and some of the major features observed during such an ex-
periment are depicted schematically in Figure 18.79. Loading started at time 10 and the
structure deformed unifonnally with a relatively stiff response. The first observed instabili ty
was unsymmetric (see Figure 18.80a). Immediately after this onset of unsymmetric defor-
mation, the panel collapsed dynamically (at time 11) and developed a local symmetric buckle
about the midspan that extended over a length of about seven to eight times the arch span
(see Figure 18.80b). Under the loadi ng conditions, this snap-through buckling was associ ated
with a significant pressure transient. Once th e local buckle formed, its geometry cou ld be
"frozen" by intcrntpting the evacuation of air (after t2 ). When the local buckle became fully
developed, further propagation of the buckle occurred at a well-defined pressure, the prop-
agation pressure P,.. which was significantly lower than the pressure observed at the initiation
of local collapse. The buckle propagated in both di.rections until it reached both ends of the
tube and the whole panel was collapsed. The rate at which the system was evacuated con-
trolled the rate of the propagation of the buckle. The transition region between the buckled
and unbuckled sections observed in the tests is shown in Figure 18.80c. O nce the buckle
reached both ends at time t •• further deformation required a signi ficant increase in applied
pressure and the deformation was uniform along the structure.
The test resu lts and the numerical simul ations presented [ 18.1 02] demonstrated that clas-
sical instabilities could trigger localization of collapse and the ini tiation of propagating buck-
les in long clastic shallow panels that exhibit a local up-down-up load- de formation resp(mse.
On the basis or the observed theoretical and experimental results, it was concluded that long
panels of other arch geometries, boundary conditions and load distributions could be antic-
ipated to show si milar local ized and propagating instabilities.

References
18. 1 Koning. V.C. and Taub, J., Stossartige Knickbeanspruchung schlankcr Sttibc im clastischcn
Bereich bei beiderseits gelenk.iger Lagenmg. Lu.fifahriforsclumg. 10. (2). July 1933, 55- 64.
translated as Impact Buckling of Thin Bars in the Elastic Range Hinged at Both Ends. NASA
TM 748, 1934.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1541

18.2 G. Herrmann, ed., Dynamic Stability of Structures, Proceedings of International Conference,


Northwestern University, Evanston, Ill., Oct. 18-20, 1965, Pergamon Press, New York, 1967.
18.3 Bolotin, V.V., The Dynamic Stability of Elastic Systems, Holden-Day, San Francisco, 1964.
18.4 Lindberg, H.E. and Florence, A.L., Dynamic Pulse Buckling, Martinus Nijhoff, Dortrecht, 1987.
18.5 Simitses, G.J., Dynamic Stability of Suddenly Loaded Structures, Springer-Verlag, New York,
1990.
18.6 Ari-Gur, J., Weller, T. and Singer, J., Experimental Studies of Columns under Axial Impact,
TAE Report No. 346, Aerospace Engineering, Technion-Israel Institute of Technology, Haifa,
Israel, Dec. 1978.
18.7 Singer, J. and Ari-Gur, J., Dynamic Buckling of Thin-Walled Structures under Impact, Jahres-
tagung der DGLR, Aachen, May 11-14, 1981, DGLR-Vortrag No. 81-007, published in the
DGLR Jahrbuch 1981, vol. I.
18.8 Evan-Iwanowski, R.M., On the Parametric Response of Structures, Applied Mechanics Review,
18, (9), 1965, 699-702.
18.9 Johnson, W. and Mamalis, A.G., Crashworthiness of Vehicles, Mechanical Engineering
Publications, London, 1978.
18.10 Lowe, W.T., Al-Hassani, S.T.S. and Johnson, W., Impact Behaviour of Small Scale Model Motor
Coaches, Proceedings Institute of Mechanical Engineers (Auto Division), 186, 1972, 409-419.
18.11 Vaughan, V.A. and Alfara-Bon, E., Impact Dynamics Research Facility for Full Scale Aircraft
Crash Testing, NASA TN D-8179, 1976.
18.12 Johnson, W. and Reid, S.R., Metallic Energy Dissipating Systems, Applied Mechanics Reviews,
31, (3), March 1978, 277-287.
18.13 Ezra, A.A. and Fay, R.J., An Assessment of Energy Absorbing Devices for Prospective Use in
Aircraft Impact Situations, Dynamic Response of Structures, G. Hermann and N. Perrone, eds.,
Pergamon Press, Oxford, 1972, 225-246.
18.14 Jones, N., Recent Progress in the Dynamic Plastic Behavior of Structures-Part I, Shock and
Vibration Digest, 10, (9), Sept. 1978, 21-33.
18.15 Jones, N., Response of Structures to Dynamic Loading, in: Conference on the Mechanical
Properties of Materials at High Rates of Strain, Oxford, March 1979, Institute of Physics,
Conference Series No. 47, chap. 3, 1979, 254-276.
18.16 Jones, N., Dynamic Plastic Response of Structures, in: Recent Advances in Structural Dynamics,
M. Petyt, ed., Institute of Sound and Vibration Research, University of Southampton, England,
1980, 677-689.
18.17 Ari-Gur, J., Weller, T. and Singer, J., Theoretical Studies of Columns under Axial Impact and
Experimental Verification, TAE Report No. 377, Department of Aeronautical Engineering, Tech-
nion-Israel Institute of Technology, Haifa, Israel, August 1979.
18.18 Ari-Gur, J. and Singer, J., Composite Material Columns under Axial Impact, TAE Report No.
462, Department of Aeronautical Engineering, Technion-Israel Institute of Technology, Haifa,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Israel, Dec. 1981.


18.19 Ari-Gur, J., Weller, T. and Singer, J., Experimental and Theoretical Studies of Columns under
Axial Impact, International Journal of Solids and Structures, 18, (7), 1982, 619-641.
18.20 Singer, J., Weller, T., Libai, A. and Ari-Gur, J., Preliminary Dynamic Stability Studies Using
Vibration Techniques for Definition of Boundaries, Internal Report ASL-95, Department of
Aeronautical Engineering, Technion-Israel Institute of Technology, Haifa, Israel, Jan. 1978.
18.21 Singer, J., Weller, T., Ari-Gur, J. and Libai, A., Preliminary Dynamic Stability Studies Using
Vibration Techniques for Definition of Boundaries (cont.), Internal Report ASL-100, Depart-
ment of Aeronautical Engineering, Technion-Israel Institute of Technology, Haifa, Israel, Au-
gust 1978.
18.22 Ziegler, H., Principles of Structural Stability, Blaisdell, Waltham, Mass., 1968, 8.
18.23 Budiansky, B. and Hutchinson, J.W., Dynamic Buckling of Imperfection Sensitive Structures,
in: Proceedings of the lith International Congress of Applied Mechanics, 1964, H. Gi:itler, ed.,
Springer-Verlag, 1966, Berlin, 636-651.
18.24 Budiansky, B., Dynamic Buckling of Elastic Structures: Criteria and Estimates, in: Proceedings,
International Conference on Dynamic Stability of Structures, G. Herrmann, ed., Pergamon Press,
Oxford, 1967, 83-106.
18.25 Hutchinson, J. and Budiansky, B., Dynamic Buckling Estimates, AIAA Journal, 4, 1966, 525-
530.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1542 Buckling under Dynamic L.oads and Special Problems

18.26 Hoff. N.J .. Dynamic Stability of Structures, in: Proceedings, International Conference on Dy-
namic Stability of Structures. Northwestern University, Oct. 1965, G. Hen·mann, ed., Pergamon
Press, Oxford, 1967, 7-41.
18.27 Maymon, G. and Libai, A., Dynamics and Failure of Cylindrical Shells Subjected to Axial
Impact, AIAA Journal, 15, (11), Nov. 1977, 1624-1630.
18.28 Humphreys, J.S. and Bodner, S.R., Dynamic Buckling of Shallow Shells under Impulsive Load-
ing, Journal of the Engineering Mechanics Division, Proceedings of the American Society of
Civil Engineers, 88, (EM2), April 1962, 17-36.
18.29 Goodier, J.N., Dynamic Plastic Buckling, in: Proceedings, International Conference on Dy-
namic Stability of Structures, Northwestern University, Oct. 1965, G. Herrmann, ed., Pergamon
Press, Oxford, 1967, 189-211.
18.30 Ari-Gur, J., Singer, J. and Weller, T, Dynamic Buckling of Plates under Longitudinal Impact,
TAE Report No. 430, Department of Aeronautical Engineering, Technion-Israel Institute of
Technology, August, 1981; also Proceedings, 23d Israel Annual Conference on Aviation and
Astronautics, Feb. 11-12, 1981, Israel Journal of Technology, 19, 1981,57--64
18.31 Hayashi, T. and Sano, Y., Dynamic Buckling of Elastic Bars, lst Report, The Case of Low
Velocity Impact, Bulletin of the JSME, 15, (88), 1972, 1167-1175.
18.32 Hayashi, T. and Sano, Y.. Dynamic Buckling of Elastic Bars, 2nd Report, The Case of High
Velocity Impact, Bulletin of the JSME, 15, (88), 1972. 1176-1184.
18.33 Gerard, G. and Becker, H., Column Behavior under Conditions of Impact, Journal of Aeronau-
tical Sciences, 19, (1), Jan. 1952, 58-·60, 65.
18.34 Meier, J.H., On the Dynamics of Elastic Buckling, Journal of the Aeronautical Sciences, 12,
(4), Oct. 1945, 433-440.
18.35 Hoff. N.J., The Dynamics of the Buckling of Elastic Columns, ASME Transactions, Journal of
Applied Mechanics, 18, (I), March 1951, 68-74.
18.36 Lindberg, H.E., Buckling of a Very Thin Cylindrical Shell Due to an Impulsive Pressure, ASME
Transactions 86, Series E, Journal of Applied Mechanics, 31, June 1964, 267-272.
18.37 Lindberg, H.E., Impact Buckling of a Thin Bar, ASME Transactions, Journal r~f Applied Me-
chanics, 32, (2), June 1965, 315-322.
18.38 Ari-Gur. J., Singer, J. and Riihrle, H., The Behavior of Composite Thin-Walled Structures in
Dynamic Buckling under Impact, in: Proceedings. I 3th Congress of the International Council
of the Aeronautical Sciences and AIAA, Aircraft System and Technology Conference, Seattle,
Wash., August 22-27, 1982, 1001-1010.
18.39 Davidson, J.F., Buckling of Struts under Dynamic Loading, Journal of Mechanics and Physics
of Solids, 2, 1953, 54-66.
18.40 Erickson, B., Nardo, S.V., Patel. S.A. and Hoff, N.J., An Experimental Investigation of the
Maximum Loads Supported by Elasttc Columns in Rapid Compression Tests, Proceedings of
the Society of Experimental Stress Analysis, 14, ( l ). 1956, 13-20.
18.41 Hegglin . B., Dynamic Buckling of Columns, SUDAER Report No. 129, Stanford University,
Stanford, Calif., June 1962.
18.42 Hoff, N.J., Nardo, S.V. and Erickson, B., The Maximum Load Supported by an Elastic Column
in a Rapid Compression Test, in: Proceedings of the First U.S. National Congress of Applied
Mechanics, ASME, 1952, 419-430.
18.43 Zhang, Q., Li, S. and Zheng, J., Dynamic Response, Buckling and Collapsing of Elastic-Plastic
Straight Columns under Axial Solid-Fluid Slamming Compression-!. Experiments, Interna-
tional Journal of Solids and Structures, 29, (3), 1992, 381-397.
18.44 Ari-Gur, J. and Weller, T., Experimental Studies with Metal Plates Subjected to Inplane Axial
Impact, TAE Report No. 580, Department of Aerospace Engineering, Technion-Israel Institute
of Technology, Haifa, Israel, August 1985.
18.45 Weller, T., Abramovich, H., Nachmani, S. and Grunwald, A., Beulen Von FVW-Platten, Internal
Report ASL-137, Department of Aeronautical Engineering, Technion-Israel Institute of Tech-
nology, Haifa, Israel, Jan. 1987.
18.46 Abramovich, H. and Grunwald, A., Stability of Axially Impacted Composite Plates, Composite
Structures, 32, 1995, 151-158.
18.47 Goodier, J.N., Dynamic Buckling of Rectangular Plates in Sustained Plastic Compressive Flow,
in: Engineering Plasticity, Proceeding; of an International Conference on Plasticity, J. Heyman,
--`,`,`````,`,````,,``,``,,`-`-`,,

and F.A. Leckie, eds., Cambridge l.Jniversity, England, March 1968, Cambridge University
Press, 1968, 183-200.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1543

18.48 Hoff, N.J. and Bruce, V.G., Dynamic Analysis of the Buckling of Laterally Loaded Flat Arches,
Journal of Mathematics and Physics, 32, (4), 1954, 276~288.
18.49 Humphreys, J.S., On Dynamic Snap Buckling of Shallow Arches, AIAA Journal, 4, (5), May
1966, 878~886.
18.50 Cheung, M.C. and Babcock, C.D., Jr., Experimental Investigations of Impulsively Loaded
Clamped Circular Arch, Report SM 68-18, Graduate Aeronautical Laboratories, California In-
stitute of Technology, Pasadena, Calif., Dec. 1968.
18.51 Vahidi, B., Non-Existence of Snap-Through for Clamped Shallow Elastic Arches Subjected to
Impulsive Load, TR No. 8, Department of Applied Mechanics, University of California, San
Diego, March 1968.
18.52 Vahidi, B., Some Aspects of Dynamic Snap-Through Problems, Ph.D. dissertation, Department
of Applied Mechanics, University of California, San Diego, 1969.
18.53 Suhara, J., Snapping of Shallow Spherical Shells under Static and Dynamic Loading, Report
ASRL TR 76-4, Aeroelastic and Structures Research Laboratory, MIT, Cambridge, Mass., June
1960.
18.54 Budiansky, B. and Roth, R.S., Axisymmetric Dynamic Buckling of Clamped Shallow Spherical
Shells, in: Collected Papers on Instability of Shell Structures, NASA TN D-151 0, Washington,
D.C., 1962, 597~606.
18.55 Simitses, J. and Tabeie, A., Dynamic Stability of Shell Structures, in: Proceedings, AIAAI
ASMEI ASCEI AHSI ASC, 38th Structures, Structural Dynamic and Materials Conference and
Exhibit, and AIAA I ASMEI AHC Adaptive Structures Forum, Kissimmee, Fla., April 7 ~ 10, 1997,
2438~2445.
18.56 Svalbonas, V. and Kalnins, A., Dynamic Buckling of Shells: Evaluation of Various Methods,
Nuclear Engineering and Design, 44, 1977, 331~356.
18.57 Humphreys, R.S., Roth, R.S. and Zatlers, J., Experiments on Dynamic Buckling of Shallow
Spherical Shells under Shock Loading, AIAA Journal, 3, (1), Jan. 1965, 33~39.
18.58 Huang, N.C., Axisymmetric Dynamic Snap-Through of Elastic Clamped Shallow Spherical
Shells, AIAA Journal, 7, (2), May 1969, 215~220.
18.59 Lock, M.H., Okubo, S. and Whittier, J.S., Experiments on the Snapping of a Shallow Dome
under a Step Pressure Load, AIAA Journal, 6, (7), July 1968, 1320~1326.
18.60 Lock, M.H., Snapping of a Shallow Sinusoidal Arch under a Step Pressure Load, AIAA Journal,
14, (7), July 1966, 1249~ 1256.
18.61 Tawadros, K.Z. and Glockner, P.G., Experiments on the Non-Linear Dynamic Response of
Shells under Blast Waves, Journal of Sound and Vibration, 26, (4), 1973, pp. 441~463.
18.62 Witmer, E.A., Pian, T.H.H. and Balmer, H.A., Dynamic Deformation and Buckling of Spherical
Shells under Blast and Impact Loading, in: Collected Papers on Instability of Shell Structures,
NASA TN D-1510, 1962, 607~622.
18.63 Burns, J.J., Experimental Buckling of Thin Shells of Revolution, Journal of the Engineering
Mechanics Division, Proceedings of the American Society of Civil Engineers, 90, (EM3), June
1964, 171~187.
18.64 Liu, T.L. and Babcock, C.D., Jr., An Energy Approach to the Dynamic Buckling of Spherical
Caps, AFOSR Scientific Report, AFOSR TR-71-1 076, Graduate Aeronautical Laboratories Cal-
ifornia Institute of Technology, Pasadena, Calif., April 1971.
18.65 Volmir, A.S., On the Stability of Dynamically Loaded Cylindrical Shells, Docl. Acad. Nauk
SSSR, 123, (5), 1958, 806~808 (in Russian); translated in Soviet Physics Doclady, 3, (6), Nov.~
Dec. 1958, 1287~1289.
18.66 Jones, N., Dynamic Elastic and Inelastic Buckling of Shells, in: Developments in Thin- Walled
Structures, vol. 2, J. Rhodes and A.C. Walker, eds., Elsevier Applied Science, Amsterdam,
1984, 49~91.
18.67 Dietz, W.K., On the Dynamic Stability of Eccentrically Reinforced Circular Cylindrical Shells,
Technical Report SURI No. 1620.1245-60, Applied Mechanics Laboratory, Department of Me-
chanical and Aerospace Engineering, Syracuse University Research Institute, 1967.
18.68 Maymon, G. and Singer, J., Dynamic Elastic Buckling of Stringer-Stiffened Cylindrical Shells
under Axial Impact, Israel Journal of Technology, 9, (6), 1971, 595~606.
18.69 Lakshmikantham, C. and Tsui, T.U., Dynamic Stability of Axially-Stiffened Imperfect Cylin-
drical Shells under Axial Step Loading, AIAA Journal, 12, (2), Feb., 1974, 163~169.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
18.70 Lakshmikantham, C. and Tsui, T.U., Dynamic Buckling of Ring Stiffened Cylindrical Shells,
AIAA Journal, 13, (9), Sept. 1975, 1165~1170.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1544 Buckling under Dynamic L.oads and Special Problems

18.71 Jones, N. and Papageorgiou, E.A., Dynamic Axial Plastic Buckling of Stringer Stiffened Cy-
lindrical Shells, International Journal of Mechanical Sciences, 24, (1), 1982, 1-20.
18.72 Wood, J.D. and Koval, L.R., Buckling of Cylindrical Shells under Dynamic Loads, AIAA Jour-
nal, 1, (11), Nov. 1963, 2576-2582; also Final Report on Buckling of Shells under Dynamic
Loads, Report 8622-0001-RU-000, Engineering Mechanics Laboratory EM 11-22, Space Tech-
nology Laboratories, Los Angeles, Oct. 1961.
18.73 Abrahamson, G,R. and Goodier, J.N., Dynamic Plastic Flow Buckling of a Cylindrical Shell
from Uniform Radial Impulse, Proceedings, Fourth U.S. National Congress of Applied Me-
chanics, Berkeley, Calif., June 1962. 939-950.
18.74 Anderson, D.L. and Lindberg, H.E., Dynamic Pulse Buckling of Cylindrical Shells under Tran-
sient Lateral Pressures, AIAA Journal, 6, (4), April 1968, 589-598.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
18.75 Florence, A.L., Buckling of Viscoplastic Cylindrical Shells Due to Impulsive Loading, AIAA
Journal, 6, (3), March 1968, 532-537.
18.76 Florence, A.L. and Vaughan, H., Dynamic Plastic Flow Buckling of Short Cylindrical Shells
Due to Impulsive Loading, International Journal of Solids and Structures, 4, 1968, 741-756.
18.77 Vaughan, H. and Florence, A.L., Plastic Flow Buckling of Cylindrical Shells Due to Impulsive
Loading, Transactions of ASME, vol. 92, series E, Journal of Applied Mechanics, 37, (1), March
1970, 171-179.
18.78 Lindberg, H.E., Dynamic Plastic Buckling of a Thin Cylindrical Shell Containing an Elastic
Core, Journal of Applied Mechanics, 32, (4), Dec. 1965, 803-812.
18.79 Vaughan, H. and Lindberg, H.E., Dynamic Plastic Buckling of Sandwich Shells, Transactions
of ASME, vol. 90, series E, Journal of Applied Mechanics, 35, (3), September 1968, 539-546.
18.80 Lindberg, H.E., Stress Amplification [n a Ring Caused by Dynamic Instability, Transactions of
ASME, vol. 96, Series E, Journal of Applied Mechanics, 41, (2), June 1974, 392-400.
18.81 Kirkpatrick, S.W. and Holmes, B.S., Structural Response of Thin Cylindrical Shells Subjected
to Impulsive External Loads, AIAA Journal, 26, (1), Jan. 1988, 96-103.
18.82 Kirkpatrick, S.W. and Holmes, B.S., Effect of Initial Imperfections on Dynamic Buckling of
Shells, ASCE Journal of Engineering Mechanics, 115, (5), 1989, 1975-1093.
18.83 Hallquist, J.D. and Benson, D.J., DYNA3D Users Manual (Nonlinear Dynamic Analysis of
Structures in Three Dimensions), Lawrence Livermore National Laboratory, Livermore, Calif.,
Report UClD-19592, Revision 2, March 1986.
18.84 Lindberg, H.E., Dynamic Pulse Buckling of Imperfection Sensitive Shells, ASME, Recent Ad-
vances in Impact Dynamic of Engineering Structures, AMD, 105, 1989, 97-103.
18.85 Coppa, A.P. and Nash, W.A. Dynamic Buckling of Shell Structures Subject to Longitudinal
Impact, General Electric Company Report ASD-TDR-62-774, Dec. 1962.
18.86 Coppa, A.P., The Buckling of Circular Cylindrical Shells Subject to Axial Impact. in: Collected
Papers on Instability of Shell Structures, NASA TN D-151 0, Dec. 1962, 361-400.
18.87 Coppa, A.P., Effect of End Conditions on Buckling of Cylindrical Shells under Axial Com-
pression Impact, in: Test Methods for Compression Members, ASTM STP 419, ASTM, Phila-
delphia, 1967, 115-136.
18.88 Lindberg, H.E., Rubin, M.B. and Schwer, L.H., Dynamic Buckling of Cylindrical Shells from
Oscillating Waves Following Axial Impact, International Journal of Solids and Structures, 23,
(6), 1987, 669-692.
18.89 Tulk, J.D., Buckling of Circular Cylindrical Shells under Dynamically Applied Axial Loads,
UTIAS Report No. 160, Institute of Aerospace Studies, University of Toronto, June 1972.
18.90 Tennyson, R.C., Interaction of Cylindrical Shell Buckling Experiments with Theory, in: Theory
of Shells, W.T. Koiter and G.K. Mikhailov, eds., North- Holland, Amsterdam, 1980, 65-116.
18.91 Kyriakides, S., Propagating Instabilities in Structures, in: Advances in Applied Mechanics, J.W.
Hutchinson and T.Y. Wu, eds., vol. 30, Academic Press, San Diego, Calif., 1984, 67-189.
18.92 Kyriakides, S. and Chang, Y.C., The Initiation and Propagation of a Localized Instability in an
Inflated Elastic Tube, International Journal of Solids and Structures, 27, (9), 1991, 1085-1111.
18.93 Kyriakides, S. and Chang, Y.C., On the Inflation of a Long Elastic Tube in the Presence of
Axial Load, International Journal of Solids and Structures, 26, (9/ 10), Babcock Memorial
Volume. 1990, 328-336.
18.94 Kyriakides, S. and Babcock, C.D., On the Dynamics and the AtTest of the Propagating Buckle
in Offshore Pipelines, in: Proceedings, 11th Annual Offshore Technology Conference, OTC
3479, Houston, Tex., April 30-May 3, 1979, 1035-1045.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1545

18.95 Kyriakides, S. and Babcock, C.D., Experimental Determination of the Propagation Pressure of
Circular Pipes, Transactions of the ASME, Journal of Pressure Vessel Technology. 103, Nov.
1981, 328-336.
18.96 Dyau, J.Y. and Kyriakides, S., On the Propagation Pressure of Long Cylindrical Shells under
External Pressure, International Journal of Mechanical Sciences, 35, (8), 1993, 675-713.
18.97 Dyau, J.Y. and Kyriakides, S., On the Localization of Collapse in Cylindrical Shells under
External Pressure, International Journal of' Solids and Structures, 30, (4), 1993, 463-482.
18.98 Babcock, C.D. and Madhavan, R., Pipe Collapse under Combined Axial Tension and External
Pressure, in: Factors Affecting Pipe Collapse-Phase II, EMRL Report No. 87/8, Engineering
Mechanics Research Laboratory, Department of Aerospace Engineering and Engineering Me-
chanics, The University of Texas at Austin, Austin, Tex., 1987.
18.99 Kyriakides, S. and Chang, Y.C., On the Effect of Axial Tension on the Propagation Pressure
of Long Cylindrical Shells, International Journal of Mechanical Sciences, 34, (1), 1992, 3-15.
18.100 Kyriakides, S., Propagating Buckles in Long Confined Cylindrical Shells, International Journal

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
of Solids and Structures, 22, (12), 1986, 1579-1597.
18.101 Kyriakides, S. and Arseculerante, R., Propagating Instabilities in Long Shallow Panels, Journal
of Engineering Mechanics, 119, (3), March 1993, 570-583.
18. I 02 Power, T.L and Kyriakides, S., Localization and Propagation of Instabilities in Long Shallow
Panels under External Pressure, Transactions of the ASME, Journal of Applied Mechanics, 61,
Dec. 1994, 755-763.
18.103 Zimcik, D.G., Stability of Circular Cylindrical Shells under Transient Axial Loading, UTIAS
Report 206, CN ISSN 0082-5255, Institute for Aerospace Studies, University of Toronto, March
1976.

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
19
Thermal Buckling and Creep
Buckling

19.1 Introduction
19.1.1 High-Temperature Effects in Structures
Structures are often subjected to high temperatures as a result of their operation. If the heating
is uniform and slow, it primarily affects the mechanical properties of the material, but if it
is non-uniform or rapid it may give rise to thermal stresses that may cause thermal buckling.
Structures used in mechanical or chemical engineering or in nuclear reactors may be exposed
to such high-temperature effects, but in aerospace structures heating is a direct result of high-
speed flight, aerodynamic heating. Since this aerodynamic heating is usually non-uniform
and aerospace structures are thin-walled, the high-temperature effects in these structures often
result in thermal buckling and therefore present an important loading case.
The high-temperature effects ensuing from aerodynamic heating became of prime impor-
tance with the advent of supersonic flight and atmospheric re-entry in the forties and fifties,
generating extensive research activities (see for example [19.1]-[19.9]).
The skin temperatures in the usual supersonic flight, at Mach 2-2.5, are still relatively
low, up to about 200aF ( ~9SOC); whereas in hypersonic flight, at Mach ::o- 5, they are much
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

higher, for example up to 1,325°F ( ~ 720aC) in the X-15 research airplane flying at Mach 5
(see Figure 19.1), or over 4,000°F (~2,200aC) in a NASA hypersonic vehicle assumed to
cruise at Mach 8 at 88,000-ft. altitude (see [19.5]). Atmospheric re-entry involves surface
temperatures of similar or even larger magnitudes. For example, the orbiter of the NASA
Space Shuttle experiences maximum temperatures of about 2,500-2,600°F ( ~ 1,350-1 ,450°C)
at its nose and wing leading edges (see for example [19.11]); the planned U.S. National
Aerospace Plane (NASP), which was to have a peak Mach number of 25, was designed to
even higher surface temperatures of up to 4,000°F ( ~ 2,200°C); while the temperature in the
nose region of the Apollo lunar return vehicle reached over 19,800°F (~11,000°C).
Aerodynamic heating has therefore been extensively studied in the last five decades, as
have its consequences, thermal protection systems, thermal structures and high-temperature
materials. The physics of aerodynamic heating or aerothermodynamics is beyond the scope
of this book (the reader is referred for example to [19.12]-[19.15]), but it may be worth
emphasizing that the aerodynamic heating in supersonic flow, though important in itself,
differs significantly in character from that in hypersonic flow, which is a high-temperature
chemically reacting flow.
In hypersonic flow the vibrational excitation of molecules and the chemical reacting effects
of dissociation and ionization become important in turn as the temperatures exceed certain
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
Buckling Experiments: Experimental Methods in Buckling of Thin-Walled
No reproduction or networking permitted without license from IHS
Structures: Shells, Built-Up Structures, Composites
Not for Resale, 02/13/2019 01:32:58 MST
and Additional Topics – Volume 2. J. Singer, J. Arbocz and T. Weller Copyright © 2002 John Wiley & Sons, Inc.
1548 Thermal Buckling and Creep Buckling

Figure 19.1 Measured temperatures on a flight at Mach 5.0 of the NASA-North American Rockwell
X-15 research airplane in 1965 (from [19.10])

high values. From a specification of these temperature ranges, a velocity-altitude map (like
Figure 19.3, reproduced from [19.15]) can be drawn, on which both the flight paths oflifting
entry vehicles (with different values of the lift parameters m/CLS) and the flight regions
associated with various chemical effects in air arc shown. The 10 and 90 percent signs
indicate the effective beginning and end of the regions in which the respective effects are
important. As can be seen in the figure, the entry flight paths cut across most of these regions.
Hence the high-temperature effects are of major importance to entry-body flows.
Significant advances have been made in CFD methods to cope with the prediction of
aerodynamic heating in supersonic and hypersonic flows (see for example [ 19 .16] and
[19.17]), but the support of high-temperature wind tunnel tests is still essential. Moreover,
there are strong nonlinear coupling effects between aerodynamic loads, aerodynamic heating
and the resulting structural response, as well as the stability and control of the vehicle, that
have to be accounted for both in design and testing. This has led to coupled design methods
and introduction of some of these coupling effects into high-temperature testing.

Figure 19.2 Peak temperatures in op measured during flights STS-2, STS-3 and STS-5, respectively,
on and near the nose cap of the Leading Edge Structural System (LESS) of the NASA
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Space Shuttle (from [19.11])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Introduction 1549

Figure 19.3 Velocity-altitude map showing the flight paths of lifting entry vehicles, as well as the
regions of vibrational excitation, dissociation and ionization (from [19.15])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

The primary effects of the high temperatures due to aerodynamic heating in aerospace
vehicles, or of high temperatures from different sources in other structures, can be classified
as follows (see for example the early studies of Hoff, such as [19.1] and [19.2], or Bis-
plinghoff's 19th Wright Brothers Lecture [19.4], or more recently [19.9]):
I. Deterioration of the mechanical properties of materials at elevated temperatures
2. Thermal stresses due to temperature gradients. which may cause thermal buckling
3. Modification of stiffness and vibration properties of structural elements due to presence
of thermal stresses
4. Aeroelastic instabilities resulting from a thermal stress-induced reduction in stiffness or
from aerothermoelastic coupling
5. Creep, flow of the material under comparatively small stresses, which may cause creep
buckling
These effects will be discussed in more detail in the next section. Prior to that, one should
remember that the analysis of stresses and deformations in the presence of elevated temper-
atures and temperature gradients represents an important branch of the theory of elasticity,
thermoelasticity, to which a number of well-known textbooks and an extensive literature are
devoted (see for example [19.18]-[19.24]). The texts and reviews on thermal stresses usually
include separate chapters discussing thermoelastic stability (as for example [19.25]).
In thermoelasticity, or thermal stress analysis, one essentially employs the ordinary equa-
tions of elasticity, but with modified stress-strain relations that can be written (with reference
to an orthogonal coordinate system x 1, x 2 , x,)
(19.1)
where a is the coefficient of linear thermal expansion, A and f.L, the classic Lame elastic
constants, are defined by

,\--~~--
vE }
- (1 +Ev)(l ~ 2v) (19.2)
!L - 2(1 + v) - G

i and j go from 1 to 3, and ou is the Komeckcr delta. Note that the influence of the tem-
perature rise T(x 1, x 2 , x,) is manifested only by the last term in Eq. (19.1), and it only atlects
the direct stresses.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1550 Thermal Buckling and Creep Buckling

This formulation is usually known as uncoupled quasi-static thermoelastic theory since it


neglects the effects of mechanical coupling on heat conduction as well as of inertia.
For more precision, one has to note that if variations of strain are produced within a body
by an external mechanical agency, these are in general accompanied by variations in tem-
perature and consequently by a flow of heat. For its determination, the mechanical coupling
term has to be kept in the heat equation. However, since the deformations due to external
load are usually accompanied only by small changes in temperature, they can apparently be
calculated with no significant error; also if the thermal expansion is not taken into account.
Similarly, the influence of the strains produced in a body by non-uniform temperature itself
would probably be minute. Hence, the coupling term in the heat equation can usually be
disregarded, and the temperature distribution can be computed independently, before calcu-
lation of the corresponding deformations, i.e. using uncoupled thermoelastic theory. Simi-
larly, the coupled inertia effects have only to be considered in the case of actual dynamic
loading or movement.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
The coupling of temperature and strain fields, its implications and the cases that require
its consideration are discussed at length in some of the textbooks (see for example Chapters
1 and 2 of [19.18] and Section 1.9 of [19.21]).
Solutions of uncoupled thermal stress problems can be obtained either by the direct ap-
proach, outlined above, or by a superposition approach, which expresses them in terms of
ordinary stress-strain relations with fictitious body and surface forces.
Energy methods are another approach, especially useful in providing approximate solutions
when other methods fail. The theorem of minimum complementary energy is the energy
theorem that has found the widest application in thermal stress analysis. It is employed in
its usual form

8U,. = 0 (19.3)

U, being the complementary energy, which have become

U,. = f
\'
(U0 + aTa;)dV (i = 1,2,3) (19.4)

where U0 is the strain energy density expressed in terms of stresses and aTa11 is the term
that has to be added for the thermoelastic problem. Approximate solutions can then be
obtained by, for example, the Rayleigh-Ritz method.
As in other structural problems, powerful computer-oriented numerical methods have been
developed for more complicated structures and many multi-purpose computer programs are
available for thermal stress and thermal buckling problems (see for example [4.48], [2.94],
[2.98], [2.99], [2.1 02] and [3.20]).

19.1.2 Structural Responses to High Temperatures


The first and obvious effect of high temperatures on a structure is the deterioration of the
materials of construction, in particular their yield and ultimate stresses and their moduli. A
great deal of data has been collected on their behavior, and special materials have been
developed for high-temperature structures (see for example [19.26]-[19.30]).
The primary effects are a sharp downward trend in ultimate and yield strengths (or the
corresponding specific strengths) with temperature and a similar, though more moderate,
downward trend in elastic moduli (or specific modulus), as can be seen in Figure 19.4 for
several advanced materials. It is apparent that above 2000°F the designer has very limited
choices. It may be recalled that in the regions of lower temperatures <300°F, efficient su-
personic transports are essentially constrained to flight at M < 2.2 because of the temperature
limitations of the aluminum alloys employed in their primary structure.

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Introduction 1551

20

0
1.5 RSR·Ti
;:;
SUPERALLOYS
~
\'1
~
u
10
;:;:
8
[);
ACC

LOWER
OENSITY
CERAMICS

1,000 2,000 3,000 4.000


TEMPERATUP.E ,'F

500

2w 400
;:;
3.
§" 300
;:;
u

7075 -T6

1,000 1,000 3.000 4.00C·


TEMPERATURE.°F

Figure 19.4 Variation with temperature of the specific strength and specific modulus (indicating stiff-
ness) of advanced materials, in comparison with that of 7075-T6 aluminum alloy (from
[19.33])

Extensive efforts continue in the search for suitable higher-temperature materials, with
emphasis on the whole spectrum of their properties and with a continuous materials-design
interaction (see for example [19.31] and [19.32]). Since a discussion of the materials is
beyond the scope of this book, the reader is referred to the abundant literature, of which
only a few examples have been cited. One may note, however, that the high-temperature
environment of gas turbines has been one of the primary driving forces for the development
of new materials for many decades (see for example [19.32]-[19.34]).
The second undesirable effect that results from an elevated temperature environment is
the development of thermal stresses, often at temperatures considerably below those that
impair the material properties. Complicated structures do not heat up uniformly, even if
subject to a uniform heat input. In practice, however, the heat input is non-uniform, as for
example aerodynamic heating in aerospace vehicles, where the leading edges of the wing
and the nose of the body are exposed to much higher heating rates than the rest of the
structure.
The thermal stresses induced by these non-uniform temperatures in a structure are self-
equilibrating since they are not caused by externally applied loads. The stress distribution
over a cross-section of the structure consists therefore of both compressive and tensile
stresses. The compressive stresses may cause buckling of the slender or thin-walled elements
of the structure. This thermal buckling is an important mode of structural failure in heated
structures like supersonic and hypersonic aircraft, nuclear reactor components or special
storage tanks. It was studied extensively for plates and shells in the fifties and sixties (see
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1552 Thermal Buckling and Creep Buckling

for example [2.87], [19.36]-[19.42] and [19.43]). Thermal buckling (which is discussed in
more detail in Section 19.3) occurs as a result of compressive thermal stresses, which arise
either from non-uniform temperature distributions or from supports which constrain the ex-
pansion even when heating is uniform.
The change in effective stiffness of structural elements due to thermal stresses, and in
particular the reduction in torsional stiffness of wings of high-speed aerospace vehicles, is
another undesirable effect of aerodynamic heating. It too can cause buckling, but the reduced
stiffness has also a broader influence on the dynamic and aerodynamic behavior of these
flight vehicles. The change in stiffness is not associated with the change in material prop-
erties, discussed above, but depends only on the state of stress and may occur at stress levels
below those necessary to initiate buckling. One of the significant types of aerodynamic
heating relates to accelerated flight, as for example the heating of a wing rapidly reaching
supersonic speeds, and therefore the time often appears as a governing parameter in the
presentation of the effects of thermal stresses.
The reduction in torsional stiffness of wings caused by aerodynamic heating was exten-
sively studied in the fifties (see for example [19.4], [19.39] and [19.44]-[19.49]). As an
example, a double-wedge airfoil of solid stainless steel, instantaneously accelerated toM=
at 50,000 ft altitude, is considered (following the discussion in [19.46]). The chord of the
airfoil is 36 in. and the midchord thickness is 1.08 in., 3 percent of the chord. On the
assumptions that the initial temperature of the wing is the stratospheric free-stream temper-
ature ( -67°F) and that the temperatures are constant through the thickness, the calculated
temperature distribution along the chord is shown in Figure 19 .Sa for several points of time
following the instantaneous reaching of M = 3. An additional assumption of a constant heat
transfer coefficient in the boundary layer along the chord led to the symmetry of the tem-
perature distribution in the figure. The heat conduction along the chord was also neglected.
The variation of the temperature along the chord in Figure 19.5a is therefore simply the
consequence of the longer time it takes the thicker midchord region to heat up than the

600

500

_ _ _ _ _ _ J."
200
-
1-----= 36.. ----nl.'lO
100 M '3
ALTIT\.0' • 50,000 FT

~160

! 120

eo
1+------
. - - -36"
1
---Jl"
- - - - 108"
M ,3
Uyo
KSI ALTITLU: , 50,000 fT
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 19.5 History of thermal effects for a solid steel double-wedge wing after instantaneous accel-
eration to Mach 3 (from [19.46]): (a) history of temperature distributions. (b) history of
thermal stress distributions
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Introduction 1553

thinner leading and trailing edges. The axial thermal stresses resulting from these temperature
distributions are shown in Figure 19.5b. Eventually, after a long time, the temperature in the
wing would become uniform at the adiabatic wall temperature of 569°F, and then all thermal
stresses would vanish.
But at the early time intervals the non-uniform temperature distribution results in com-
pressive stresses near the leading and trailing edges with equilibrating tensile stresses around
the midchord (as can be seen in Figure 19.5b). Such a stress distribution, however, reduces
the torsional stiffness according to a simple formula

Glen = GJ0 + L ai 2 dA (19.5)

where GJ0 is the torsional rigidity (or stiffness) of the wing according to St. Venant, ux is
the spanwise thermal stress (negative when compressive), dA is an infinitesimal area of the
cross-section, r is the distance of dA from some reference point in the plane of the cross-
section, and the integration has to be carried over the entire cross-section. This formula was
derived independently by Hoff [19.45], Budiansky and Mayers [19.46] and Bisplinghoff
[19.4], on the basis of the idea that the first torsional mode is a sufficiently close approxi-
mation to the more precise deformation according to plate theory, and assumes a very long
wing. The idea was later shown by Singer [ 19.48] to be indeed a good approximation for
wings of large aspect ratios, but for short wings, as often used in missiles, a very significant
end effect occurs for smaller aspect ratios, say below 5, that has to be taken into account.
The effective torsional stiffness history, according to Eq. (19.5), for the solid steel double-
wedge wing instantaneously accelerated to Mach 3 is shown in Figure 19.6. Thus, about 1
minute and 15 seconds after the sudden reaching of Mach 3, the wing section sustains its

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
maximum loss of torsional stiffness, leaving only 22.5 percent of its original value. As flight
at Mach 3 continues, the temperatures become more uniform, the thermal stresses diminish
and the torsional stiffness gradually increases tending towards its original value GJ0 • The
shear modulus G has been assumed to remain constant with temperature, and any reduction
in G with temperature would lower the torsional stiffness even further.
Figure 19.6 also shows the effective torsional stiffness history for the same wing suddenly
accelerated to Mach 2 and 4. For flight at Mach 4, the effective torsional stiffness reduces
to zero after 21 seconds. The vanishing stiffness represents torsional buckling, and theoret-
ically the wing would unbuckle after 2 minutes and 23 seconds and start to regain its torsional
stiffness.
Since instantaneous acceleration to the respective speeds is not feasible, calculations were
also carried out with finite accelerations, which showed only slightly smaller reductions in
torsional stiffness. Further calculations for other solid wings, with constant thickness and
parabolic cross-sections (see [19.46] and [19.48]), exhibited similar effects.

10

~~~--~~2~£_~3--~~4~~-~~
TIME, MN.

Figure 19.6 Effective torsional stiffness, Glcrr• its history for a solid steel double-wedge wing after
instantaneous acceleration to Mach 2, 3 and 4 (from [19.46])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1554 Thermal Buckling and Creep Buckling

The reduction in torsional stiffness of fiat plates subjected to thermal stress distributions,
similar to those predicted for double-wedge and parabolic cross-section wings, was verified
experimentally by Vosteen and Fuller [19.44] in the mid-fifties. These experiments were
among the earliest thermal buckling tests and have been widely quoted. They are discussed
below in Subsection 19.2.1.
The reduction in torsional stiffness is accompanied by a corresponding reduction in the
frequency of torsional vibrations. Large deflections, or large amplitudes of vibrations, as well
as small pretwist, alleviate the effect of aerodynamic heating (see for example [19.47],
[19.49]-[19.51]).
Though not as pronounced as the loss of torsional stiffness, flexural stiffness may also be
affected by thermal stresses caused by aerodynamic heating (see for example [19.52]).
The fourth undesirable effect of thermal stresses caused by aerodynamic heating is the
promotion of aeroelastic instabilities, caused either by the thermal stress induced reduction
in torsional stiffness, discussed earlier in this subsection, or by aerothermoelastic coupling.
Since, however, aeroelasticity and the here-relevant extension, aerothermoelasticity, represent
a separate prime discipline that is beyond the scope of this book, the reader is referred to
the literature (for example [19.4], [19.53]-[19.57]). It may be mentioned only that panel
flutter at elevated temperatures, which was one of the major research topics in the sixties
and early seventies, is also influenced by panel buckling and postbuckling behavior.
The research emphasis then shifted to other aeroelastic subjects. However, with the re-
newed interest in supersonic and hypersonic flight in the late eighties and nineties, extensive

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
studies of the effect of high temperatures on panel flutter and other aeroelastic phenomena
were again initiated (see for example [19.58]-[19.63]). Some of these investigations also
considered the potential of active controls (for example, piezoelectric actuators) for sup-
pression of nonlinear panel flutter and improvement of aeroelastic performance.
The fifth undesirable effect of a high-temperature environment is creep, another phenom-
enon in the high-temperature behavior of materials. This is the flow, or continued change in
the shape, of a structural element when it is subjected to constant loads.
When a tensile test specimen of a metal is placed in an oven, heated well above room
temperature and maintained at a constant temperature, the application of a load to the lower
end of the specimen results in an elongation as shown in Figure 19.7. The instantaneous
elastic elongation, or possibly also plastic one if the load is high, are followed by an addi-
tional elongation that increases with time-creep.
Immediately after loading, if the load is not removed, the specimen continues to elongate
but at a decreasing rate. This range is called a primary or transient creep. Later the rate of
elongation, or creep rate, becomes constant, as shown by the straight portion of the curve

z0

~
3...

PRIMARY SECONDARY TtRTIARY

TIME

Figure 19.7 Creep of metal rod in tension (from [19.56])


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Introduction 1555

in Figure 19.7. This is known as secondary creep. Just before failure by fracture, the elon-
gation accelerates again and the creep curve becomes steeper again. This range is called
tertiary creep.
The speed with which creep deformations grow also depends on the temperature and stress
in the specimen. The complex relationship between creep deformation and the various par-
ameters influencing it was subject to extensive investigations in the fifties and sixties (see
for example [19.1], [19.4], [19.39], [19.64]-[19.70]), primarily because of the importance of
creep in supersonic aircraft and missiles as well as in gas turbines.
The same creep phenomenon as that described for tension, appears also in bending. But
when the load is compressive, a special time dependent buckling process, creep buckling,
may occur. Creep buckling is discussed in detail in Section 19.4, where the basic creep
phenomenon is further expounded.

19.1.3 Thermal Protection Systems


Before we consider the methods of high-temperature testing, it may be useful to discuss the
types of structures involved in these tests. The emphasis will be on aerospace structures
because of the unique character of their heat source, aerodynamic heating, and the predom-
inant requirement of low weight.
At present, there are three main concepts for the primary structure of supersonic and
hypersonic vehicles (see Figure 19.8): (a) insulated structures employing a thermal protection
system, (b) actively cooled structures, and (c) hot structures using high-temperature materials.
For supersonic flight the hot structure is the simplest approach conceptually, and indeed
has been used for example on the British-French Concord (limited to flights at M < 2.2),
on the Lockheed SR-71 and on the NASA-North American Rockwell X-15 aircraft. The
advantage of the relative simplicity of the hot-structure approach is offset by the disadvan-
tages that most materials suitable for elevated temperatures are heavy and that the high
thermal stresses present many design problems, which are made more difficult by the re-
quirement of a relatively smooth outer surface.
In short-duration, say 10-12 minutes, missions, like those of the X-15 research airplane,
a thick nickel alloy skin and titanium airframe served as a heat sink. For long-duration, but
lower Mach number, flights, like those of the SR-71, a high-temperature titanium structure
was employed, together with a high-emissivity surface coating at radiation equilibrium to

Beaded
km

.
Corruga!ed
(a) • webb

Shlle~ed
sl<rn

Durable thermal
(b) protection layer

Secondary coating

l1Quid
coolant
CirCUli

Lzquzd hydrogen
(c) from tank

Figure 19.8 Typical airframe structural concepts for supersonic and hypersonic aerospace vehicles
(from [19.34]): (a) hot structure, (b) insulated structure, (c) actively cooled structure (with
secondary coolant) --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1556 Thermal Buckling and Creep Buckling

reject part of the aerodynamic heating. But for hypersonic speeds some kind of thermal
protection of the primary structure has been regarded a more promising solution and has
been the main approach in recent decades (see for example [19.11], [19.33], [19.34], and
[19.71]-[19.73]).
A thermal protection system (TPS) may be defined as any device that prevents the tem-
perature of the primary structure from reaching undesirable high values. Both insulated struc-
tures and actively cooled ones are included in this definition, but here the active cooling will
be considered separately. The main functional elements of a thermal protection system are
shown in Figure 19.9 (reproduced from an early survey paper by Heldenfels [19.5]). A more
detailed and up-to-date presentation of thermal protection system concepts is displayed in
Figure 19.10 (taken from a NASA workshop on Current Technology for Thermal Protection
Systems [19.71]). The incoming heat of the airstream is disposed of by radiation or absorp-
tion in the vehicle, the system being classified as radiative or absorptive, depending on which
function predominates, although both are present in all applications.
A radiative system is inherent in the concept of the radiation equilibrium temperature.
Such a system in its simplest form might consist of as little as a highly emissive surface
coating. It has the advantage of being practically independent of time but is limited to a
maximum temperature or heat flux, determined by the materials of the structure.
The main feature of a thermal protection system is the barrier between the incoming heat
and the primary structure of the vehicle. When this barrier takes the form of an insulating
material, it operates at high external surface temperatures, which promotes dissipation of
most of the heat input by radiation. Since the required thickness of an insulating barrier is
determined by the amount of heat that can be absorbed in the underlying primary structure,
the insulation weight depends upon the total heat load and therefore on the duration of flight.
The insulation barrier does not have to share the load-carrying functions of the structure, but
only has to be able to withstand the exposure to the outside airstream.
Some types of thermal protection systems combine the barrier function and the heat ab-
sorbing function, but as they act primarily as heat absorbers, they are classified as absorptive
systems. Such systems (see Figure 19.10) include ablation, transpiration cooling (in which
gas is forced through a porous external surface of the vehicle) and film cooling, all of which
inject material into the hot boundary layer. The most important of these is ablation, with
ablative heat shields having been widely used in the short re-entry periods of spacecraft (like
Mercury, Gemini or Apollo) and ICBM's.
The insulated structure concept uses a structurally efficient material for the primary struc-
ture and protects it from high temperatures by a special thermal protection system, an in-
sulation barrier in Figure 19.9. The insulated structure approach has the advantage that it
eliminates the need for a smooth external surface and may therefore yield more efficient
structural designs. This is the approach adopted in the U.S. Space Shuttle Orbiter, where
rigid and flexible ceramics have been used for the TPS that protects the aluminum alloy
primary structure.

AERODYNAMIC 1ADIATED
HEATING

Figure 19.9 Functional diagram of thermal protection systems (from [19.5])


--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Introduction 1557

THERMAL PROTECTION SYSTEM CONCEPTS


SURFACE
SURFACE
HEATING
HEATIN~GRADIATION
RADIATION
SURFAC RAOIA TION CONDUCTION
HEATI INSUlATION
PASSIVE: . STRUCTURE

AIRFLOW

STRUCTURE
HEAT PIP E
ACTIVE :
SURF~
SURFACE D
HEATING
SURFACE~.
AIR FLOW HEATING ~ RADIATION

COOlA~
A IRFLOW

COOLANT COObAN
FLOW _ FLOW FLOW

TRANSP IRATION COOLING Fil M COOLI!iG CONVECTIVE COOllN<l

Figure 19.10 Types of passive, scmi·pa>~ive and ac1ive thermal protection syMem concepts (from
(19.711)

The orbiter tlremwl protection system (see [19.11] and 119.73]) was one of the major
technical accomplishments of the Space Shuu le program. Its main desig n requi reme nts were
(I) that the prime struc ture must be mnintained a t te mpe rat ures below 350°F (- Jso•q, (2)
that the TPS nn1st also pe rform satisfactorily in the other induced environmenL~ (suc h as
launch acoustics, detlcclions caused by aerodynamic loads. on-orbil cold soak a nd on-ground
salt. fog, wind and rain). and (3) be re usable for I00 missions, wilh minimal weigh!, main-
lcnance and refurbishment
The locmions of 1he various temperature proteclion malcrials employed in lhc Orbiter are
shown in Figure 19. 1 I. Three of lhe material systems of the Orbiter TPS are characterized
as reusable surface insula tions (RSI): 1\vo of the m are low-de nsity silica cera mic insulation.
and the third material consists of a coated Norncx (a type o f nylon) felt syste m. The silica
ceramic ti les are c l :1~~ificd in two categories: the high-temperature reusable surface insulalion
(HR$1). employed predominantly on lhe lower surface of tl1e Orbiter (sec Figure 19.11 ). and
the low-temperature reusable surface in$ulation (LRSI). used on lhe holter pans of the upper
surface, on the fuselage and the tail. The HRSI tiles arc coated with a black borosilicate
glass. whereas the LRSI tiles have a wh ite coating wi th the optical propenies required fo r
correct (ln-orbit tempe ratures.
The fl exible reusable surface insulation (FRS!) is the simplest TPS used o n the Orbiter.
It consists of a need led Nomex felt coated with a lhi n silicone elastomeric fi lm and protects
the structure in the areas of relatively low heat load (see Figure 19.11 ). However. in Orbiter
103 and subsequent orbiters an advanced Hexible surface in~ulation (AFRS I) replaced LRSI
and FRSI in many areas.
The HRSI and LRS I tiles are bonded to the Orbi ter struc ture with a sili cone adhesive and
a n interveni ng layer of Nome x felt. There a re many insu lation panels (approximately 24,000
tiles). and their joi nts can be a source of thermal gradienls. ln partic ular, in the higher-
pressure regions open tile-to-ti le gaps could result in ingestion o f high-temperaltlre gas flow
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1558 Thermal Buckling and Creep Buckling

ffiffi!U~ AtWorc.o c..boftooc,,~ fiiiOO)

_
c::::J :::.:.-===~·
~ ~.:.-:~~~=
c:J

-
= =
.. ......
=
.~~
,!.
l
:r::"--:::'~
fl
,
...
::*t

Fi(,'Ure 19.11 1be subsystems that constitute the Space Shuttle Orbiter thermal protection system (from
Orbiter 103 onward) and their materials (from [19.71J)

that would cause local tempcraiUre peaks. To prevent this. special gap fillers are used (as
detailed in [ 19.11)).
The nose cap and wing leading edge, in !he regions of highest temperature on the Orbiter
(sec Figure 19.2), are protected by a unique structural material cal led reinforced carbon-
carbon (RCC). This material (fabricated in a special manner detailed in [ 19.11]) yields a
hard carbon structure possessing reasonable strength and a low coefficient of themtal expan-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

sion. which provides excellent resistance to lhermal stress and !henna! shock.

Tlanlum Muklwal

Figure 19.12 Advanced metallic thermal protection system concepts ( from [ 19.72]. ( 19.741 nnd
[19.34J): (a) titanium multi-wall system. with a corresponding bayonet clip attachment.
Copyright Wiley (b) supemlloy honeycomb system. with a typical through-panel fa.tcncr
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
High- Temperature Testing 1559

The RSI tiles used on the Space Shuttle have performed well, and many satisfactory
vehicle tumaround operations have been accomplished. The tiles are excellent insulators but
are very fragile and may not be durable enough for second-generation space transport ap-
plications. Hence, new titanium multiwall, superalloy honeycomb and advanced carbon-
carbon multipost thermal protection systems are being developed at NASA Langley to pro-
vide durable protection for future space transportation systems (see for example Figure 19.11
and [19.72]).
In the third concept, actively cooled structures, the fuel, such as hydrogen, or a secondary
coolant such as water-glycol, is passed through tubes or cooling passages to convect the heat
away from the structure. It is the most complex concept, but because it has great potential
for the severely heated areas in hypersonic aerospace vehicles it has been extensively studied
in recent years (see for example [19.33] and [19.71]). A simple schematic of airframe direct
active cooling is shown in Figure 19.14a. The fuel, say hydrogen, is circulated through a
skin heat exchanger to absorb the heat load acting on the structure. The fuel is then routed
to the engine, where it is burned. Besides cooling the structure, the added thermal energy
absorbed by the fuel from the structure increases the specific impulse of the fuel, which
provides extra thrust.
The alternative to direct fuel cooling is secondary active cooling, in which another liquid
coolant flows through the heated structure to a separate heat exchanger, where the heat is
dumped into the fuel (see Figures 19.14b and 19.8c). Direct active cooling appears the most
promising for the engine structure, while secondary cooling may be as efficient as direct
cooling for absorbing the airframe heat load.
The use of hydrogen as the fuel provides an efficient coolant with a heat capacity 3.5
times greater than water. Hydrogen, however, presents some significant design problems. It
is a hazardous fluid, is difficult to contain, tends to embrittle most metals and subjects the
structure to very cold temperatures since it is stored cryogenically.
Active cooling, both direct and secondary has great potential, but considerable research
efforts are still required before it will be ready for flight.

19.2 High-Temperature Testing


19.2.1 Early Thermal Stress and Thermal Behavior Experiments
Investigations of structural effects of aerodynamic heating began in the late forties, and the
related aeronautical high temperature testing followed very soon (see for example [19.8],
[19.36], [19.77] and [19.78]). By the mid-fifties, substantial elevated temperature test facil-
ities had been developed at various research centers (such as NACA Langley, Wright Pat-
terson Air Force Base and Polytechnic Institute of Brooklyn in the United States and the
Royal Aircraft Establishment, Farnborough, in the United Kingdom) and considerable re-
search equipment had been generated (see for example [19.79]-[19.82]).

,~ RIB STIFFENED
ACC HeAT SHIELD

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Figure 19.13 Advanced carbon-carbon multi-post standoff thermal protection system concept (from
Copyright Wiley [19.74])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1560 Thermal Buckling and Creep Buckling

(a)

Figure 19.14 Schematic diagrams of (a) airframe direct active cooling and (b) airframe indirect active
cooling (from [19.71])

At NACA Langley Research Center a substantial research effort was in progress in the
mid-fifties (about 50 man-years, constituting the majority of the structures research group;
see [19.8]). Some theoretical and numerical thermostructural studies had already been carried
out by the early fifties (see for example [19.83], [19.84] and [19.36]), and exploratory tests
had been initiated (see [19.8] and [19.85]-[19.87]).
In 1951 a simple experiment (shown in Figure 19 .15) was devised at NACA Langley to
verify their thermal stress analysis methods (see [19.36]). Steady-state thermal stresses were
induced in a large, thick 7075-T6 aluminum alloy plate (36 X 24 X 0.25 in.) by heating the
center and cooling the edges.
The heat was supplied to the center line by Nichrome wire imbedded in porcelain cement and was
removed by water flowing through two plastic tubes glued to the plate. The size and material of
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

the attached apparatus were such that negligible resistance was offered to the thermal expansion
of the plate. The sides and ends of the plate were insulated with asbestos to minimize heat loss
through these surfaces and to ensure that the instrumentation was at the same temperature as the
plate. The flow of cooling water was controlled so that the temperature distribution was symmetrical
about the longitudinal center line and essentially constant in the longitudinal direction. No meas-
urable temperature variation occurred through the thickness of the plate.

The applied tentlike temperature distribution (see Figure 19.16a), and hence also there-
sultant stress distribution, was symmetrical about both centerlines of the plate. Strain gages

7)S-T6 PLATE AND


I INSULATING COVER

~ "36 in.

"'>
~
~ElECTRICAL
./· HEATER
L THERMOCOUPlES
AND STRAIN
GAGES

Figure 19.15 NACA Langley 1952 thermal stress tests-diagram of test specimen and test setup (from
[ 19.36] and [ 19.8])

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
High-Temperature Testing 1561

-4

(b) -8
SECTION -A SECTION - B

Figure 19.16 NACA Langley 1952 thermal stress tests: (a) the applied tentlike temperature distribution
(from [19.37]), (b) agreement between theory and experiment for longitudinal direct
thermal stresses (denoted D and 0 in the figure, from [19.8])

were therefore mounted only on one quadrant of the panel. At each measurement point three
strain gages were arranged in a fan rosette, with a similar group bonded to the opposite side
of the plate. Iron-constant thermocouples were mounted at the center of each strain gage fan
(see Figure 19.15), but only on one side of the plate.
The thermal stresses induced in the plate tended to deform it out of its plane, a combined bending-
buckling phenomenon associated with the initial imperfections of the plate. Since the present tests
were concerned with plane stresses, this deformation was minimized by mounting the plate in a
jig designed to provide the required restraint without introducing extraneous stresses. To further
insure that the experimental stresses would closely approximate a state of plane stress, the readings
of strain gages on opposite sides of the plate were averaged before the stresses were computed.
Three tests were made and in each the panel was heated until the temperature distribution
stabilized; then the strain gages were read.

After some processing of the data (detailed in [19.36]), the experimental results were
compared with calculations by an approximate complementary energy method (see for ex-
ample Figure 19 .16b) and were found to agree well. The experimental points shown in the
figure, and in similar ones in [19.36], are the average of the three tests. Separate points are
not shown for each test, since most of them are too close to be easily distinguishable. At
the time, though the maximum temperature ditierence was only 150°F, there were problems
with strain gage temperature errors, which were overcome by a special calibration technique.
Today such problems would be avoided by choice of the correct temperature-compensated
strain gages.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1562 Thermal Buckling and Creep Buckling

Because of their relative simplicity and well-defined tentlike temperature distribution, these
Heldenfels and Roberts tests have been widely quoted and employed as benchmarks for later
methods of thermal stress analysis.
The effect of steady (stabilized) elevated temperatures on local instability and compressive
strength of H-section plate assemblies, in addition to their direct influence on the material
properties, was also studied experimentally at NACA Langley in the late forties (see for
example [19.87]). With the addition of a small furnace to a well-tried room temperature test
setup, as well as a rigid-lever system to transfer lateral displacements to a gage below the
furnace and mounting thermocouples on the specimens, local instability tests were carried
out on 7075-T6 extruded H-section columns at up to 600°F. The test results showed that
available room temperature methods for calculating the critical compressive stress (see
[6.18]) can also be used for elevated temperatures, provided that the compressive stress-strain
curve for the material at the relevant temperature strain rate and exposure time is available.
It should be pointed out that the predictions include a small approximation due to an em-
pirically assumed increase in Poisson's ratio f.L with temperature.
The NACA Langley Structures Division also carried out a different type of exploratory
test in the early fifties, in which a series of small multiweb wings were subjected to super-
sonic jets (see Figure 19.17 and [19.85] and [19.86]). A NACA blowdown jet facility on
Wallops Island, Virginia, that incorporated a heat accumulator for stagnation-temperature
control was used. The wings were placed in the free jet at the exit of a Mach 2 square (27
X 27 in.) nozzle, which provided a temperature potential of 425oF to heat the model, that
could be maintained for about 9 seconds (with a 2-second starting period and 3-second
shutdown time).
The aim of the first of these tests was to obtain data on the temperature distribution in a
small multiweb wing. However, the aerodynamic heating and aerodynamic loads caused an
unexpected dynamic failure, which motivated the extensive later studies (at NACA and other
research centers) on the stiffness and frequency reductions of wings caused by aerodynamic
heating (see for example [19.44]-[19.51] and [19.8]). From the study of the results of that
test and of others that followed (sec [19.85] and [19.86]) it was concluded that ''the model
failed near the end of the test as a result of the combined action of aerodynamic heating
and loading. The rapid aerodynamic heating apparently induced thermal buckling of the
model skin, which in turn led to an unstable aeroelastic condition. The final result was a
dynamic failure that appeared to be some form of flutter."
The reduction in the effective torsional stiffness and the torsional vibration frequencies
caused by thermal stresses, induced by aerodynamic heating, was clearly demonstrated in
the 1955 Vosteen and Fuller experiments on a cantilever plate under rapid-heating conditions,
mentioned in Subsection 19.1.2 [19.44].
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 19.17 NACA 1952 supersonic Mach 2 free jet tests of multiweb wings-schematic diagram of
typical model in the jet (from r19.85])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
High-Temperature Testing 1563

The type of non-uniform temperature distribution produced by the aerodynamic heating


of a thin missile wing was simulated in these tests by rapid heating along the edges of a
cantilever plate (see Figure 19.18). The heat was applied to the plate for about 16 seconds
by carbon-rod radiators, and the temperature history on an edge and at midchord is shown
in Figure 19 .19a. Along the span the temperature distribution was constant, except for a
slight decrease toward the tip.
The chordwise variation in temperature at three instants in the heating cycle is shown in
Figure 19.18: (a) after 10 seconds of heating, (b) at the time of maximum edge temperature
(16.5 seconds), and (c) during cooling (30 seconds). The temperature remained relatively
low over the center half of the plate but rose sharply near the heated edges.
Two effects of the nonuniform temperature distribution on the stiffness were examined:
(1) the deformations due to the thermal stresses and (2) the changes in natural frequency
during heating. The tip rotation histories are shown in Figure 19 .19b, where they are plotted
for no external load and for applied torques of 400 in.-lb in each direction. In each case the
cantilever plate deformed by rotating torsionally about the midchord line. As the plate cooled,
the torsional deformation decreased and eventually the plate returned to its original position.
Note that the plate underwent a substantial rotation without the application of external
load-it experienced thermal buckling. Because the plate represented a typical solid fin of
a missile, this thermal buckling could significantly influence the control of the missile. Due
to an initial twist, the plate rotated in the same direction regardless of whether heat was
applied symmetrically (as shown in Figure 19.18) or asymmetrically on either edge. The
combined application of a mechanical torque and the thermal stresses indicated an approx-
imate superposition of deformation.
The measurement of the natural frequencies under transient heating conditions was difficult
because of the time required to establish resonance. As the research continued, the research-
ers at NACA Langley developed a system for following a resonant frequency as it changed
during a heating test (see [19.8]). Today the modern sophisticated instrumentation could
overcome this difficulty. But at the time, in the 1955 NACA tests, the study was limited to
the first bending and torsion modes (see Figure 19.20, showing the frequency history of the
first torsion mode). During the heating the plate was periodically struck to excite these
fundamental modes. The peak reduction in torsional frequency was 35 percent, whereas the
first bending frequency decreased only 21 percent at the point of maximum temperature
gradient. --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

These essentially simple experiments became benchmarks for many later studies.
A number of hot jets and hot tunnels were operational at NACA Langley in the fifties, in
particular the 9 X 6 ft Thermal Structures Tunnel, and additional ones were planned (see

T °F

25 50 75 100
PERCENT CHORD

Figure 19.18 NACA Langley 1955 tests on the behavior of a cantilever plate under rapid-heating
conditions-measured chordwise temperature distributions of the radiantly heated plate
at three instants in the heating cycle (from [ 19.44])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1564 Thermal Buckling and Creep Buckling

300
EDGE

200

MID CHORD
T, °F

(a)
10 20 30 40 50
TIME, SEC

APPLIED TORQUE

10
TIME I SEC

Figure 19.19 NACA Langley 1955 tests on the behavior of a cantilever plate under rapid-heating
conditions (from [19.44]): (a) measured temperature histories for an edge and the
midchord line, (b) measured tip-rotation histories for applied tip torques of 0 and
±400 in.-lb
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 19.21). After various alternatives were studied and tested, it appeared that a true-
temperature, Mach 3, blowdown wind tunnel was the best approach for high-temperature
research on large structures. This concept materialized in the 9 X 6 ft TST, which became
operational in 1957 and was used to test a variety of structural models till 1977, when a
structural failure in the air storage field made further operations impractical.
The ethylene jet and ceramic heaters were very high-temperature supersonic jets for testing
materials and small structural models, a capability subsequently carried to extremely high
temperatures by the electric-arc powered jets. The 7 ft HTF concept eventually became the

1.00

·-
I

.25
\I I
I
\
I
!
I

10 20 30 40 50
T, sec

Figure 19.20 NACA Langley tests on the behavior of a cantilever plate under rapid-heating condi-
tions-change with time in first torsion mode frequency and comparison with calcula-
tions (from [19.50] or [19.8])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
High-Temperature Testing 1565

[ UIYlftl( .IET
lf l
)/4'
[
CERAMI C tt:AitRS
'-4"
-

r KTF
\ J
-.J!I!:'I
11~" ~ 1-'L!lQ"
ARC~
Fi~.ture 19.21 NACA Langley opcrntional (black) and planned hot jets and hot tunnels in April 1959
(from [19.8]). Note the well-known 9 x 6 ftthcrmal structures tunnel ("rST)

prc~enl NASA Langley 8-Foo1 High-Temperature Structures Tunnel. again a true-


temperature. Mach 7. blowdown wind tunnel. which became operational in 1968 (~ee 119.8)).
A ~ystema1ic search for ways 10 simula1e or duplicme aerodynamic healing in 1he labo-
ralory began al Langley in 1951. A variety of devices were evaluated for radiative and
convective heating of struciUres. A goal wa.' set to achieve initial heating rates of I00 Btu
per square foot per second ( - 106 kW /ft' or 1136 kW /m 2). a va lue derived from calcu lations
of the heat transfer rate to :ti rplanes accelerating to M ach 3 or Mach 4 at 50,000 ft altitude.
It may he noted that at the same time the researchers at RAE Farnborough in the U. K.
comidcrcd similar heating rates. about 100 kW lf't' (sec 119.801).
Radian! heating soon became the most widely employed method of rapid heating. simu-
lating aerodynamic heating (or kinetic heating, as it i$ $Ometimes referred to). At 'ACA
Langley the first extensively used rapid heating device wa.s a carbon-rod radiator. It was
soon replaced by tungs1en filament lamps, which were much better radian! heaters. They in
tum were replaced by the more powerful q uanz tube lamps that General Electric was de-
velopi ng in the early fifties. S imu ltaneously, quartz tuhe lamps also served us the prime
radian! heating tool at the RAE Farnhorough in the U.K. As more experience was gai ned at
various research centers, these high -intensily in frared radiant heaters became the universal
equipment for rapid heating of structures. They will therefore he d iscussed in more detai l in
the neu subsec1ion.
Wright-Paucrson Air Force Ba~e was another very active U.S. high-lempemturc research
center iniliated in lhc mid-fif1ie~. The U.S. Air Force Air Research and Development Com-
mand decided in 1952 to establish a laboralory for full-scale slruclural lesiS. combining
mechanical and 1hermal loads, in preparmion for future supersonic aircraft, and directed
WADC to bui ld a 3 MW pilot heati ng facili1y. Originall y it was planned, after develo pment
of the test methods at WPAFB. to locate 1he 50 MW full -scale hot-test facility at Arno ld
Engineering Development Center (AEDC) in Tennessee. However. when it was discovered
thai AEDC was limited in :tvailablc elec1ric power. it wa~ decided 10 build the full-scale hm-
teM setup at the Wrighl Field Structures Test Facility. where it became operational in 1961
(sec 119.77)). No1e the characterislic very high electric power demand of large high-
temperature test facilitic;.
Three rne1hods for heating were studied at the WPAFB high-tempemlure facili1y in the
early sixlies: radiant. conduction and induction. For radiam hearing. various infrared heat
lamps were 1ried wi1hout success until 1he T-3 quartz envelope lamp appeared, whi ch proved
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,

very successful for temperatures up to IOOOoF (- 640°C). With special rencctors and lamp
cooling systems, which wi ll be d iscussed in 1he next subsection, 1he T-3 lamps could operme
al temperatures up to 3200°F ( - 1760°C).
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1566 Thermal Buckling and Creep Buckling

Conduction heating through bonded pads or heating blankets to reach uniform tempera-
tures up to 500°F ( ~260°C) was also investigated, but the method was not successful because
pad bonds failed during heating and the large thermal inertia prevented rapid heating. In-
duction heating appeared to be more successful but aroused some difficult problems, which
researchers at other centers also encountered. These are discussed in the next subsection.
The WPAFB pilot heating facility was also used for certification tests almost as soon as
it became operational. Among these were tests on well-known aerospace vehicles, such as
the Titan ICBM, the F-106 fighter and the B-SS bomber (see [19.77]). For example, the aim
of the Titan test was to simulate the severe thermal gradients due to internal cryogenic fuel
and external aerodynamic heating simultaneously with the mechanical compression loads in
the skin caused by the engine thrust. The safe fuel simulant of liquid oxygen used in these
tests was liquid nitrogen.
As another example, the hot-test condition for the F-106 was a transient condition rep-
resenting takeoff on a cold day and a rapid acceleration to Mach 2. The internal fuel of the
aircraft was at oop ( -17 .S C). A Titan fuel tank section was used for the fuel simulant feed
0

tank, ethylene glycol being the fuel simulant. Adding dry ice reduced the temperature in the
feed tank to -sop (-20HC). The cold ethylene glycol was then pumped into the wing fuel
tanks of the F-1 06. Wben it reached the required OoF (- 17. SoC), the heating was programmed
at a specific rate until the peak skin temperatures reached 260oF (127°C). The mechanical
loads on the aircraft were increased during the heating from limit to ultimate load, and the
F-106 withstood the combined thermal and mechanical loading successfully. In the B-SS test
a similar procedure was employed. In both tests the mechanical loads in the heated area
were applied through small silicone bonded tension pads that could sustain loads at temper-
0
atures up to 550°F (22S C).
In the early sixties large components of the hypersonic X-20 Dynasoar project (of both
the hot-structure concept and the actively cooled one) were tested at the WPAFB high-
temperature facility at peak temperatures of up to 3000oF ( ~ 1650°C). Later, in the mid-
sixties, two major hot tests were carried out: one, an actively cooled, ceramic-coated, metal
honeycomb structure, the so-called thermantic structure; and the other a one/third scale wing-
fuselage component of the Advanced Structural Concepts Experimental Program (ASCEP,
which was part of the first aerospace plane program).
Then, by the end of 1969, the U.S. Air Force suspended its interest in hypersonics, and
hot-structures research at WPAFB hibernated for 20 years, till it was awakened again by the
National Aerospace Plane (NASP) project in the early nineties.
In the sixties the WPAFB high-temperature facility also dealt with the simulation of nu-
clear thermal effects on structural panels. The severe requirements of this simulation led to
the development of very high-heat flux test techniques, for example graphite heaters that
produced up to 300 Btu/ j2 sec ( ~ 3400 kW I m 2 ) for the short test times required for nuclear
simulation (about three times the heat rate of the usual quartz tube lamps).
The effects of high-temperature gradients occurring in nuclear power plants have also been
studied experimentally in a number of research centers. For example, at the French Atomic
Energy Authority (CEA) Saclay Research Center, thermal buckling in thin shells due to
thermal gradients was investigated in the eighties. The thermal gradients were simulated in
those tests (see [19.SS]) with the aid of induction heating, discussed further in the next
subsection.
Another important U.S. high-temperature research center was the NASA Dryden Flight
Research Facility at Edwards, California. Dryden's principle mission has been to conduct
flight research on high-performance aircraft, but operational problems on the X-15 research
aircraft in the early sixties required the support of extensive laboratory thermostructural
testing, which led to the development of the facility to an active research center (see for
example [19.SS]-[19.97]).
The method of using calibrated strain gages for measurement of flight loads was developed
by NACA in the fifties and has been employed extensively at NASA Dryden. While the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
High- Temperature Testing 1567

method was being used to measure X-15 wing and tail loads at the higher Mach numbers,
extremely large errors were observed. The early efforts therefore focused on the development
of a laboratory thermal simulation technique to determine the response of the calibrated
strain gages to the aerodynamic heating effects. This gage output, as found in the laboratory
heating simulation, was then used to correct the flight load measurements. After some pre-
liminary tests, the X-15 horizontal stabilizer, and later its wing, were employed to develop
the thermal loads calibration technique and then other thermal simulation heating procedures
(see [19.90]-[19.92]).
In the early seventies the thermal loads calibration technique was applied to the YF-12
aircraft as part of its flight load program. The laboratory investigations of the flight vehicle
included a Mach 3 thermal simulation, which involved heating the entire 5000 ft 2 ( ~465
m2 ) of the airplane's surface, after which the aircraft was returned to flight status.
After completion of the extensive laboratory heating tests of the YF-12 airplane, hot-
structures testing at the Dryden center focused on components tests, like the Hypersonic
Wing Test Structure (HWTS). This was a full-scale Rene 41 wing component of a hypo-
thetical Mach 8 flight vehicle, which served as a test bed for over 15 years, during which
time it accumulated over five hours of simulated rapid heating at temperatures in excess of
lOOOoF (-~540°C) (see for example [15.65], [15.66] or [19.98]). The Dryden thermostructural
facility and its instrumentation are discussed further in the subsections below.
Other components were tested in the Dryden facility, like the Space Shuttle elevon and a
thin-shell Inconel X-750 leading-edge concept for the X-15. In the nineties, the NASP project
motivated the development of a new Liquid Hydrogen Structures Test Facility at Dryden,
which initiates a new class of thermostructures testing with liquid hydrogen.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
The Royal Aircraft Establishment at Farnborough in the U.K., one of the pioneering high-
temperature research centers, was already very active in the mid-fifties (see [19.78]-[19.82]).
For both steady and transient heating on complete structures and structural components,
infrared radiant heating with quartz lamps (gas-filled lamps, consisting of a coiled tungsten
filament held in the center of a clear quartz tube by a series of tantalum disk spacers) had
become the usual practice at Farnborough. These radiant heaters were arranged in arrays
(see Figure 19.22), which were the building blocks of the heating simulation systems. Two
standard lamp arrays were established: the basic fiat panel system (Figure 19 .22b) and an
adjustable versatile curved array (Figure 19.23). Special support frame elements were also
developed, essentially simple braced framework beams whose top and bottom tubular flanges
served to carry cooling fluid. The lamp arrays were attached to these Warren girders by a
system of tubes and clamps that facilitated their assembly. Figure 19.24 shows a typical
open-ended oven surrounding a structure that used these support elements.
The quartz lamps were designed for 220-230 V and an output of 1 kW per lamp. At the
design voltage a long life, of up to 5000 hours, was expected. However, for the aeronautical
short-time transient heating applications a shorter life was acceptable and considerably
greater outputs could be achieved by "overvolting." The lamps could be operated at up to
double voltage for short periods, and then they had an output of 3 kW each. The normal
lamp array had eight quartz lamps, but for higher heat concentrations the lamps could be
placed closer together and could even be run in multibanks. The radiant heating rate was
then 130 kW /ft 2 • If the surface of the heated structure was well blacked, it was possible to
absorb about 60 kW /ft2 of the emitted radiation.
The Farnborough researchers therefore set a heating rate of about 100 kW /fe as a practical
goal. As pointed out earlier, this was also roughly the goal set by the NASA Langley ther-
mostructural investigators at the time. One may observe the similarities between the G.E.
quartz lamps and the Phillips ones developed across the ocean. Essentially, the techniques
and performance limits of radiant heating with quartz lamps established at NASA and at
RAE Farnborough in the fifties remain unchanged to this day, although better and more
efficient reflectors and in particular vastly improved control systems have been developed
since (see for example [19.98]).
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1568 Thermal Buckling and Creep Buckling

(a)

(b)

Figure 19.22 Radiant hea1e1· arrays as developed al the Royal Aircraft Establishmenl. Farnbomugh in
the fifties (from 119.821): (a) Phillips quartt tubular heater array componcn!S (the ap-
proximate output of the array was I kW al 230 V and 3 kW at440 V). (b) the •~sembled
hea1er array with a pyrometer anached

Fi~turc 19.23 Radiant in frared hcming at the RAE Fa111burough. in the fi fties-an infinitely vari able
curved reAeciOI' and lamp array (from [ 19.821)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
High- Temperature Testing 1569

Fig ure 19.24 Radiant infrnrcd heating at the RAE Fambomugh in the fiflies-an open-ended oven
surrounding a struciUre. which shows the swnd:u·d supp<>rt clements (from [19.80])

The RAE Famborough re~earchers also considered electromagnetic induction heating as


an alternative method for heating rates beyond the 60-100 kW/ft1 limit of infrared radiant
heating. as required. for example. for simulation of re-entry conditions. Induction heating.
sometimes referred to as radio frequency heating or eddy current heating. essentially consists
of generation of power in the form of periodic electric or magnetic fields of high frequency.
it$ translerence to the heated structure and its absorption by it. Metal structures, which are
conductors. cannot be penetrated by electrical fields and can therefore only absorb power
from magnetic fields, via the eddy currents generated in the metal.
Wi th induction. or radio freq uency, heating very high rates could be obtained, up to about
1500 kW/ ft2 (see for example 119.79]). But the RAE investigators soon discovered the
difficulties with induction heating: the heating depended on the material of the specimen.
the design of the induction coils and the assessment of the eddy curreots were difficult. the
cost of the apparatus was higher than that of other forms of heating. and radio frequency
interference was experienced over a wide area unless expensive screening of the entire spec-
imen was made.
Si milar problems were also encountered by WPAFB researchers (and their University of
Florida colleagues) at the time (sec for example [ 19.77]), as well as at other laboratories
that tried to exploit the potent ial of induction heating. For instance. the senior author re-
members vividly that when. as an M.Sc. student at the Polytechnic Institu te of Brooklyn in
the early fifties. he used a very small induction heater in a creep test, he had to work in a
shielded "cage" to prevent upsetting all the Brooklyn Poly communications systems.
In parallel with the large high-temperature research facilities discussed. a number of
smaller aerospace thennostructural faciHties were established at some universities in the
fifties and sixties (for example, the Polytechnic Institute of Brooklyn. Stanford University.
Cran fi eld Institute of Technology in the U.K., the Royal Institute of Technology, Stockholm,
and the Techni on in Israel).
Having looked back at the early thermal behavior experiments and havi ng perceived their
dominant inH uence on today's thermal simulation metltods, we can proceed to discuss these
method~ in detail.

19.2.2 Methods for Rapid Heating


Nowadays. thermal ground testing of high-speed night vehicles or component~ can use dif-
ferent means of heating: high-temperature wind nmnels, graphite heaters, arc lamp heaters.
l a~crs, induction heating, convection heating and convent ional infrared (TR) radiant heating
wit h quartz lamps. However. IR quartz heaters arc universally accepted as the preferred
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1570 Thennal Buckling and Creep Buckling

method of realistically simulating the effects of aerodynamic heating on high-speed flight


vehicles' stmctural test components since they offer the most efficient (approx imately 90
percent efficiency) and diverse method of testi ng. For example. at the NASA Dryden Thcr-
mostructures Research Facility quartz lamps are used exclusively as a means of heating (see
[ 19.97]). Modem thermal testing with infrared lamps is therefore discussed in some detail,
folJowing the presentation by Fields [ 19.98], while the other methods are brieny considered
later in this subsection.
Infrared energy produces electromagnetic radiation just below the visible spectrum, mostly at wave
lengths from 0.5 to 2 m. The radiation can easily be focused on small areas for maximum heat
flux and temperatures, or it can be distributed over the surfaces of large test articles to duplicate
flight heating variations.
j1be aerospace experimenter should note that] there is a significant difference between aerody-
namic and IR heating. ln·fiight aerodynamic heating is a function of the difference between the
recovery temperature and the component surface temperature. Laboratory radiant heating is a func-
tion of the difference between the quartz heater temperature to the fourth powel' and the surface
temperature to the fourth power. An appreciation for this fundamental difference is necessary in
the design of m heater systems. For instance, pl'oper distribution of quartz lamps in a system may

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
be required to accommodate this difference. pruticularly in locations where there is a significant
heat sink.
. . The quanz lamps commonly used for IR heating tests consist of a tungsten filament enclosed
in a quartz glass rube. These tubes are approximately 3/8 in. (-9.5 mm) in diameter and of varying
lengths (Figure 19.25). The rated wauage of the lamps can be user selected to meet the specific
test J'equil'ements. Figure 19.26 shows the characteristics of a 200 W /in T-3 10-in. lighted length
quartz lamp with high-temperature end seals.... this lamp produces 2000 W at rated voltage and
6000 W at double rated voltage. The lamps offer the most efficient heating when operated at the
highest possible temperature. Operating at rated voltage, usually 240 V, the lamp filaments reach
nearly 4000' F (- 2200'C). If the lamps are operated at double rated voltage or 480 V, the filaments
can reach 5400' F (- 2980' C).

However, as mentioned in the previous subsection, overvolting significantly shortens the


lamp life. But quartz lamps are usually quite durable, with a life span from a few hours to
several hundred.
The types of reHectors used depend upon the heating requirements. There are generally
three classes of reflectOrs employed in directing the radiant energy from the lamps to the
test specimen. For low-temperature tests, say less than SOO' F (- 430' C), simple polished
metal reflectors work satisfactorily. For example, for the YF-12 forebody the reflectors were

,,

Figure 19.25 Typical infmed quartz lamps in use today (courtesy of NASA Dryden Flight Research
Cemer)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
High- Temperature Testing 1571

12,000 20

8000
Watts 10 Amps

4000

L-~~--~L------L--L----L------_J o
0 200 400 600 800
Volts

Figure 19.26 Charactc1istic• of 200 W/in. T-3. 10 in. quartz lamps (from [19.98])

0.075 -in. ( - 1.9-mm) thick type 310 stainless-steel sheets. bent to confonn the contour of
the airplane. Because reHectors of this 1ype get hot. gaps have to be left between panels to
allow for thermal expansion. which leads to designs with swivel attachments to provide for
expansion and adjustments.
[For medium]temperatures. about ISOO'F (- 81SOC). gold plated stainless-steel reflectors (Figure
19.27) can be used. These rellectors. as with the previous ones, arc passively cooled with back
side radimion and natur.ll convection. These units have air cooling for the ends of the quartz lamps.
For higher temperatures, above ISOO'F, it is necessary to use either active cooling or a heat
resistant material such as ceramic (foamed silica).

A large array of water cooled polished aluminum refleclors, which was employed to heat
a test article to 1900°F (- 1040"C) is shown in Figure 19.28. A copper line was added to
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

the reflectors of thi s array to provide cooling for the lamps and seals. Some type of effective
cooling is usually required for lamps and end seals at high temperatures. and this is often
provided by the reflectors.
Many test setups usc ceramic reflectors. which operate at very high temperatures. say 2500°F
(- 1370"C), without the need for cooling. Tite reflectivity of these units is much less than the
previously discussed reflectors. so the initial elliciency at low temperatures is less. At higher tcm·
peratures. the surface absorbs sufficient energy >O that it begins to reradiate and improve the
efficiency. The reradiation is a longer wavelength which is absorbed by the quanz tubes, and hence
creates a problem of lamp heating.
Tite radiant energy or hear flux provided to a test specimen is dependent upon the entire test
system. which is unique for each test setup. In addition tO the quartz lamps, the test system includes:
( I) the type and efficiency of the reflector used to direct or focus the lamp radiation. (2) lamp-to-

Figure 19.27 Gold-plated stainless steel reflector for medium temperatures, about ISOOF (-8150C)
(counesy of NASA Dryden Flight Research Center)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1572 Thermal Buckling and Creep Buckling

Figure 19.28 An array of water-cooled polished aluminum reflec10rs. employed to heat a test article
to 1900• F (-1040°C). (counesy of NASA Dryden Flight Resea,·ch Center)

lamp spacing, (3) lamp-to-specimen d istance. (4) boundary conditions (end effects), (5) test spec-
imen surface emissivity, (6) convection (natural or forced) around the test article. (7) test
specimen-setup discominuities, such as heal sinks or variable geometry. and (8) possibly a moveable
tesl article. These factors delennine the electrical power requirements of a syslem. lhe maximum
lcmperaiUres. and lhe maximum heating rmes which can be achieved nn any given test componem.

Typically. temperatures of 2000-3000"F (- I I00-1650"C) and heat fluxes up to 80 Btu/


ft2 sec ( - 900 kW /m 2) are easily achievable.
The heat flux loss of the heater system. which is the cause of the difference between the
flux emitted by the lamps and that absorbed by the test specimen, includes conduction and
convection. To avoid elaborate heat transfer calculations, an empirical curve of total heat
loss as a function of specimen temperature was derived at NASA Dryden from past heating
tests employi ng water-cooled pol ished aluminum reflectors or stainless steel reflectors. T his
curve. shown in Figure 19.29, is based on a specimen emissiv ity of approximately 0.85.
Since the heating tests used to derive this curve were conducted to only 250<rF (- 1370°C),
the curve is verified up to that temperatu re on ly, and the part of the curve from 2500-4000•F
(- I 370-2200"C) is an extrapolation. The curve docs not include heat losses due to end
effects.
"End effects on the boundaries of TR heater systems can be a major source of temperature
en-or" and are considered in some detail in the relevant reports (for example in [ 19.91].
119.98 ] and [19. 105]). These studies ill ustrated that boundary conditions have to be taken
into account in the design of IR heaters, and that as a minimum the retlector should extend

100
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

qk).ss '
~so

f1 2-sec

0 2000 4000
Specimen temperature. • F

Fig ure 19.29 Heal flux loss-empirical curve derived rrom pas1 heal ing 1csts at NASA Dryden. using
wmer-cooled polished aluminum or SLainlcss steel reflectors (from [19. 98])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
High-Temperature Testing 1573

approximately 10 in. ( ~ 250 mm) beyond the test specimen. Furthermore, end reflectors are
often required to minimize the end-effect heat losses.
"Figure 19.30 shows the power required as a function of heater area for specimen tem-
peratures up to 3800oF ( ~2090oC) and heating rates from steady state (SS) up to 10 Btu/
ft 2 -sec ( ~ 114 kW /m 2 ). The calculations were again made for a specimen emissivity of 0.85.
At very high temperatures (3500°F to 3800°F, or 1930oC to 2090°C), the slope of the curve
is such that the heater area is quite small." One may again note the substantial power
requirements of the heaters (tens of megawatts). For instance, the maximum power available
at the NASA Dryden Thermostructures Facility is 20 MW [19.97].
The infrared heater is (usually) controlled by a closed-loop system, as illustrated by the block
diagram of [Figure 19.30 (reproduced from [19.91))]. A control thermocouple at a particular spec-
imen location, chosen to be representative of the zone, generates a millivoltage which is compared
with the function-generator output; this output represents the simulation temperature time history
programmed for the control thermocouple location. If there is a difference (error) between these
two signals, the controller commands the ignitron power regulator to supply more or less power
to the heater, maintaining the error near zero.
The quartz lamps are generally grouped or electrically wired together into zones. Each zone is
controlled to provide a specified temperature-heat-flux time history. Control methods vary [from
manual control to simple closed-loop systems (like Figure 19.31) or sophisticated computer con-
trolled closed-loop systems with complex algorithms (see for example Figure 19.32). The latter
method is mandatory for tests with a large number of control zones.
The data acquisition and control system currently in use at the NASA Dryden Thermostructures
Research Facility is summarized in [Figure 19.32]. The unique feature of this system is its ability
to conduct real-time simulations of thermal load on aerospace vehicle structures. With quartz lamps
placed around the test specimen, the surface temperature within each zone of lamps is forced to
follow a desired time history, based on feedback from a control thermocouple. An adaptive algo-
rithm is used to compute commands to silicon controlled rectifier power controllers. As many as
512 independent thermal zones can be controlled to distribute up to 20 MW of available power.
The real-time system resources may be dedicated to a single test or may be shared among as many
as three autonomous test activities conducted simultaneously.
The suiface emissivity affects the power required by the infrared heating system. "Part of
the total radiant heat flux from the heater is absorbed by the test specimen and the remainder
of the heat flux is reflected. Since absorptivity is equal to emissivity, the amount of the heat
absorbed depends on the emissivity of the test specimen and the heat reflected depends on
the reflectivity of the test specimen." Since emissivity plus reflectivity equals 1.0, it follows
that the minimum power required will occur when the emissivity of the test specimen ap-
proaches 1.0. On the other hand, when the reflectivity approaches 1.0, the maximum power
will be required. The emissivity of the test specimen should also be as uniform as possible.
The application of combined IR heating and mechanical loading is a common problem in
hot-structures testing. Some possible approaches include: "(1) hard points may be designed

50

40

Power, 30
MW
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

20

55 = 51eady slale
10 Emissivity =0.85

500 1000 1500 2000 2500


Heater area, ft 2

Figure 19.30 Power required for IR heating as a function of heater area (from [19.98]). Note the large
power requirements
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1574 Thermal B uckling and Creep Buckling

41)~11 a. c. powtr

Conhtl
lhtlt'IIOCOU,It
lttdbKl

- ....
Figure 19.31 Block diagram of a simple closed-loop laboratory infrared heating system (from [19.91))

inlo a 1es1 struclure for load in1roduc1ion [see for exa mple ( 19.94)], (2) cx1ernal load pads
can be used 10 distribule load inlo wucturc, (3) whipplc-lrees can be used 10 distribule loads
10 1he 1es1 article 1hrough discrelc auachmenl poi nts [sec for example Figure 19.991. and (4)
buffer bays or load cxlcnsions may be included with the 1cs1 article design 10 provide load
1hrough an adjoining slruclurc. No maner which approach is sclec1ed. imerferencc with the
IR healer system and the desired 1emperature distribution will be encountered." Conse-
qucmly, early planning during the design of structural test components for combined loading
and healing is essemial.
Some details of 1ypical IR heati ng tests carried ou1 hy NASA Dryden are now presented.
10 illumate practical applicmions.
The first example is the wing of the X-15 hypersonic research vehicle (shown in Figure
19. 1).
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`

Fil(ure 19.32 Data acquisition and computer-controlled closed-loop TR heating ;ystcm in usc at the
NASA Dryden ThennostrucLUres Research Facility (from [19.98))
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
High- Temperature Testing 1575

The wing [see Figure 19.331 was a short-span, lhin. low-aspecl ralio, mu ltispar strucwre. It had
lhree main ribs: a root rib, a midspan rib, and a lip rib. The leading edge was a segmented slug
or heat sink having a conslant radius. The wing-to-fuselage anachmenlS consisled of five A-frame
assemblies which were an integral pa11 of lhe wing. The wing skins, tip rib, from spar, and Structure
fonvard of lhe front spar were c<mslructed of lnconei-X (a nickel alloy): the remainder of Lhe wing
struciUre was titanium aHoy.
. . . Preliminary healing 1ests [see ( 19.93)] showed discrepancies between lhe simulation lem-
pcratures measured wilhin each zone. olher lhan control lhermocouples, and wilh calcul31ed ICm·
peratures. The lack of uniformity in the surface emissivity was the mos1 likely problem. [Indeed,
a fresh coat of high-emissivity paint solved lhe anomalies. ]
The quanz lamps in the IR healer system for 1hc wing were divided into 24 upper and lower
surface zones. as shown in [Figure 19.34a]. Two addilional zones were added 10 provide 1he
necessary heal flux to the leading-edge area. There were more zones chordwise than spanwisc tO
bener simulale the in-fligh1 surface lemperaiUre gradienls. Figure [ 19.34b] shows a cross-section
of the wing and IR healer a1 a roo1 s1a1ion. Since 1he wing Sll11clure does nol represenl a uniform
heal sink. the additional heat fl ux 10 ma1ch 1ha1 imposed in night and which is required by 1he
highest mass localions (leading edges and primary spars) was provided by concemnuing lamps
over !hose areas. Dividers between zones were required in some locations because of 1he wide
variation in chordwisc heat nux. Without the dividers, a thennal control problem known as cross-
talk [19.9J]may exisl. ln lhat case, inlerfercnce heating from a higher 1empcra1ure zone can prevem
a lower lemperaiUre zone from being conlrolled properly.

T he problem of application of mechanical loads during a heating s imu lation he re becomes


that of making attachments to the wi ng w ithout adding a heat s ink, which creates unwanted

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Lherm al stresses. The loading mechanism s hould also be designed to cause o nly minimal
interference to the heating system. The solution adopted was to apply the mechanical load
d irectly to the wi ng tes t s tructure tlu-oug h a 0.75-in. ( 19-mm) wide, 0.040-in. (1 - mm) thick
4130 steel corrugated strip a t the midspan rib location (see Figure 19.35b and [ 19.99]).

This attachment caused a minimum heal sink in comac1 wilh lhc surface of lhe teSI SLructure. Rivets
were removed from lhe wing. and the con11ga1ed SU'ip was fas1ened wilh blind rive1s through the
rcsuhing holes. Cables were auached 10 the tops of the COITugalions and passed through lhe reflec10r
with a minimum of aheration 10 1he heal ing sys1em. The cables were thermally insulmed 10 prevem
lhem from becoming overheated as they passed bclwecn lhc heat lamps. No deuimenlal elTects
were observed in lhe flight healing profile on any part of the sLruclure resuhing from lhe loading
syslem auachmem. Figure [ 19.35aj shows lhe three hydraulic acwa1ors. !heir load cells, and whip·
plelree arrangement used lo apply equal loads to each of the tesl panel auachmem cables. Each
actua10r was conu-olled by a closed-loop system lhat used load measuremems indicaled by a load
cell as feedback. All three ac1ua1ors were operated by a single programmed loading ti me hislory.

A typical arrangement of the IR heater assembly is shown in more detail for the X-15
horizontal s tabilizer in Figure 19.36 (see a lso [ 19.91)).

Figure 19.33 X- 15 wing SlruciUre (from !19.93])


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1576 Thermal Buckling and Creep Buckling

(a)

Figure 19.34 Heater system zoning for X-15 wing IR heating tests carried out at NASA Dryden (from
[19.98]): (a) plan view of wing and zoning, (b) root station structure and heater

The simulation at a particular data event must include all significant environmental conditions
beginning from some data reference. For strain gage loads measurements, the simulation must be
started at the last available zero-load condition; for the X-15, this was just prior to B-52 [its launch
vehicle] takeoff. Thus, the wing simulation had to start at ambient conditions, cool down and cold
soak to the launch condition, and then proceed to the X-15 flight heating profile.
A recirculating cooling system was constructed to provide the cool-down and cold-soak portion
of the simulation. During a particular test, the cooler mixes ambient air with liquid nitrogen (LN 2 ),
which evaporates and extracts the heat of vaporization from the ambient air, cooling it to the desired
temperature. This air is then directed to the test specimen. The upper and lower IR heater reflectors
were used as part of the cooling system enclosure.

The second example is the YF-12 airplane. A program consisting of flight and laboratory
tests was conducted at NASA Dryden in the early seventies to measure flight loads on a
YF-12A airplane.
During the flight tests, temperatures throughout the structure of the airframe were recorded con-
currently with strain gage bridge measurements on the wing, fuselage, and control surfaces. The
laboratory tests had two parts. First, a loading calibration of the strain gage bridges was performed.
Second, thermal calibrations [discussed in the following paragraphsj were performed, which sim-
ulated the measured skin and structural temperature profiles of flights at Mach 2.5, 2.75, and 3.0.
The heating control performance during the Mach 3 simulation was excellent, with a control error
standard deviation for all channels of 5.6°F. The temperatures resulting from the heating tests were
compared with many flight measurements throughout the aircraft. Those comparisons showed the
simulation of temperatures on the substructure thermocouples at strain gage locations to be good.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
High-Temperature Testing 1577

(a)

(b)

Figure 19.35 NASA Dryden X· l5 wing IR heating tests (coUJ1csy of NASA Dryden Flight Research
Center): (a) combined heating and mechanical loading test setup. (b) wing loading at·
tachme nl

Figure 19.36 NASA Dryden X·l5 horiwntal stabi lizer IR beating tests-heater assembly for heating
simulation (COLII1esy of NASA Dryden Flight Research Center)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1578 Thermal Buckling and Creep Buckling

[The YF-12 airplane had a large surface area, about 5000 ft 2 (~465 m'), which required zoning
of the reflectors.] Since the safety aspects of the heating made quick access to the airplane nec-
essary, the heater support structure was divided into sections, as shown in [Figure] 19.37. There
were two forebody sections, one for the left side and one for the right; four central heaters, a top
and a bottom heater on each side; and a rear heater that covered top and bottom on both sides.
The forebody sections and the lower central sections were mounted on V-grooved wheels that ran
on inverted angle track and retracted sideways. To clear the nacelles, the inboard ends of the lower
central sections had to be lowered by hydraulic jacks built into the frames, to a position near the
floor before retraction. Separate main wheel well heaters were built to allow the lower central
heaters to be retracted independently. The upper central sections were supported by towers and guy
cables. Each section was counterbalanced about a hinge point at the towers and retracted upward.
The rear section was mounted on V-grooved wheels and retracted rearward. The purpose of using
V-grooved wheels and angle track was to make sure that the sections would return to their original
position after being retracted, making it unnecessary to readjust the reflectors. Stops were perma-
nently fastened to the tracks as an additional measure to insure proper positioning. The guy cables
served a similar purpose for the upper central sections.
Figure [19.38] is an overhead view of the immense IR heater system and all of the electrical
wiring. The reflectors for the heaters were made of type 310 stainless-steel sheets that were 0.075
in. (1.9 mm) thick. The sheets were designed to fit the contour of the airplane 6 in. (152 mm)
away from the skin surface. The only forming operations to match the surface contours of the
airplane were rolling and bending; however, to produce conical shapes the rolling operation had to
be performed with different radii at each end of the panel. Gaps were left between each panel to
allow for thermal expansion and to permit minor adjustments to be made after the panels were
installed. The gap size, which was approximately 0.5 in. (13 mm), was determined by assuming a
panel temperature of 900oF (~490°C) and calculating the amount of growth from the coefficient
of thermal expansion for the material. A total of 444 panels and 15,870 quartz lamps were used
in this heater system, excluding the internal nacelle heater [described below] .
. . . Thermocouples used for temperature control of the heater system were attached to the outer
skin surfaces of the aircraft; and the entire vehicle was given a fresh coat of high-emissivity black
paint during test preparations.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

[Further to the primary division into support sections (shown in Figure 19.37), the heaters were
subdivided into 450 zones; half of these] were located on the left side of the airplane, the other
half were positioned at the identical location on the right side. Since the airplane is symmetrical,
zones in identical positions on each side of the airplane required the same heat input. Consequently,
control thermocouples were installed only on the left side of the airplane, and data from these
thermocouples were used to compute the power required for the entire zone. For safety reasons,
29 monitor thermocouples were located on the right side of the airplane in selected locations. A
typical zone contained 24 12-in. (305 mm) lamps and 8 18-in. (457 mm) lamps and covered an

Area@

Figure 19.37 NASA Dryden YF-12 heating tests-primary zoning areas and heater panel configura-
tions (from [19.98])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
High- Temperature Testing 1579

Figure 19.38 NASA Dryden YF-12 heating tests-overhead view of lR heater system and its electrical
wiring (counesy of NASA Dryden Flight Rescanch Center)

area of approximately 12 fl2 (- 1.1 m 2). TI1e power required for the total heating system with all
zones o perating at fu ll power was estimated to be 18 MW.

An internal nacelle heater was designed to account for the effects of stmctural heating
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

from the left-hand engine. It was constructed of 20 stainless steel panels that were rolled to
form a cylinder; which carried a total of 560 quartz lamps, divided into 20 zones of equal
size, 10 zones on the inboard side of the nacelle and I 0 zones on the outboard side. A
special installation procedure was set for these heaters (sec [ 19.98)).
The entire IR heater system consisted therefore of I6,430 quartz lamps and 464 reflector
panels, divided into 470 control zones, which were each progranuned to follow a designated
temperature-time history.
''Although the system was capable of deliveting over 18 MW of power. the maximum
power used for a Mach 3 flight simu lation was 3.5 MW. The relatively low power usage
was primarily the result of a much higher heater efficiency than design calcu lation indicated."
Furthermore, in addition to the good simulation of the Hight measured temperatures obtained.
the aircraft retumed to flight status after these tests, and performed many research flights in
subsequent years.
These two NASA Dryden applications of TR heating have been discussed in considerable
detail since they represent modern practice and emphasize the problems likely to be en-
countered.
As mentioned in S ubsection 19. I. I, simi lar and larger IR heating tests were carried out
at WPAFB in the same period. Figure I 9.39 shows a recent thermal structural test in oper-
ation at the WLFDD high-temperature facility. Though not all the detai ls can be discem ed
in thi s photograph, the size and capabi lity of this 55 MW test faci lity is apparent.
At the beginn ing of this subsection the different means of heating for ground testing were
enumerated, followed by a detailed discourse on the presently preferred IR rad iant heating.
The other means will now be brieHy discussed.
T he early direct simul ation of aerodynam ic heating to stmctural components by hjgh-
tcmperaturc wind tunnels was represented. for example, by the NASA Langley 9 x 6-ft
Thermal Structures Tunnel, mentioned in the previous subsection (see Figure 19.21). Though
the tme-temperature blowdown tunnel approach (like this NACA 9 x 6-ft, Mach 3, TST. or
its successor the NASA 8-ft round, Mach 7. High-Temperature Stmctures Tunnel, the HTST),
is the most realistic si mulation of aerodynamic heating, it is also expensive and has severe
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1580 Thenna/ Buckling and Creep Buckling

FiRure 19.39 U.S. Air Force Wright Laboratory Flight Dynamics Directorate (WLFDD). Aerospace
111ennal Structures Tests Facility- a large IR heating and mechanical loading tests in
operation (counesy of Dr. E.R. Anselmo. WPAFB)

limitations on specimen si.-:c and heating duration. For higher Mach numbers these limitations
and related problems become even more dil1icuh. lienee the high-temperature wind tunnel
is nowadays not a widely used testing tool for primary structural components. and among
the many hypersonic tunnels in operation, very few arc applied to thcrmostructural investi-
gations except for ablation testing (see for example 119. 100]).
Hot wind tunnels and other hypervelocity facilities arc. however, used ciTcctivcly to study
local high-temperature phenomena. such as themml ~hock. Among these facilitie~ arc hotshot
tunnel~ and plasma arc tunnels. Hotshot tunnels arc shon duration tunnels in which the high
temperatures and pressures required for operation arc obtained by rapidly discharging a large
amount of electrical energy into an enclosed small volume of pressuri1ed air. which then,
after rupture of a diaphragm. expands through a no1.1.lc and test sectio n (sec for example
[1 9. 1001 and 119. 101]). The characteri stics of a typical ho tshot tunnel are Mach numbers
ncar 20, stagnation temperatures of 3500-7500°F (- 1950-4170°C) and runs of abo ut 50
mscc. and many arc in operation in industry. rese:uch centers and university laboratories.
Plasma arc runnels use a high-current electric arc (e mployi ng either direct or alternating
current) to heat a test gas to a very high temperature. up to 25.000°F (- 13.9()()0C). The
electric arc raises the gas temperature to an ionization level. at which the ga;, changes to the
state of plasma. The hot pla,ma from the arc chamber then flows via a noule to a low-
density test chamber (see for example [19.100)). These tunnels may be operated for periods
of time up to many minutes. This and the hig h heating rate~ that can be developed (about
300 Btu/f! 2 sec. or 3400 kW /m 2 ) make them sui table for surface material ablation tests.
Ma ny plasma arc tunnels have been bui lt at industry labo ratories and uni vers ity research
centers in several count ries.
Graphite heaters (sec for instance [19.102] and 119. 1031). as mentioned in the previous
subsection. were developed to produce very high heat fluxes (up to about th ree times that
of quartz lamps) at temperatures up to 4000°F (- 2200"C). They have many ~im ilar features
to quanz lamps but requi re a low-oxygen environment to avoid oxidation of the graphite
filaments. They respond more slowly than quart.-: heater. and require more electric power.
However, their high heat fluxes. of about 300 Btu/ft 2 sec (- 3400 k\V/m 2 ). si milar to those
obtained in plasma arc tunnels. discussed above. but with broader versatility of application,
si ngle g raphite heaters out as useful means for obtaini ng very high heating rates in ther-
mostructural tests.
--`,`,`````,`,````,,``,``,,`-`-`,

Arc lamp heaters provide much higher heat fl uxes. of up to 2600 Btu/ft2 sec ( - 29.500
kW / m ' ) over small speci mens. with an area of say 0.2 x 4 in. (5 x 100 mm). They are,
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
High-Temperature Testing 1581

however. much more expensive than quartz heaters or even graphite heaters. Since arc lamp
heaters are also limited to heating very sma ll areas, they are only employed for special
purposes requiring simulation of extremely high localized temperat ures (see for example
[19.102]).
A recent example of use of arc lamp heaters is the Yo1tek Arc Lamp Heating Facility,
developed by Vortek Industries, Vancouver, B.C., for the U.S. Air Force WLFDD high-
temperature laboratory (sec Figures 19.40-19.42 and [ 19. I04]). The facility provides con-
tinuous high-i ntensity radiation for thermal testing. It utilizes an advanced high-energy
plasma arc heat source (shown schematically in Figure 19.41), which cons ists of an argon
gas mixture plasma arc in a quartz tube, surrou nded by a spiraling gas column and a spiraling
water wall. TI1e radiation from the arc lamp is focused with gold clad reflectors. With an
ellipsoidal reflector, the smaller 300 kW System 1 (Figure 19.40) achieves a heating rate of
14,000 kW/m' (- 1230 Btu/ ft2 sec) over an area of 800 mm' (- 1.24 in.'). With a gold
electroplated reflector, the larger 600 kW System II (Figure 19.42) achieves a heating rate
of 30.000 kW / m2 ( - 2640 Btulft 2 sec) over an area of 800 nun' ( - 1.24 in. 2) . These heating
rates are 12- 26 times that of the usual quartz lamps. At present, however, the cost of the
Vottek arc lamp system is quite high, about 10 times that of an equivalent system employing
tungsten filament quartz lR lamps (see [ 19.110]).
Another means of providing very high but very localized heating fl uxes is lasers. wh ich
are sometimes used for special cases.
On the other hand, electromagnetic induction healing, as pointed out in the previous
subsection. has the potential for very high heating rates in a broad spectrum of appl ications,
but it also entails some serious problems.
Induction heating has been widely applied for many decades to a variety of industrial
processes, such as melting furnaces; through-heating for forging, form ing and annealing;
surface hardeni ng and soldering, brazing and tube weldi ng (see for example [19. 105)).
For simulation of aerodynamic heat ing on thin-walled structures, high-frequr.ncy systems
(at 50 kHz 10 I 0 MHz, usually called radio frequencies) are generally suitable. They have
very shallow current depth (0. 1-2 mm), fast heating rates and power levels up to I MW.
Though favored in the fifties (sec for example [ 19.39). referring 10 the 20 kW electromagnetic
induction heater of the Polytechnic Institute of Brooklyn), induction heati ng is nowadays an
infrequently employed alternative method. used mainly, but not only, for material investi -
gations. The experimental thermal buckl ing studies at the French CEA Centre d'etudes nu -
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 19.4() U.S. Air Force WLFDD Aerospace Structures Tests Faci li lics, Vanek Arc Lamp Healing
Facili ly- Vortck lamp head mounted in a parabolic rcOccJor (courtesy of Dr. E.R. An-
selmo. \VPAFB)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1582 Thenna/ Buckling and Creep Buckling

Fi~:urc 19.41 U.S. Air Force WLI0 DD Vonck Arc Lamp Heming Facility- schemat ic diagram of Vonck
lamp head opemtions (courtesy of Dr. E.R. Anselmo, WPAFB)

cleaires de Saclay are. however. an example of fairly recent thermostructumltests. in which


induction coils are employed 10 produce a very strong axial thermal gradient in cylindrical
;hell~ (see 119.88} and [19.106)).
Convective heating. another means of heating simulation. applies primarily to lower tem-
peratures and larger structuml components. A remarkable usc of convective heating was the
usc of hot air to heat full -sca le components of the Concord supersonic transport (sec [ 19. 107]
or [19.103)). The convect ive heating system was also employed to cool fuselage sections of
the aircraft 10 complete the clcsirccl thennal cycle for fatigue tests. S ince both heating and
cooling is needed in the thermal cycle. economic; c:m be achieved if the same method is
used for both heating Hnd cooling. The main advantage of such a convective heating system
i> that it is necessary to control only the inlet tempcr:uure and mass flow of the hot or cool
air over the specimen in order to simulate the correct heat transfer conditions.
For larger structural components such a simplified com·ective heating system would be
even more attractive because it requires only a few controllers. compared to the many hun-
dreds of control points needed in an equivalent quart£ lamp radiant heating system.
The Concord convective heat ing and coo ling system (sec [19.107]) used effectively a
closed-circu it wind tunnel around the specimen (a section of the fuselage) wit h two bypass
systems that can be altemately connected to this drcuit. Through one of these. air heated
by ga.~ heaters to 1600C (3200F) was passed into the main circuit during the acceleration

Fi~:ure 19.42 U.S. Air Force WLFDD Vortck Arc Lamp Heating Facility-Vonek System II nrc lamp
head mounled in gold electroplated rellector (counc,y of Dr. E.R. Anselmo. WPAFB)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
High-Temperature Testing 1583

phase of the flight cycle. In the cooli ng phase, this hot air was replaced by ai r, cooled by
injection of liquid nitrogen, introduced through the other bypass.
A more recent application of convective healing is the USAF WLFDD Convection Heating
Test Facility (see Figure 19.43 and [ 19.104)). This faci lity is designed to lest structures
requiring heating and cooling resulting from air flow. It is currently configured for assessment
of full-scale aircraft transparencies. Tests are performed in the facility by controlling the
external transparency temperature. the interior air temperature and exterior air pressure. T he
test faci lity offers full-scale fl ight mission profile simu lation of heat and pressure for trans-
parencies ranging in Sil.e from F-16 to B- I aircraft. The total temperature range of this facility
is - 100 to IOOOoF ( - 73 to 540°C) and its pressure range is -10 to 25 psi (- -70 to 170
kN/m2 ) . The overall dimensions of the faci lity Jlre quite large, 98 fl (- 30m) long, 50ft
(- 15m) wide and 50ft (- 15m) high.
Before closing this subsection. we should mention a historical review on the development
of structural facilities at Wright-Patterson Air Force Ba.'e [ 19. 108] and a survey of the NASA
Dryden flight loads research faci lity ( 19. 109], a significant part of each being devoted to
elevated temperature testing. These reviews trace the evo lution of high-temperature testing
techniques and describe the facili ties at these two main thermal structures research centers
in the United States. The detai led presentation of the problems and experience gathered over
the years can provide some guidance for researchers and engineers setting out to develop
new advanced faci lities.
Another interesting study, likely to he of considerable value in the development of new
elevated temperature test techniques, is that carried out by Hanson and Casey of FluiDyne
E ngineering Corporation for the Air Force Wright Aeronautical Laboratories in the mid-
eighties [ 19. 110]. Its objective was to determine and evaluate the state of the art for high-
temperature fatigue and static testing of full-scale aerospace vehicle structures and structural
components, the relevant temperature range bei ng 1000-3000°F (-500- 1650°C).
The study surveyed all the existing facilities and techniques in the United States and
defined the requirements for a high-temperature test facility that would have the capability
to accommodate typical future hypersonic aerospace vehicles. Two different-sized test cham-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 19.43 U.S. Air Force WLFDD Convection Heating Test Facility, configured for testing of full-
scale aircraft transparencies, shown with a l)•pical fighter transparency installed in it
(courtesy of Dr. E.R. Anselmo, WPAFB)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1584 Thermal Buckling and Creep Buckling

bers were considered. The larger, a horizontal vessel, 100 ft diameter X 150 ft long ( ~ 30.5
m diameter X 45.7 m long), would be used for full-scale wings or a major portion of the
fuselage. The smaller test chamber, 30ft diameter x 60ft long (~9.15 m diameter X 18.3
m long), would be employed for smaller components. Both facilities proposed by the study
would have vacuum facilities for altitude simulation, cryogenic systems (liquid hydrogen
and nitrogen) for fuel temperature and load simulation, radiant heating for aerodynamic
heating simulation and a mechanical loading system for aerodynamic and inertial load sim-
ulation. The detailed requirement are tabulated in Table I of [19.110].
Based on these requirements, detailed estimates yielded a cost of more than $300 million
(in 1987 dollars) for the full-scale facility and about $70 million for the airframe component
test facility. For the larger setup a power supply of about 350 MW would be needed for
operation.

19.2.3 Measurements at High Temperatures


Temperature, strain and deflection are the three basic quantities to be measured in elevated
temperatures structural testing. Load measurements, when a combination of simulated heat-
ing and mechanical loads is applied, do not differ from those in the usual static tests. The
instrumentation for all three basic quantities should in general satisfy the following require-
ments (see [19.110]):
1. Suitability for measurements at temperatures from 1000-3000oF ( ~500-1650oC)
2. Capability of being and remaining attached to likely test article materials
3. Durability at the maximum test temperature
4. Suitably small to fit into the crowded environment close to the test vehicle or article
5. Reasonable ruggedness
6. Either no requirement for calibration after installation, or ease of calibration
7. Capability of operating at high vacuum
8. Reasonable cost
Each type of instrumentation will now be discussed separately. For temperature measure-
ment, thermocouples are the principal sensors, up to about 3000°F ( ~ 1650°C), though ra-
diation, optical or infrared pyrometers are also used. For temperatures above 3000°F
( ~ 1650°C), optical instrumentation is suitable, but it requires line-of-sight access, which will
usually not be available in the region of the specimen, which is crowded with radiant heating
devices.
Thermocouples are temperature-measuring devices based on the Seeback thermo- electric
effect (when two dissimilar metal conductors are joined at their ends but separated in be-
tween, so that the junction at one end is at a different temperature than that of the junction
at the other end, an e.m.f. is set up in the circuit and a current will flow).
In early tests the thermocouples were installed by drilling a small hole the size of the two
thermocouple wires fused together to form a junction, inserting the junction into the hole
and peening the edge. This worked well but was time-consuming and not suitable to very
thin surfaces. In the fifties a method for flash welding the thermocouple wires directly to
the specimen was developed. This improved installation was done with a small capacitance-
type welder, which supplied the necessary instantaneous current for fusing the wire to the
specimen.
Iron and constantan wires (type J thermocouples) are generally used for temperatures up
to 1000°F ( ~500°C) and Chromel-Alumel (type K) for temperatures up to 2000°F ( ~ 1100°C).
Above this temperature, noble metals are employed (for example type S, platinum and I 0
percent platinum/rhodium). Thermocouples are defined nowadays by the U.S. National In-
stitute for Standards and Technology (NIST) as seven basic types, S, R, J, T, K, E and B,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
according to their dissimilar materials joined together. The most common types used in
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Table 19.1 Strain gages used in hot-structures testing at NASA Dryden
Type Method of Temperature Temperature Signal Approximate
attachment range, oF range, oc conditioning cost, each gage
( 1990)
Foil Adhesive -320-600 (-195-315 ) DC $20
-320-1,200

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Weldable Spot welding (-195-650 ) DC $700
Capacitive Spot welding -320-1,500 ( -195-815 ) AC ($1,000) $1,500
Fe-Cr-Al-alloy Plasma-flame spray 70-1,900 ( 60-1,340) DC 20 to 200

....
~
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1586 Thermal Buckling and Creep Buckling

elevated temperature structural testing are J, K and S, and Figure 19.44 summarizes the
temperature range and associated error limits for two of them, K and S.
Temperature instrumentation is required both for temperature control of the heating system
(as pointed out in the previous subsection for the radiant heating systems described) and for
evaluation of the measured strains. In a typical radiant heating test of a full-scale vehicle or
a large component, the heaters are divided into hundreds of separately controlled zones in
order to obtain a realistic heating simulation, as pointed out in the previous subsection.
Furthermore, since the control function of the temperature instrumentation requires reliable
information throughout the entire temperature-loading cycle, redundancy of sensors is re-
quired. Hence, hundreds of thermocouples are needed for temperature control, and a similar
quantity to provide diagnostic information for strain and deflection measurements. For ex-
ample, in the full-scale high-temperature test facility planned in [ 19.110], 4000 temperature
sensors were specified, 2500 of them on the specimen surface for radiant heat control. Tem-
perature measurements are also needed in the highest-temperature regions, and therefore
some thermocouples have to be attached to carbon-carbon composites, which make up most
of the stagnation regions. Such bonding of thermocouples to composites with suitable ad-
hesives and heating to 2500°F ( ~ 1400°C) has been performed, but there are still some un-
certainties concerning possible chemical reactions between the thermocouple material and
carbon at high temperature.
There is another type of widely used temperature sensor, related to electrical resistance
changes with temperature changes. These thermal resistors, or thermistors, are semiconduc-
tors (usually a mixture of cobalt, nickel and manganese oxides with finely divided copper)
whose resistance is very sensitive to temperature. Thermistors are available in many forms,
such as probes, rods or disks, and have extensive use in electronic systems, such as wave-
guide systems, and in temperature compensation devices, as well as for temperature mea-
surement. In elevated temperature structural testing, however, thermistors have not been
employed.
Strain instrumentation is probably the dominant measurement system in structural testing,
and considerable efforts have been devoted to extending the applicability of the principal
sensor, the electric strain gage, to high temperatures. The specific additional requirements
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
for suitability of strain gages to elevated temperature structural testing are (see [19.110]):
1. Maximum strain approximately 5000 microstrains (0.5 percent)
2. Relatively small variation of gage factor with temperature
3. Only slight zero shift with temperature
Nowadays, the strain gages for hot-structures testing satisfy most of these specific as well
as the general requirements. For example, Table 19.1 (reproduced from [19.97]) shows in-
formation on the strain gages used for elevated temperature structural testing at the NASA
Dryden Flight Research Facility. The table includes the method employed for attachment of

20
Type K, standard

15
limits of
temperature 1o
error,
"F

0 500 1000 1500 2000 2500 3000


Indicated temperature, oF

Figure 19.44 Thermocouple specifications, temperature range and associated error limits for two com-
monly used types K (chromel-alumel) and S (platinum and 10 percent platinum/rho-
dium), (from [19.97])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
High-Temperature Testing 1587

the gages, their operational temperature range and the signal conditioning used. The ap-
proximate cost in 1990 of each gage is also shown, as it may influence the selection. One
may notice that the high-temperature gages are still quite expensive. Extensive experience
has been gained at NASA and elsewhere in the employment of foil, weldable and capacitance
gages over a wide range of temperatures. The fourth type, the iron-chromium-aluminum (Fe-
Cr-Al) alloy high-temperature gages, have only been introduced in recent years and warrant
further testing.
Despite ongoing research, no practical strain sensor for temperatures up to 3000°F
( -1650°C) has yet materialized. Some experimental high-temperature gages, other than the
Fe-Cr-Al alloy ones, have been developed, like the Boeing/Hitec capacitance gage for up
to 1500oF (-8l5°C) and the Battelle-Columbus free filament gage up to 2000°F (-ll00°C),
but they still require excessive care in application (see [19.110J). Hence, further research is
needed in this field.
Returning to the well-tried high-temperature gages extensively used at NASA Dryden and
elsewhere (the first three in Table 19.1), one should remember that their accuracy when
applied to a hot structure is dependent upon several factors. These include the gage factor
and its change with temperature, the drift of the gage and the apparent strain characteristics
of the gage when attached to test specimen. Following De Angelis 's and Fields's presentation
in [19.97], it is noted that "satisfactory gage factor and drift characteristics have been dem-
onstrated for the foil, weldable, and capacitive gages." For the fourth type, the Fe-Cr-Al
alloy gages, only limited data are available, but it too "suggests satisfactory gage factor and
draft characteristics."
Hence the apparent strain, various strain gage outputs due to temperature, become the
gage characteristic that has the dominant influence on the accuracy of strain measurement
at high temperature. Several techniques have been used over the years to reduce these tem-
perature effects. such as temperature-compensated strain gages, special strain gage configu-
rations, and the wiring of a number of strain gages into Wheatstone bridges. But often these
techniques are insufficient, since the apparent strains can be larger than the intended strain
measurements. Therefore extensive calibration tests had to be resorted to (see for example
[19.91]-[19.93] and [19.99]).
For resistance strain gages, the apparent strain "is a function of the difference between
the coefficient of thermal expansion of the gage and the material to which is attached, and
the change in resistivity of the gage with change in temperature." The capacitive strain gage
(see Figure 19.45), on the other hand, has no significant apparent strain, because of its design.
Here the rod material and the material of the test specimen are the same, and therefore no
apparent strain corrections should be needed. However, since the rod is positioned somewhat
off the surface of the test specimen, there is usually a difference in temperature between the
rod and that surface, which can produce a measurement error. Hence, two thermocouples,
one attached to the rod and one to the surface directly below it, are usually added to provide
the temperature difference required for correction of the strain data.

Outer capacitor plate


Inner capacitor plate

Rod material same as material -""'


of specimen under test ""-

Nominal gage length: 2.54 em (1 ln.)

Figure 19.45 High-temperature capacitive strain gage, suitable for temperatures up to 1500oF ( ~815°C)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(from [19.97])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1588 Thermal Buckling and Creep Buckling

Prior 10 an elevated temperature test on any structure instrumented with strain gages, an
apparent strain test has to be carried oul. " For foi l gages, these tesL~ consist of mounting
several strain gages from a common batch on a sample material of the test article. The
sample material or cOuJXln is then heated slowly to achieve a thermal-stress-free temperature
rise. after which the coupon is allowed to cool slowly. Several cycles arc run and the data
arc collected and analyzed ." Below, a typical apparent strain test setup is described (see
Figures 19.50a and 19.51a).
Typical apparent strai n curves for different types of resistance gages, obtained in such
tests, are shown in Figure 19.46. T he apparent strain curves, from room temperature to SOO"F
(- 260"C), for eight foi l gages attached to a titanium coupon are shown in Figure 19.46a.
As seen in the figure. the foil strain gages exhibited little variat ion in apparent strain among the
eight sensors. The apparent strain variation with temperatures seen in these tests cotTesponded
closely 10 the manufacturer's prediction.
Weldable strain gages also tend to generate repeatable apparent strain curves for multiple thcnnal
cycles. However, the curves tend to vary widely from gage to gage. as shown in [Figure 19.46bl
by the range of apparent strain for 24 weldable strain gages. The figure also depicts a method to

Apparent
strain
~in.lln.

(a)

~
~=
Weld1bl t strain 9191
A.ppller'll
t iHIIn,
J"in.lln.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

W~.":!':... Tomo"''"''· •F
Clt.mped

-::~~·
pln.lin.
':[:==-· -----:-=;---
500
. L
WetdecsJ
I _
<::::::
J
- 0 200 400 600 BOO 1000 1200
C~mplng Oo~<e TempefehJie, •F
(b)

-- tat cycle htat.•l.4)


.... ... • I at cyelt cOOI·down & 2nd cyclt fl. .t.up
--- 2nd cyc:lt c:ool·down 6 3rd cyelt heat-up
-··- 3td cycle eool-down

S ltaln,
IJinftn.

·•000 L:=:L~.J__.J__.J.__
0 . ..
(c) Tt mptrtture, "F

Figure 19.46 Typical apparent strain curves, obtained in specific apparent strain tests. fo r three types
of resistance gages (from r19.97)): (a) foil gages, (b) weldable gages. showing also a
typical gage and a clamping device. (c) Fe-Cr-AI alloy gages

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
High-Temperature Testing 1589

establish an apparent strain curve for each gage before it is welded to the test article. A clamping
device is manufactured from the same material as that to which the gage will be attached. Then,
using the clamping device as a test coupon, apparent strain curves are obtained for each gage to
be used. This approach may seem to be tedious and costly, but it is necessary to obtain accurate
strain measurements when using these gages at high temperatures.

The apparent strain values measured for the Fe-Cr-Al alloy gages are significantly larger,
as can be seen in Figure 19.46c. The magnitude of the apparent strain values for these strain
gages varies according to the material to which they are attached. However, the characteristic
behavior of the apparent strain with temperature, shown in Figure l9.46c, is the same.
[The] figure shows the apparent strain for three heat-up and cool-down cycles over a temperature
range from room temperature to 1900°F (~1040°C). The gages were mounted on an Inconel 601
bar (a nickel-based heat resistance alloy). As seen in the figure, the second cycle heat-up traces
the first cycle cool-down and the third cycle heat-up traces the second cycle cool-down. This
characteristic exists because the strain at the maximum temperature is invariant, and essentially all
metallurgical phase transformations occur during the cool-down portion of the cycle below about
1200°F ( ~650°C). Furthermore, as more cycles arc run, the data for both heat-up and cool-down
cycles converge into one curve .... The large apparent strain values suggest that generating apparent
strain data from coupon tests to correct data from gages installed on a test panel is not a valid
procedure [in this case].

Hence, a more accurate procedure would be to generate individual apparent strain curves
from each gage as installed on a test structure, but the requirement of such a procedure
would restrict the scope of the test programs.
However, as pointed out in [19.97], there are cases where the apparent strain is not a
governing factor, and then these gages are appropriate. For example, Fe-Cr-Al alloy strain
gages were used at NASA Dryden in a force-stiffness technique for measuring the buckling
load on a buckling-critical panel (sec Subsection I 5.3.2 of Chapter 15 for details of this
technique). In this application, mechanical loads were applied after the panel was stabilized
at an elevated temperature. This meant that the temperatures were uniform and therefore the
apparent strains were not a factor, as they represented a uniform reference strain.
The calibration of strain gages attached to structures subjected to elevated temperatures
has been extensively investigated and still involves considerable test efforts, as indicated
above (see also [19.97] and [19.99]). In slowly varying temperature environments, the tech-
niques used to correct strain gage enors are generally well established and reliable. However,
in hypersonic or transatmospheric vehicles, highly transient temperature conditions are en-
countered that may pose additional calibration problems (see for example [19.111]).
A test component of such a vehicle, instrumented with a bonded electrical-resistance strain
gage, is shown in Figure 19.47 (reproduced from [19.1 I 1]).
As heating rates applied to a strain-gage installation increase. temperature differences between the
strain gage and the substrate will increase according to Fourier's law. These differences shown in
[Figure 19.47] become increasingly significant because the materials used to insulate gages elec-
trically from the substrate are usually good thermal insulators as well. The lower the thermal
conductivity of the strain-gage insulating material, the greater the temperature difference through
the strain-gage installation (for the same heat flux imposed).
Conventional strain-correction procedures currently neglect temperature gradients by assuming
that these temperature differences are insignificant and, therefore, do not adversely affect strain-
gage measurement accuracy. This assumption may be valid for slowly varying temperature envi-
ronments, but it is uncertain at what heating or cooling condition this assumption becomes inap-
propriate.

Therefore, the conventional strain-correction theory has been modified to account for the
additional errors introduced in transient-temperature environments, and a new experimental
method has been developed at NASA Dryden to correct strain-gage measurements in the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1590 Thermal Buckling and Creep Buckling

HMt tiUJI

I
Tem.,.r.ture

,..___..,_L diffe~I'ICe

Figure 19.47 Strain gage installation on a typical hypersonic vehicle test component (from [ 19.111 ])

presence of signi ficant temperature transients (sec [ 19. 11 I )). The discussion here follows that
reference.
In order to facilitate the presentation of the new approach, the conventional strain correc-
tion procedure is first recapitulated. The apparent strain error is graphical ly illustrated in
Figure 19.48. " The original position of a strain gage and substrate is shown at point A
[Figure 19.48a]. When an instntmented test article is subjected to a uniform temperature
change of the substrate from an ini tial reference temperature !lT,. the strai n gage and the

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

f igure 19.48 Conceptual illustrJtion of apparent strain (from [19.111]): (a) gage and substrate at am-
bien! condilion. (b) free thennal expansion of gage and subslrate. (c) subslrate-induced
stmin on gage, (d) final state of apparent strain

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
High-Temperature Testing 1591

substrate will auempl to expand o r contrac t by an amount corresponding to their coefficients


of thennal expansion."
The thermal expans ion of the gage and s ubs trate, if both arc heated and allowed to expand
freely in the positi ve X direction, is s hown in Fig ure 19.48b.
The gage expands to point B and the substrate expands to point C. Because of its greater Stifl'ness,
the substrate will force the gage to conform to its expanded position, as shown at point C [Figure
19.48c]. Any local stiffening of the substrate by the gage is neglected. An additional strain of
(a, - a. J6T,. that is not representative of the stress state in the test a11icle. is therefore applied tO
the gage.
Rcla1ivc gage-tcrnpcralurc changes also produce changes in 1he 1emperature coefficient of resis-
tivily. ')',and 1he gage fae10r. GF [see Figures 19.48b and c). [Figure 19.48d) shows the final stress
condition of the gage after responding to free thennal expansion of lhe subs1ra1e."
Apparent strain e"'"" as shown graphically in Figure 19 .48d, canno t be accurately ca lculated
because considerable variation ex ists with the temperarure dependence of the gage prope1t ies
from lot to lo t. The apparent s train tests, described above, are there fore re lied upon to achieve
an accurate correction for the e,,.
c haracteristics of the gage.
The new approach [ 19 .111] accounts for the strain e rror produced in transient temperature
e nvi ro nments. The transient temperature strai n error can also be illustrated g raphical ly (see
Fig ure 19.49) in a manner s imi lar to that of the apparent strain.
Figure [19.49a) shows lhe same gage and substrate ins1allation as [Figure 19.48d), with lhe strain
gage subjected 10 the same e,•• error as before. [Here il is assumed that 1he installed gage is
exposed 10 a highly transient heating profile that produces a positive 6T, ,]. In this case. only 1he
gage will respond to the temperature increase by anempting to expand from point C 10 point D.
Because the gage and substrate are bonded together. 1hc substrate will prevent the gage from
expanding.
Figure [19.49b] shows 1ha1. in this example, a compressive strain - apT,, is sensed by the strain
gage 1hat is not a result of e. in the subs1ra1e. Gage sens itivity 10 strain also changes with 6T,,

bAT.. •I --1
c 0
~<;. AT8,

==]
1 -·

• • -~ 4r ,
0

'+·(0:.-"t)•'oo

i I

Figure 19.49 Conceptual illustration of the transient-temperature Slrain en·or (from [ 19.11 I]): (a) free
thennal expansion of gage due to tempermure difference between gage and subs1rate, (b)
substrate-induced strain on gage. (c) final stress state of gage under transienl heating
conditions
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1592 Thermal Buckling and Creep Buckling

and further contributes to the s 1 . Figure [19.49c] shows the enor and the final equilibrium positions
of the gage and the substrate.
Summarizing, the corrected stress-induced strain e" is related to the indicated strain by

(19.6)
where earr is the apparent strain, obtained from apparent strain tests, and e 1 is the transient
temperature strain error. By combining all the temperature-induced strains into a combined
strain error due to temperature
(19.7)
Eq. (19.6) can be written as
(19.8)
Equations (19.6) and (19.8) represent two alternative formulations, which both require
measured results for calculation of the corrected stress-induced strain. Details of the com-
putational procedures are presented in [ 19 .111].
A series of tests was conducted at the NASA Dryden Flight Research Center to demon-
strate the correction theory and experimental procedures. Two ovens were required for heat-
ing tests: one for low temperature-rise rates to provide slow, uniform heating (the conven-
tional apparent strain tests) and a second capable of achieving temperature-rise rates greater
than 100°F/sec (~56°C/sec) to simulate highly transient temperature conditions.
A ceramic oven (see Figure 19.50a) was used for the low heating rates of less than 20°F/
sec ( ~ ll oc/sec), this being a typical apparent strain test setup. "Heat was applied to both
sides of the test coupon by 42 quartz infrared lamps (21 per side) spaced 1.25 in ( ~32 mm)
apart. Each lamp provided a maximum of 2500W and 440V. All surfaces of the oven were
ceramic. Steel baffles were inserted between the coupon and heating elements for the 0.3°F/
sec (~0.17oC/sec) test (the test with the lowest heating rate, which served as the baseline)
to diffuse heat uniformly to the coupon."
For transient heating tests, the researchers at NASA Dryden used an oven, capable of high
temperature-rise rates (20-100°F/sec, ~ ll-56°C/sec), and shown in Figure 19.50b. In that
oven, as can be seen in the figure, "the test coupon was mounted vertically on an insulated
stainless steel frame to provide stability and prevent lead wire damage during heating." The
accurate orientation of the coupon between the two banks of parallel lamps was facilitated
by the maneuverability of the frame. "Each side of the coupon was heated by 14 quartz
lamps spaced 1 in. (25.4 mm) apart. Each lamp was capable of producing power greater
than 6000W at 440V (double the rated voltage). Water/glycol-cooled aluminum reflectors
were used to reflect radiant heat to the coupon surfaces. The oven was enclosed with 1.5 in.
( ~ 38 mm) thick ceramic blocks, each lined with 0.00 I in. ( ~0.025 mm) thick nickel foil to
improve oven wall reflectivity."
The test coupon (see Figure 19.5la) was made of titanium (Ti 5Al-2.5 Sn) and measured
3 X 50 X 25 in. ( ~ 76 X 127 X 6.4 mm). It was first used in the usual apparent strain test,
and then in the transient-heating tests. For the earr tests, the coupon was instrumented with
30 spot-welded type K thermocouples (half on the top surface of the coupon and half at the
same locations on the bottom surface) and 6 foil strain gages, arranged in two rectangular
rosettes in the middle of the coupon (one of each surface), as shown in Figure 19.51 a. This
instrumentation is typical of the usual isothermal apparent strain tests.
For the transient temperature tests, an indication of the gage temperature is required, in
addition to the surface temperature measured by the 30 spot-welded thermocouples. Here
the gage temperature was represented by a commercially available type K foil thermocouple
attached to the substrate with the same adhesive and technique as the foil strain gage. The
--`,`,`````,`,````,,``,``,,`-`-`,,`

foil thermocouples were assumed to measure the gage temperature !1T" because of its sim-
ilarity in materials and cross-sectional dimensions to the strain gage (see Figure 19.5lb).
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
High-Temperature Testing 1593

r 25 ln., 2500.W,

rT
0 0
0
quartz lam.pa 0
o (211alde) 0
0 0
0 0

~ \':i
0
0
0
0
251n. 0 0
1 .,..
311n.
...
0
0
0
0

1
0
0
0

·~r ::1
0
0
141n.

1 1.2Sin.,
typlc.lly
0
0
0
0

31n.

t IUin.

(a)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 19.50 NASA Dryden strain gage correction tests-test setups (from [19.11 1]): (a) diagram of
low-heating rate oven. (b) diagram of high-heating rate oven

Two foi l lhermocouples were inslalled in the cen1er of the coupon (one on each surface). A
simple one-dimensional lhermal analysis showed thai the difference be1ween the foi l s1rain
gage and foi l lhermocouple temperatures for a gi,•en heat flux should indeed be less lhan 5
percent The difference belween the lemperatures measured by lhc foil and spo1-welded
thermocouples could there fore be used to define the aT,. term in e, (see Figure 19.49b).
After all lhe sensors were auached. the instrumenlecl coupon was pai nled with a high·
emiuance painL, which ensured the application of a more uniform heat flux onto the coupon
surface and improved lhe radiative heating efficiency.
In tests with temperature rises bel ween 10 and IOO' FI sec ( - 6- 56' C I sec) significant errors
due 10 transient healing were observed. oflen of simi lar magnitude to the apparenl strain
errors. Comparison of lests resu lts with a finile-difference analysis (using STNDA '85/
FLUINT; see (19.111 (), showed belter correlation between analysis and tesl resuhs corrected
with the new method (which takes the transient temperalure strai n error into accou nt) than
when lhey were corrected with the convenlional method (which neglects the lransienL tem-
perature effects). This was noticeable in particu lar for the higher healing ra1es (see for ex-
ample Figure 19.52, for a healing rale of IOO' F /sec ( - 56°C/sec).
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1594 Thermal Buckling and Creep Buckling

3 1n. •I

(b)

Figu re 19.51 NASA Dryden strain gage correction tests- test coupon and instnmJentation (from
f 19. 111 ]): (a) location of thcnnocouples and str.tin gages on coupon. (b) comparison of
foil strain gage and foil thermocouple cross-sections

This de tai led discussion of correction techniques for strain gage measurements in an el-
evated temperature environment high lights one of the pro minent instrumen tation problems
in high-temperature testing .
A nothe r example of a recent investigation of the performance of new h ig h-temperature
Slfain gages is th at carried out by the S tructures Test Branch. FDDWL at Wright-Patterson

-~ -~· · r C~nvenUon.al rn e1h0d I


0 1

.. ~

~::1---'-~\-+\:-,--
. _r--~oly~ls-1--
_ I
• • •• • · Hewmelhod I

-300 --J-\~
~- -·00~---+--~~--~--~~--~--~
- --- - 0

1J•tr1ln \

-oool---~---4~~~·~-4----+---~ ••
-··· f---~----~-~"<""
1\
'- 1---+----t----1
··-~
- 100 - - · -- ---1--- ~ - -- - - -

Figure 19.52 NASA Dryden strain gage COITect ion tests-comparison of stress-induced strain with
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
analysis at a heating rate of JOO"F/sec (- 56C/sec) (from [19. JJ I])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
High- Temperature Testing 1595

Air Force Base [19.112], in which five recently developed strain gages were compared. These
were:

1. Micro-Measurement 350 ohm Karma alloy WK-03 temperature-compensated series for


strain measurements at up to 350°F ( ~290°C), bonded with an epoxy-phenolic adhesive
2. NZ-2104 free-filament Fe-Cr-Al alloy 120 ohm strain gages, applied by the flame-spray
0
technique, for use at up to 1500°F ( ~820d C)

The above two types were the basic strain gages employed in the experiments, 100 type WK
and 80 type NZ were installed on the test article (the lightly loaded splice subcomponent).
In addition, three developmental half-bridge temperature compensating strain gages were
installed by the flame-spray technique in sets of 10 each.
3. NASA-Langley developed sensors, consisting each of two Kanthal alloy gages (a Fe-
Cr-Al alloy similar to the NZ-2104), installed side by side
4. NASA-Lewis developed half-bridge sensors, each consisting of a flame-sprayed palla-
dium-13 percent chromium 130 ohm strain-sensitive arm and a flame-sprayed 10 ohm
resistance element arm
5. Wright Laboratories half-bridge temperature compensating sensors, each consisting of
two side-by-side flame-sprayed NZ-2104 Fe-CR-Al alloy gages

Details of the tests and the behavior of the gages can be found in [19.112]. Some of the
conclusions there reconfirm the assessment of the state of the art discussed above. In partic-
ular, the following should be noted:

1. "More laboratory work needs to be done before untested strain gages are used en masse
on real structures."
2. "High temperature strain measurement above 600°F ( ~ 320°C) is still not accurate or
reliable."
3. "The temperature compensating half-bridge (dummy strain gage concept) ... still offers
the best control of undesirable thermal output."
4. "Temperature measurement for apparent strain correction needs to be very carefully
considered" (as pointed out earlier).
5. "Deflection measurement may be the only accurate method of structural response mea-
surement at high temperature for quite some time."

Deflection is the third basic quantity to be measured. Deflection instrumentation should


satisfy the general requirements specified in the U.S. Air Force-sponsored survey of high-
temperature technology [ 19.11 0] discussed above, and for large-scale structural components
deflection sensors should be capable of measuring deflections of about 3 ft ( ~900 mm) and
deflection rates of 250 in./ sec ( ~6500 mm/ sec).
This 1987 survey found that none of the available deflection sensors could operate in the
required high-temperature environment of up to 3000°F ( ~ 1650°C). Therefore, cooled shields
are necessary for the deflection sensors, with an attachment to the specimen, which may, but
need not, be cooled as well. For example, cables or low-coefficient expansion rods, such as
quartz or lnvar, have been successfully used as deflection connectors.
The sensors generally used for deflection measurement in high-temperature structural test-
ing (applied via cooled or low-thermal expansion coefficient rods, such as quartz or Invar,
or via piano wires) are linear variable differential transformers (LVDT's) or wire-wound
potentiometers. The resolution of the LVDT's is much better than that of the string poten-
tiometers. However, sometimes obstruction to the quartz compression rods during heating
distorts the LVDT measurements (see for example [19.112]), while the small wire diameter
of the wire-wound potentiometers permits unobstructed deflection measurements, and then
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

the potentiometers are preferable.


Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1596 Thermal Buckling and Creep Buckling

For load measurements in elevated temperature structural testing, the usual load cells are
employed, obviously outside the hot region.

19.3 Thermal Buckling


19.3. 1 Origins of Thermal Buckling
Thermal buckling is probably the most important effect of high temperatures in thin-walled
structures, besides the degradation of material properties. It was recognized in the early fifties
as the most serious problem associated with the transient aspects of aerodynamic heating
(see for example [19.113]), and was therefore the subject of many investigations in the fifties
and sixties, as mentioned in the previous section.
The restraint to thermal expansion is one of the two basic causes of thermal buckling. Its
effect can be clearly demonstrated by consideration of a uniformly heated straight slender
bar attached to immovable pin joints. The thermal stresses, due to the prevented expansion,
make the bar a column and result in a compressive force
P = EAaT (19.9)
where a is the coefficient of thermal expansion of the column, E its modulus of elasticity,
A its area of cross-section and T the temperature.
The simply supported column will buckle when Preaches the Euler load ( 7T 2El I U), where
I is the moment of inertia of the cross-section and L its length. The critical temperature is
therefore
T"' = ( 7T 2 EII F-)/ EAa (19.10)
2
= ( 7T / a)(l/ AL 2 )

As an example, consider a 400-mm long steel column with immovable pinned joints,
having a cross-section 10 X 36 mm. For steel, the linear coefficient of expansion is a =
0.000012 per oc (~0.0000067 per °F), and therefore from Eq. (19.10) the uniform temper-
ature at which the column buckles is Tcr = 42.8°C ( ~ 109°F).
The second basic cause of thermal buckling is non-uniform heating. When a uniform
structure is heated with a heating rate variable over the surface, or when a non-uniform
structure is heated uniformly or non-uniformly, the temperature varies from point to point
in the structure. To preserve the continuity of the structure, internal stresses arise, which are
known as thermal stresses.
Comparatively small temperature differences can give rise to significant thermal stresses,
as is evident from a simple example: a long rectangular thin plate subjected to a chordwise
temperature variation. The length of the plate is 2a, its width 2b and its thickness t, where
2a >> 2b >> t. The in-plane coordinates are x andy, with the origin at the center of the
plate. T = T(y) and plate stress is assumed. If the body forces and the stresses u, and T"
are assumed to be zero, u, = f (y) and the equilibrium equations are satisfied. The compat-
ibility equation reduces then to
u,_, + EaT,,., = 0 (19.11)
If the conditions of zero resultant force and moment are invoked, the solution for the thermal
stress becomes

u, = -EaT+ (112b) rl> EaTdy + (3y/2b 3


) r, EaTydy (19.12)

For a temperature distribution of the type obtained experimentally by rapid heating of leading
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Thermal Buckling 1597

and trailing edges of a flat plate, discussed in Subsection 19.2.1 (see Figure 19.18 and
[19.44]),
T = T0 + T 1(ylb) 6 (19.13)
the thermal stress obtained from Eq. (19.12) is
a:, = EaT1[(1/7) - (yi!J)6] (19.14)
The maximum tensile stress, a:unax.ten = EaT1(1 /7), occurs at the centerline (y = 0), while
the maximum compressive stress, a:,maxwmp = -EaT1(617), occurs at the leading and trailing
edges (y = ±b). For instance if T1 = lOOoC (180°F), the maximum compressive stress in a
steel plate is a:, max "'"'~' = -213 X 106 N I m 2 ( ~- 21.7 kg I mm2 or -30,850 psi), a substantial
value that can cause buckling, or even approaches yield.
The solution of Eq. (19.12) applies only to very long plates and is therefore often referred
to as infinite plate thermal stress. For finite rectangular plates the end effects have to be
accounted for, which become more important the smaller the aspect ratio of the plate. Several
methods have been employed to calculate these end effects (see for example [19.19], [19.36],
[19.48], and [19.114]-[19.117]).
In aerospace applications the thermal buckling of wings was the first concern of the
designers of supersonic aircraft as their Mach numbers exceeded 2. The early studies (see
for example [19.38] and [19.39]) dealt with the basic physical aspect of the problem: "The
coverplates of the wing are exposed to the airflow and are heated by it while the interior
elements, such as the shear webs, remain cold. The thermal expansion of the coverplates is
partially restricted by the webs, with the result that compressive stresses are produced in the
coverplates and tensile stresses in the shear webs. Under the action of the compression the
coverplates are likely to buckle." This was indeed confirmed by heating tests of a model of
a box beam in an electromagnetic induction heater at the Polytechnic Institute of Brooklyn
[19.118], and later by radiant heating tests of a box beam at NACA Langley [19.124].
The coverplate was assumed to be represented approximately by a simply supported long
plate, subjected to a thermal stress that varied across the width of the plate [19.38]. This
non-uniform stress was approximated by a uniform compressive stress, upon which was
superimposed a second harmonic of the compressive stress, whose amplitude wasp times
the average stress. The buckling stress of the long rectangular plate, of width 2b and thickness
2t, was then:

(19.15)

where the subscript c of the material constants refers to the coverplate. Equation (19 .15) was
found to be accurate enough for practical use when -2 < p < 1.
As an example, buckling was examined [19.38] for a typical steel wing, whose center
section was represented by a box wing, with coverplates 12 in. (~305 mm) wide and 0.125
in. ( ~ 3.2 mm) thick, and webs 8 in. ( ~203 mm) deep and 0.0625 in. ( ~ 1.6 mm) thick. If
the wing was subjected to a constant heating intensity q = 28.7 Btulft2 sec ( ~ 326 kW 1m2 ),
a very feasible magnitude, which would result in a temperature rise in the cover plate of
about 40aF I sec ( ~ 22°C I sec), the cover plate would buckle in less than 10 seconds. The
actual thermal buckling stress would be 10,700 psi ( ~73,800 kN 1m2 ), whereas for uniform
spanwise compression a:,, would be slightly (6.5 percent) higher, 11,400 psi ( ~78,600
kN I m2 ). This example indicates that thermal buckling of coverplates may occur after a short
time even at fairly low heating intensities. It also shows that the assumptions leading to Eq.
(19.15) are accurate enough for engineering purposes and that the elementary uniform com-
pression case represents a reasonable first approximation.
Thermal buckling of plates has been the subject of many other studies since the early
fifties (see for example [2.87], [19.19], [19.25], [19.37], [19.40] and [19.119]-[19.126]),
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1598 Thermal Buckling and Creep Buckling

though only very few of them were experimental, which will be considered in the following
sections.
For shells, the early studies examined the likelihood of buckling in cylindrical shells
subjected either to axially varying temperature distributions or to circumferential variation
(see for example [19.19], [19.25], [19.41]-[19.43] and [19.127]-[19.129]).
Cylindrical shells may experience non-uniform hoop compression, when either they are
subjected to heating with an axial temperature variation or the temperature rise is uniform
but restraint is provided by cooler bulkheads or stiffening rings against the uniform expansion
of the shell. Hoff [19.41] found that for a simply supported shell with a uniform temperature
rise that is constrained at the cooler supports, elastic thermal buckling is very unlikely. The
reason for this is the shift of the thermal stresses towards the supports, which induces per-
manent local plastic deformations long before elastic buckling can occur. Other investigators
(see for example [19.127]-[19.129]) also studied clamped ended shells and arrived at essen-
tially similar conclusions, though clamped shells were found to be more prone to buckling.
Circumferential temperature variation induces non-uniform axial thermal stresses in cylin-
drical shells that can cause buckling. Early studies (see for instance [19.42], [19.129] and
[19.130]) concluded that the critical axial thermal stress under circumferentially varying
temperature does not differ very much from the critical uniform axial compressive stress.
This conclusion was substantiated by later investigations (for example [19.131] and
[19.132]), provided that the stress variation was not abrupt, or in other words, that the
intensity of variation over one-half buckling wavelength was not large.
The very few thermal buckling experiments on cylindrical shells, carried out in the early
sixties (for example [19.127], [19.130] and [19.136]), roughly verified the theoretical pre-
dictions but were not conclusive. Hence, additional tests were performed in the late sixties
and seventies (to be discussed in Subsecrion 19.3.4), most of them considering circumfer-
entially varying temperatures, which were of greater practical significance. The interaction
between thermal and mechanical loads, which may result in buckling of the shell, were also
extensively studied in the sixties and seventies, and it too is considered in that section.
Though the focus of the discussion is on aerospace applications, on buckling caused by
aerodynamic heating, thermal buckling is also prevalent in power plants, in particular nuclear
power plants, in propulsion units and in the structural elements of the chemical industry.

19.3.2 Early Thermal Buckling Experiments on Plates and Shells


For plates, the classic 1952 Heldenfels and Roberts NACA Langley thermal stress place
benchmark test [19.36] was extended to a plate buckling test in the same year [19.37]. The
same heating and cooling arrangements were used, yielding the same tent-like temperature
distribution (of Figure 19.16a), but the deflections of the plate were now measured by nine
dial gages (see Figure 19.53a). The deflection and postbuckling behavior of the correspond-
ing perfect and imperfect simply supp01ted plate (the experimental boundary conditions
closely represented simple supports, with virtually complete freedom in the plane of the
plate) were calculated by an approximate Galerkin solution of the von Karman large-
deflection equations, modified for the effects of nonuniform temperature distributions. A
comparison of the calculated and experimental deflections at the plate center with increasing
temperature difference T0 (between center and edge temperatures), depicted in Figure 19.53b,
shows good agreement. The deflections plotted are the differences between the total center
deflections w, and the initial one w,, ( = 0.045 in.). The well-defined Gossard, Seide and
Roberts tests [19.37] served as a benchmark for verification of theory in the years that
followed.
Another series of exploratory, more design-oriented tests on thermal buckling of plates
was carried out at NACA Langley in the mid-fifties [19.124]. Kotanchik, Johnson and Ross
subjected four multi-web box beams (see Figure 19.54), two with ribs and two without, to
heating at a rate of 24-90 Btu/ft2 sec ( ~:n0-1010 kW /m 2 ). Heat was applied to one cover
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Thermal Buckling 1599

300

\ w1 ~ 0 ---------····
200 ----------------------- _.\_:_ ----------------.

100
=f CALCULATED
0 EXPERIMENTAL

.I .2 .3
(b)

Figure 19.53 NACA Langley 1952 thermal buckling experiments (from [19.37]): (a) test setup with
location of dial gages and thermocouples, (b) comparison of calculated and experimental
deflections at plate center

1,2 "
:
j___

I
<t- 3' v- Thermocouple

r
Rivet

r~8.
j1
f---~t--"

2i-.j_i ' 0:11t;5 r---~---


L211
r-:~J'
2" 16
5Ll
I
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,

Figure 19.54 NACA Langley 1955 rapid radiant-heating tests of multiweb beams-the box beams
tested, those with ribs, showing the location of the sensors (from [19.124])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1600 Thermal Buckling and Creep Buckling

only of each beam, and no mechanical loading was applied during the heating cycle. Heating
was by radiation from an array of graphite heaters, described in detail in Appendix A of
[19.124]. Each beam was instrumented with 15 thermocouples and 10-14 strain gages, 4 of
which were cantilever-type gages for buckling detection. The latter consisted of aluminum
alloy cantilever beams, attached to the unheated side of the beam, and steel pins with conical
ends, one end of which rested against the inner surface of the heated cover while the other
end rested against the tip of the cantilever through small holes in the unheated cover. Dis-
placements of the heated cover out of its original plane produced deflections of the cantilevers
and thus changes in the strain at their roots, which indicated buckling.
The typical sequence of events during the heating cycle was as follows. Upon exposure
to heat, the cover temperature increased rapidly, causing it to expand. The longitudinal ex-
pansion of the cover, restrained by the cooler webs, induced increasing compressive stresses
in the cover that caused buckling. During cooling, after relatively long times, the webs
developed buckles near the hot cover.
A typical temperature distribution, that in beam 1 with ribs, at a heating rate of 90 Btu/
fe sec ( ~ 1020 kW /m2 ), is shown in Figure 19.55. In (a) the distribution is shown after 2.3
seconds, when the cover buckled. As the temperature increased further, the buckles grew till
the maximum cover temperature was reached after 4 seconds. During cooling, after about
60 seconds, the web buckled, and the temperature distribution is depicted in Figure 19.55(b).
Upon cooling to room temperature, deep permanent buckles remained in the heated cover
and in the webs.

:::L-!'~"-
200 :
1 (o) Cover buckled, 2 3 sec
2
0 --~~--~--~~
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

1,000

800
."-
600
"'
U\

"' 400
:;"'
0 200
t
Q.
Maximum cover temperature, 4 sec.
E
~ 0

800

600

400

200
(c) Web buckled.

~----~----L--~4--~5---6~~
0
Distance from center of web, in.

Figure 19.55 NACA Langley 1955 rapid radiant-heating tests of multiweb beams-temperature dis-
tribution in beam I (from [19.124]): (a) after 2.3 seconds, when the cover buckled, (b)
after 4 seconds when the maximum cover temperature was reached, (c) during cooling
after about 1 minute, when the webs buckled
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Thermal Buckling 1601

A further design-oriented experimental investigation on thermal buckling of plates was


carried out at NACA Langley by Zender and Pride a couple of years later [19.119]. They
tested stainless steel square tubes, 32 in. ( ~813 mm) long and with sides of (bl t) = 40-80,
heated along two opposite walls and simultaneously subjected to axial compression. With
the aid of thermocouples, LVDT's and a pantograph (which multiplied the amplitude of the
buckles by a factor of 11), they determined the change in the onset of permanent buckling
with temperature. During heating, two opposite walls were exposed to radiation from two
banks of quartz lamps, while the other two were shielded from it by lengthwise aluminum
plates, projecting diagonally outward from the comers.
The purpose of the tests was to verify a simple approximate method, developed for pre-
diction of the onset of permanent buckles in plates. The results indeed validated the method
of prediction and indicated that the compressive load that a plate can support at the onset
of permanent buckling is substantially reduced with increasing temperature difference be-
tween plate and adjoining members.
These exploratory experiments provided designers with a qualitative and order-of-
magnitude assessment of the phenomena.
Thermal buckling and postbuckling of rectangular and circular plates was analyzed in the
early sixties for various temperature distributions and boundary conditions (see for example
[19.40], [19.43], [19.122] and [19.133]-[19.135]), but except for the two above-mentioned
NACA experiments and the 1960 Cranfield College of Aeronautics study [19.135], no ex-
perimental verifications were reported till the early nineties.
For shells, there were very few early thermal buckling experiments ([19.127], [19.130],
[19.36cb and perhaps one or two other rare tests), as mentioned in the previous subsection.
Of these, the 1962 NASA Langley tests by Anderson and Card [19.130] were probably the
most important. The tests included 13 19-in. ( ~483-mm) diameter stainless steel, ring-
stiffened cylindrical shells with (R/ t) ~ 300 and two ring spacings, (LI R) ~ 0.5 and 1.0.
The wall material of the shells was extra-hard type 301 stainless steel, chosen so that the
variation of E with temperature would be small and the expected stress levels would be well
within the elastic range of the material, in order to eliminate significant dependence of the
results on changes in material properties. The Z-section rings were spun from type 304
stainless steel and the cylinders were fabricated by spot-welding.
The cylinders were tested in the small NASA Langley bending test rig for stiffened cy-
lindrical shells (see Figure 9.75a of Chapter 9). A detailed view of the heater installation for
one of the cylinders in this bending test rig is shown in Figure 19.56.

One end of the cylinder is supported by a heavy backstop; a pure bending load is applied at the
other end by a weight cage acting through a loading frame with pin-connected linkages. Rollers
between the frame and support allowed lateral and longitudinal movement which minimized any
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

extraneous loads. The tip fixture was counterbalanced to eliminate any shear force on the cylinder.
The cylinders were heated rapidly by a 25-inch-long quartz lamp heater. For most of the tests
the heater extended over approximately one-third of the circumference and was symmetrically
located about the bottom extreme compression fiber of the cylinders, as shown in [Figure 19.56].
In a few tests the heater extended around the entire cylinder in order to produce uniform heating
around the circumference.
The procedure for each rapid heating test was to apply a bending moment less than the room-
temperature bending strength and then heat the cylinder at a rate of approximately 20oF/sec
(~ 11 oc I sec) until buckling occurred. During each test temperatures at several ring and skin lo-
cations were recorded, and in some tests strains were measured with resistance-type strain gages.
The vertical deflection of the tip fixture was continuously recorded in order to identify the instant
of buckling. By varying the applied load in different tests, a buckling interaction curve for load-
induced stress and temperature was obtained.
Conventional room-temperature wire or foil gages were successfully used in regions where tem-
perature did not exceed 175oF ( ~sooc). The gage response due to temperature alone was not large
compared with the overall strains being measured; therefore, the strains determined from these
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1602 Thermal Buckling and Creep Buckling

Figure 19.56 NACA Langley 1962 combined bending and radiant heating tests on stainless steel cy-
lindrical shells- details of heater installation in the Langley bending test rig (from
[ 19.130])

gages by using an appropliale lempcralurc· rcsponsc calibration curve were believed to be reason-
ably accurate.
An attempt was also made to measure strains in the heated po11ions of the cylinder with
high-temperature strain gages, but the data obtai ned were inconsistent.
Three room temperalllre tests preceded the heating tests in order to assess the experimental
scaller and detennine the knock-down factor from classical buck ling stress prediction fo r the
type of specimen and test sclllp used .
The experi mental results for the two unifotm ly heated cyl inders, the eight nonuniformly
heated o nes and the three shells tested at room temperature are ploned in F igure 19.57. "For
each specimen the maximum load-induced compressive stress is plotted as a function of the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

maximum cylinder temperature at time of buckling. The maximum cylinder temperature was
taken as the temperature at the bo ttom or extreme compressive fiber of the cylinder and
midway between rings in the bay nearest to the middle of the cyli nder. This temperature
was the actual maximum o r nearly the maximum for all cylinders tested."
The figure also depicts predictions based on room temperature tests of alum inum cy lin-
drical shells, carried o ut at NACA Langley some years before [9.247[, as well as NACA
Langley theoret ical calculations of the effect of rapid unifonn heating [ 19.1 28 ]. It may be
reiterated that if a cylinder were heated slowly and uniformly around the circumference so
as to produce no thermal st ress. the cylinder buckling stress wou ld be reduced in proportio n
to the reduction in E. The amount of this reduction is indicated by the short-dash curves in
Figure 19.57.
For non-uniform heating. the load-induced stress at buckling is further reduced, as evi-
denced by the square test points in the figure. For this condition axial thennal stresses are
present, which might be expected to add directly to the load-induced stress. so that buckling
would occur when the total compressive stress (load-induced stress plus axial thermal stress)
is equal to the buckling stress fo r uniform heating, as calculated from [ 19. 128].
Anderson and Card f 19.130] also developed a method for calculating thermal stresses in
cy lindrical shells subjected to a general temperature distribution. With this method they
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Thermal Buckling 1603

50

40

30
a,'
ksl
20 Test

0
10
c Nonuniform heatIng

200 400 600 800 1,000

(a) L/R • 1/2.

50

40 ""
~
~
----- ~-----
30

------
----_,_ -.........____, --
a,'
ks I
20
o·~-
----
----.._

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
I0

200 400 600 800 1poo

(b) L/R " 1.

Figure 19.57 NACA Langley 1962 combined bending and radiant heating test on stainless steel cylin-
drical shells~test results (from [19.130]): (a) close ring-spacing (LIR) = 0.5, (b) wider
ring-spacing (LI R) = 1.0

obtained reasonable agreement between their predicted stresses with the experimental results
(the correlation between the square test points and the long-short-dash curves in Figure
19.57).
Though it was an early thermal buckling test program, it was carefully executed, and since
many of the relevant difficulties and problems that troubled later experiments were already
dealt with and discussed in [19.130], it justifies meticulous study today.

19.3.3 Thermal Buckling Tests on Plates


A good example of recent thermal buckling tests on plates is those carried out by Thornton,
Coyle and McLeod at the University of Virginia in the early nineties [19.123]. The experi-
ment was essentially a modern version of the classic 1952 Gossard, Seide and Roberts NACA
Langley thermal buckling tests [19.37], discussed in the beginning of Subsection 19.3.2. The
major emphasis in the University of Virginia investigation was, however, placed on deter-
mination of the transient inelastic response of the rectangular plate, which necessitated the
application of significant levels of heating.
These were now obtained with a concentrated high-intensity infrared heat lamp, a tungsten
filament quartz bulb with an elliptical reflector that focused the applied radiant heat flux
along a narrow strip (see Figure 19.58), which replaced the electric heater of the 1952 tests.
In order to ensure a well-defined thermal environment, lamp characterization tests, using heat
flux gages, preceded the buckling tests. The calibration apparatus consisted of an incident
heated surface on an 8 X 4 X 0.5 in. ( ~203 X 102 X 13 mm) aluminum bar, which was
moved under the IR lamp, using an x-y cross-table. The bar was actively cooled by chilled
water, and the incident heat flux was measured with foil-type heat flux gages attached to the
heated surface between the two coolant tubes.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1604 Thermal Buckling and Creep Buckling

Figure 19.58 University of Virginia thermal buckling experimem on a Hastelloy-X four-point-supported


plate, heated along the cemerline by a quartz lamp-schematic of test fixture ( from
[ 19. 123])

In the plate buckli ng tests the edges of the Hastelloy-X (a nickel-based alloy commercially
available as sheet) plate parallel to the heater were maintained at a constalll temperature by
chilled water flowing through plastic tubes attached to the edges. The tubes were slotted.
fitted around the edges of the plate and sealed with a silicone-based adhesive so that the
coolant directly contacled the plate. The plastic tubes had negl igible bending stiffness and
therefore did not significantly restrict the deformation of the plate. To simulate an insulated
boundary, ceramic fi ber blanket insulation was used to cover the entire panel ( I in. insulation
o n the upper surface and 2 in. on the lower o ne), except for the narrow heated str ip. The
coolant tubes were encased in the insulating blanket. The insulation was light and flexible
enough to eliminate any effect on the plate deformations. T he heated strip was painted with
a high-emissivity paint, thereby increasing the heat absorbed.
The plate was supported at o nly four poi nts, by rods that contacted both the upper and
lower surfaces (see Figure 19.58), to provide well-defined stntctural boundary conditions and
minimize heat losses. At three of the points the rods contacted the plate with small spherical
tips to reduce contact area and thus minimize heat transfer from the plate. The spheti cal tips
also provided as little bending restrai nt as possible and allowed the plate to slide by the
supports when there was different '' thermal growth" between the plate and the support
fixture. At the fourth point the tips of the rod were conical and fitted into a sl ight indentation
in the plate, thus preventing in-plane mo tion and providing a reference point.
The interactions between the thennal and structural boundary conditions were kept at a
minimu m in this test setup. The thermal boundary conditions were appl ied without intro-
duci ng additional unwanted sttuctural stiffness, while stmctural restraints were designed to
allow in-plane thermal expansion. The con flicts between thermal and structural boundary
conditions. outlined by Blosser [19.1031 and discussed below, were therefo re minimized,
which contributed to a well-defi ned experiment whose resu lts provided useful data for com-
parison with analysis.
The plate was instrumented with 29 thermocouples and 15 LVDT's to measure temperature
and transverse displacements, respectively. The locati ons of these sensors is shown in F igure
19.59. To document the temperature gradients through the plate thickness two pairs of ther-
mocouples were located back to back o n the top and bottom surfaces of the plate.
In a typical test the chill water system was operated initially for 30-60 minutes to bring
the test plate to a uniform temperature of about 58-65•F (- 14-JS•q. T he coolant flow rate
was sufficient to maintain the edges of the plate at the initial temperature during the entire
heating test.
Five tests were catTied out. Two were elastic, with maximum temperatures of 250 and
375°F (- 120 and 190•c), respectively, while in the last three tests, with maximum temper-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Thermal Buckling 1605

ature of 500, 700 and lOOOF ( ~260, 370 and 540°C), respectively, increasing levels of
permanent deformations were induced. Center temperatures plotted versus displacement for
a typical elastic test (test 2), with a maximum temperature of 37SOF ( ~ 190°C), are shown in
Figure 19.60. For the lower temperatures one sees only a little out-of-plane displacement in
the figure, while above 200°F ( ~93°C) a sharp increase in deflection can be observed, in-
dicating the onset of thermal buckling.
Before discussing other recent thermal buckling experiments on plates and fiat panels, it
may be expedient to reconsider more generally the influence of boundary conditions in
thermal buckling. In thermal-structural testing the importance of the boundary conditions
exceeds even that emphasized in Chapter 11 and other chapters for the usual buckling ex-
periments, since here thermal boundary conditions are superimposed on the structural ones.
Blosser [ 19.1 03] reviewed the formulation and requirements for thermal and structural
boundary conditions and identified the possible conflicts between them. He pointed out some
of the problematic characteristics and interactions of experimental thermal and structural
boundary conditions. For example, severe thermal stresses are often the result of transient
temperature differences between parts of the structure, whose temperature rises at different
rates; or differences in the coefficient of thermal expansion of specimens and fixtures result
in significant boundary constraints, even with a uniform temperature rise.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
Blosser also reiterated two approaches to providing boundary conditions for thermal-
structural experiments, used also in buckling tests of nonheated structures, as pointed out,
for instance, in Chapter 9 for shells.
One approach is to use a specimen limited to the size of interest and to provide experimental
boundary conditions that match, as closely as possible, those of the analysis. The second approach
is to build a specimen larger than the region of interest so that there is transition structure sur-

- - 1-· ·--'--!-
0 0 0 0
a T25
T24
T27 T26 T23
oT22

oT21
oT20
OT19

0 0 0
TtS
T12 ~-T17
-o--o-o- -x
T16 T15 T14 T13 T6 T7 T11 T10 T9 TB
o TS
OT4
0 T3

0 T2

oT1
(a)

t =--=;---t-
0 L1J o L14 0 L15

I L12
o

0 0 0
I
o--o-o--o X
L5 L6 L7 L8 L9 L10 L11

o L4

o L1 Ol.J
0 "·
(b) -t---= --

Figure 19.59 University of Virginia thermal buckling experiment on a four-point-supported rectangular


plate, of dimensions 15 x 10 x 0.125 in. ( ~ 381 x 254 x 3.2 mm), heated by central
quartz lamp (from [19.123]): (a) locations of thermocouples, (b) locations of LVDT's
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1606 Thermal Buckling and Cref~P Buckling

400

0 0 0 o ooo
350

300
Temperature,
Of 250

150

100

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
50 ~---L--~.----L---~---L--~----L---i-

0.000 0.0:20 0.040 0.060 0.080


Deflection, in.

Figure 19.60 University of Virginia thermal buckling experiment on a four-point-supported rectangular


plate-plot of center temperature versus displacement for a typical elastic test (copyright
1994, from [19.123]. Reproduced by permission of Taylor & Francis, Inc., http://
www.routledge-ny.com)

rounding the region of interest. The intent of the larger test specimen is to have any undesirable
effects of imperfect boundary conditions to occur within the transition structure, so that the region
of interest will have the desired boundary conditions. The first approach uses a smaller, less ex-
pensive specimen and test apparatus, however, the second approach may often provide a more
accurate representation of the desired boundary conditions.

He pointed out that one often aims at limiting the heat loss at the boundaries to avoid
unwanted temperature gradients. Light, fi,~xible fibrous insulation can be used to cover un-
heated surfaces on hot structures and thus reduce the heat loss from natural convection or
radiation. Where the specimen contacts the boundary fixtures, these may be made of low-
conductivity material, or a guard heater may be employed to heat the fixture to the same
temperature as the specimen.
Finally, Blosser summarized seven general types of conflict:
1. Thermal expansion of the fixture may not match that of the specimen.
2. Thermal and structural fixtures need to occupy the same physical location or have access
to the same location on the specimen.
3. Thermal fixtures may reinforce or stiffen a specimen.
4. Heat transfer paths through the fixtures may affect the temperature distribution in the
specimen.
5. Acceptable temperature limits for fixtures or surrounding test equipment may be ex-
ceeded because of thermal loading.
6. Deformation of fixtures and/ or specimen may affect loading.
7. Changes in material properties with temperature may affect interaction of specimen and
test apparatus.
The task of the designer of test setups for thermal bucking experiments is to eliminate or
alleviate these conflicts, often a very difficult mission.
Other typical examples of recent thermal buckling experiments on plates and fiat panels
are the tests on titanium single-face hat-stiffened panels carried out by Teare and Fields at
the NASA Dryden Flight Research Facility [19.125]; the tests on titanium matrix composite
hat-stiffened panels performed by Richards and Thompson at NASA Dryden [19.96] and
those carried out by them on a titanium honeycomb sandwich panel [19.137]; or the studies
on estimation of the stiffness of test fixtures carried out by Polesky and Shideler at the
NASA Langley Research Center [19.126].
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Thermal Buckling 1607

The boundary conditions in most of these experiments received special attention along the
lines outlined by Blosser [ 19 .I 03] and discussed above. For instance, the test fixtures for the
titanium honeycomb sandwich panel investigated by Richards and Thompson [19.137] were
the subject of a significant development process. The sandwich panel, which was to serve
as a wing panel for a Mach 5 aircraft, was designed to withstand thermal stresses due to a
thermal gradient through the honeycomb panel thickness, with an outer skin temperature of
900°F ( ~480°C). In the aircraft wing the 23-in. ( ~585-mm) square panel would be sur-
rounded by other, similar panels. The surrounding panels would impose a symmetry bound-
ary condition along all four edges, which would imply an insulated thermal boundary and a
clamped structural boundary. In the experiment, however, the panel was allowed to expand
in-plane and the edges of the panel were not forced to remain straight. Thermal stresses
developed when the panel attempted to bend upward because of the temperature gradient
through the thickness but was restrained from rotation at the edges by the test fixture. The
preferred approach mentioned in [19.137] would have been to test the panel in the center of
a multipanel array (transition structure approach). Because of budget constraints, however,
only two panels were tested individually.
The text fixture for each panel was designed to apply uniform temperatures across the
upper surface, to prevent all four panel edges from rotating and allow the panel to expand
in-plane. The upper surface of the panel was heated by an array of 12 in. ( ~ 305-mm) long
quartz lamps, located 6 in. ( ~ 152 mm) from the surface.
As pointed out by Blosser, the most challenging portion of this test was designing a fixture
to restrain the edge rotation, since providing a perfectly clamped-edge boundary condition
is usually extremely difficult even for a room temperature test. The fixture shown in Figure
19.6la was used in the initial phase of the experiments.

- Back A('Jgle
"'la~ ____!l~(_lm ~

I beam If
Load
y_!_~-

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

\. C beam
stiffener
(a) (b)

Figure 19.61 NASA Dryden thermal stress and buckling tests on titanium honeycomb sandwich panel
(from [19.103] and [19.126]): (a) initial test fixture (with side heaters) moment restraint,
(b) final configuration of test fixture, without side heaters but with heavy stiffening 1-
beams, (c) top view of final test f]xture, showing the 52 moment restraint mechanisms
(13 on each side) around the perimeter of the panel
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1608 Thermal Buckling and Cre1~p Buckling

The angle beam, back plate, and C-beam stiffeners were intended to provide stiffness to react the
thermal moment. Bolts were used to transfer the load from the panel to the stiffened fixture, thereby
limiting the contact area and thus the heat loss into the fixture. There were 13 pairs of bolts along
each panel edge. Each pair of bolts was offset, as shown, to provide a restraint to edge rotation.
The spherical end of each bolt was seated in a spherical depression in a small load pad. The load
pad, in turn, pressed on a bearing plate which reinforced the panel. A high-temperature lubricant
was used between the load pad and the bearing plate to allow free in-plane expansion. Because
the fixture overlapped the edges of the test panel, the edges of the panel were shaded from the
applied radiant heat. These cooler edges could cause additional compressive stresses which reach
a maximum value in the center of the panel. Side heaters were installed in the fixture, as shown
in [Figure 19.6la], to counteract the effects of shading. Because the side heaters provided heat to
both the supper and lower surfaces, they did not maintain the temperature gradient through the
thickness of the panel and, thus, did not completely counteract the effects of shading. The unheated
surface was not insulated because the heat loss helped to develop the desired through-the-thickness
temperature gradient.

Preliminary test results showed large ditTerences between experiment and predictions.
[However] the major discrepancy between the anticipated thermal stress levels and those obtained
experimentally was the stiffness of the edge restraint. The edge fixture was much more flexible
than desired and, consequently, the thermal stresses were much lower than predicted.
After considerable evaluation of the experimental setup, the decision was made to simulate the
design thermal stresses, rather than attempting to match the design environment of the panel. The
edge fixture was modified, as shown in [Figure 19.6lb]. The side heater was removed for two
reasons: to allow the edges to be cooler so that the compressive stresses in the center of the panel
would be higher, and to allow the upper and lower angle beams to be more directly connected,
which greatly stiffened the fixture. In addition, large I beams were used to further stiffen the fixture.
The additional stiffening improved the edge fixity from 35 to 75 percent of a perfectly fixed
restraint. ... The edges of the panel remained nearly at room temperature, primarily due to the
shading by the test fixture. However, these modifications did not produce the desired thermal stress
levels. To reach the desired levels of thermal stress, the bolts in the mechanical restraints were
used to induce a mechanical preload and the temperature rise rate was increased from 3.5 to lSOF/
sec [ ~ 2 to soc; sec].

This fixture clearly demonstrates a conflict between the requirements of thermal and struc-
tural boundary conditions. The thermal requirement to minimize the contact area and the
need for a side heater inside the fixture softened the structural restraint. To stiffen the bound-
ary fixture, the required side heaters were eliminated and mechanical preload was induced
to arrive at a correct simulation of the thermal stress. This represents an example of the type
of compromise the test engineer may have to make.
In the NASA Dryden buckling tests on a titanium hat-stiffened panel [19.125], the panel
was mechanically compressed between fiat and parallel load platens in a hydraulic testing
machine. Since a large contact area was required for the transfer of the mechanical load, the
load platen had a relatively large mass and represented a significant heat sink at the ends of
the panel. To eliminate this heat sink and reduce the resulting heat loss, an independently
controlled heated test platen was introduced between the panel and the load platen. To protect
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

the test machine from overheating and minimize thermal warping of the load platens, the
heated test platen was separated from the load platen by a I /4-in. thick insulation and an
actively cooled platen (see Figure 19.62) . Additional edge cooling, with the aid of coolant
passages embedded in the edge frames of the panel, was employed in some of the tests to
induce thermal stresses in the panel.
The panel and the test platens were enclosed in an IR quartz lamp oven (as shown in
Figure 19.62). The lamps in the oven were located horizontally on either side of the test
panel and were grouped into eight equally sized control zones. The panel was tested under
eight different combinations of thermal and mechanical loads, consisting of three distinct
thermal test conditions and two mechanical ones. The three thermal conditions were: (l)
heating both sides of the panel to a uniform temperature, (2) heating of the panel on the
skin side only to provide a through-the-thickness thermal gradient, and (3) cooling the edge
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Thermal Buckling 1609

Test Panel

Quartz
Lamp CNen

Heated Load
Platen

1/4 ~ Insulation
Act1vely Cooled _
Load Platen Figure 19.62 NASA Dryden thermal buckling tests on a titanium
hat-stiffened panel-schematic diagram of test setup
(from [19.125])

frames were cooled along the free edges while heating the panel on both faces. The two
mechanical load conditions were: (l) axial loads (parallel to the hat stiffeners), (2) cross-
corrugation loads (perpendicular to the hat stiffeners).
The panels were extensively instrumented, and several measurement techniques were em-
ployed, such as the moire grid shadow method, elevated temperature strain measurement
methods and a single-gage force stiffness method to assess the buckling load. The moire
grid shadow method (mentioned in Chapters 8, 9 and 14) is an optical displacement mea-
surement method that relies on the superposition of a grid pattern and its shadow on a test
article, with the resulting pattern indicating the out-of-plane deflections. The moire grid
method could, however, be used here only for room temperature tests since the oven inter-
fered with the optical lines of sight. The elevated temperature strain instrumentation and
measurement techniques employed at NASA Dryden were discussed in detail in Subsection
19.2.3 (see for example Table 19.1) and need not be recapitulated here.
The single-gage force stiffness method, on the other hand, is a useful extension of the
force stiffness (FIS) method (discussed in Subsections 15.3.2 and 15.3.3 of Chapter 15)
motivated by the special limitations of thermal buckling tests, or by limitations imposed by
the fabrication process of the specimens. The single-gage force stiffness method then replaces
the usual dual-gage method, which is based on two back-to-back strain gages that measure
bending strains.
Here the single-gage method was necessary because instrumenting hat-stiffened panels
with back-to-back strain gages would have to be performed prior to the fabrication of the
panel, and the gages could not survive the fabrication process. In the single-strain-gage
method [ 19. 96] and [ 19 .125], a single gage determines the local bending strain. The method
divides the strain gage output into two parts: a linear and nonlinear response.
Figure [19.63a] shows the typical response of a strain gage to load under uniaxial compression in
a buckling situation. The gage responds linearly up to a load where bending is introduced, after
that point it responds nonlinearly. The single-strain-gage method uses the linear response of the
gage to determine the bending strain during the nonlinear response portion. By fitting a straight
line through the linear portion, an extrapolation beyond the bending introduction point can be
made. The assumption is that if the gage would continue to respond linearly with load, the output
of the gage would follow the dashed line. The extrapolation beyond the bending introduction point
enables the bending strain to be computed as the difference between the measured strain and the
strain from the linear extrapolation. The bending strain can then be used in the F IS plot.
Figure [19.63b] is a typical FIS plot in which FID is plotted as a function ofF, where F
represents the compressive load and D the bending strain found in Figure 19.63a. The curve moves
downward to the right as the critical buckling load is approached. By selecting a linear-fitted range
near the bottom portion of the F IS curve, a line can be extrapolated down to the load axis to
predict the buckling failure load. The fitted range is usually the lower portion of the F IS curve
and is based upon judgment, experience, and upon the assumption that no load path or mode
changes will occur before the intersection.

It may be noted that the single-strain-gage F IS method, outlined above, can also predict
local buckling. It therefore complements the application of the F IS plot to complex buckling
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

behavior, mentioned in Subsection 15.3.2 (see for example Figure 15.45 there).
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1610 Thermal Buckling and CreE!p Buckling

Measurwdll1111n
Linear extrapolation

Bending
\ lnlrodueHon
St111ln

Llnear111nc~ ',, Bending


''' ll1111n (D)

---,,!_

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
JComp,...lon lotldJ, (F)
(a)

-111<1 atr•ln
Linear ext111poletlon
Bending Introduction

f.
D
----1
" I
"'- !
r
Linear lmed 111nge

Pl'lldlcted
buckling

(b)
"'----,:1
!Compression 1011111, (F)

Figure 19.63 The single-gage force stiffnes~. method (from [19.96]) (a) typical strain output as a
function of load, (b) typical F IS plot

After completion of the thermal-mechanical test program on the monolithic titanium


(Ti5Al-4V) hat-stiffened panel, the test techniques developed were applied at NASA Dryden
to a similar nondestructive experimental program on a corresponding titanium matrix com-
posite (TMC) hat-stiffened panel [19.96]. Among the techniques employed were single-
strain-gage F/S buckling predictions at room temperature and at elevated temperature equi-
librium conditions, representing a practical example of the method discussed.
Figure [19.64a] is a typical plot of strain as a function of load for the case of axial loading at a
steady-state equilibrium condition of 500oF ( ~260°C). The gage used for the plot is located on the
skin-side near the center of the panel and between the legs of a hat stiffener. The gage response
is nearly linear up to a certain load, after which bending is introduced and the gage responds
nonlinearly. The bending strain is determined by the difference between the extrapolation of a
linear fit through the linear range and the measured strain beyond the bending introduction point.
The bending strain is then used in a single-strain-gage FIS plot shown in [Figure l9.64b]. This
plot shows F I D plotted as a function of F, where D is the calculated bending strain of [Figure
19.64a]. The lower limit of Fl D values is reached as a result of experience gained from previous
tests. The desire is to get the F I D value as low as possible without buckling the panel, to achieve
a good prediction of the critical buckling load. To obtain an accurate approximation of the (critical)
buckling load, a fitted range of the F IS curve needs to be selected. For this case a straight line
was drawn through the fitted range and the (critical) buckling load was predicted. Figure [19.64b]
does not provide the exact local buckling load but this local prediction does provide an estimate
of the general instability occurring in the panel. A good estimate of the overall panel buckling load
can be obtained by averaging local buckling load predictions over an area of the panel away from
the edges. The single-strain-gage method provides a good understanding of the local behavior of
the panel (load path and mode changes can be readily observed) and is used to make real-time
Copyright Wiley judgments of panel integrity [19.1381.
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Thermal Buckling 1611

_ _ _ ;7 Maximum
bending

- Measured slr81n

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
- - - Linear extrapolation

(a) !Compression loedj,lb

II -e- LlnNrty tmed nne


Measured str8ln
LlnNr ext,.polatlon

F
D.
111/IJ.ltVIn.

[ JCompresslon load!, lb
(b)

Figure 19.64 NASA Dryden elevated temperature tests on a titanium matrix composite (TMC) hat-
stiffened panel-application of the single-strain-gage force stiffness method, buckling
prediction at an elevated temperature of 500oF (- 260°C) equilibrium condition with the
panel in the axial mode (from [19.96]): (a) typical strain output as a function of load,
(b) typical F/S plot

Additional examples of thermal buckling experiments on flat panels are the tubular Rene
41 panels (which were essentially flat panels reinforced by tubular stiffeners; see Figure
15.49 and [15.66]), tested at NASA Dryden, and the Rene 41 (a nickel alloy) honeycomb
sandwich panels, also tested at that research facility (see [19.139]). The former represented
the NASA hypersonic wing test structure (HWTS) of the eighties and were subjected to
combinations of different temperatures and mechanical loads. The latter simulated the ther-
mal-structural loads on an integral tank-and-fuselage structure for an earth-to-orbit vehicle.
The thermal loads were obtained by immersing one face sheet of the panel in LN 2 to achieve
a cryogenic temperature representative of that associated with an LN 2 propellant while ra-
diantly heating the other face sheet to the temperatures produced by aerothermal heating
during ascent or reentry of the vehicle. The first of these two panels failed by local core
crushing, whereas the second sustained the imposed load cycles, though with large plastic
stains. Except for core buckling, buckling was actually not a significant failure mode in these
tests, but the test setup and the behavior under combined thermal-structural loading are
relevant also to proper buckling experiments.

19.3.4 Thermal Buckling Tests on Shells


The few early thermal buckling tests on shells (such as [19.127], [19.130], [19.136] and
[19.140]), briefly discussed in Subsections 19.3.1 and 19.3.2, were followed by a second
small group of experiments in the mid-sixties (see for example [19.141]-[19.145]) and by a
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1612 Thermal Buckling and CreE'P Buckling

third in the seventies (see for example [19.146]-[19.149]). But the harvest of experimental
investigations on thermal buckling of shellls has still been rather meager and has provided
insufficient data, pointed out by reviews of the field (see for example [4.48] and [19.43]).
The second group of experiments was carried out at Stanford University and at the nearby
Lockheed Missiles and Space Company Palo Alto Laboratories. The Stanford investigations
actually began in the late fifties with Hill's tests on aluminum and steel circular cylindrical
shells heated along very narrow axial strips [ 19 .140]. The boundary conditions in these tests
were intended to simulate simple supports and clamping. The clamping support rig was,
however, far too flexible to adequately simulate clamping, and hence only the shells with
simple supports can be considered valid tests.
Ross, Mayers and Jaworski continued and extended the experimental work of Hill some
years later (see [19.141]). They tested 16 type 304 stainless steel cylindrical shells (with
Rlt = 334) and 8 type 1006 cold-rolled steel ones (with Rlt = 291), subjected to IR lamp
heating along and axial strip. All shells were 10.4 in. ( ~ 264 mm) in diameter and 48 in.
( ~ 1220 mm) long, and their end supports were designed to simulate fully clamped condi-
tions. This was achieved by supporting the test specimens "on cylindrical rims which pro-
vided an inner bearing surface of 1.5 in. ( ~ 38 mm) at either extremity of the shell. Each
rim was formed as an integral part of a h<:avy head piece which could, in tum, be bolted to
the test frame to provide rigid and immovable end support. By means of steel clamping
bands, the test cylinders were securely bound at each end to the supporting rim such that
boundary displacements and rotations relative to the axial, tangential and radial directions
were eliminated."
The strips of the specimens to be heated were painted with a black stovepipe enamel prior
to the test, on both the inside and outside . The thermal load was applied simultaneously by
IR lamps and reflectors to both inner- and outer-painted axial strips. The heated axial strips
were located diametrically opposite to the longitudinal splice of the test specimens. The
width of the heated strips varied from 1.5 percent to 18 percent of the total shell circum-
ference.
Instrumentation included Chromel-Alumel thermocouples, whose output was autophoto-
graphically recorded during the test by means of a CRC oscillographic recorder, as well as
a sensitive microphone, which provided an audio signal at the instant of buckling.
In the tests the cylindrical shells were subjected instantaneously to the maximum output
energy of the lamp heaters, which continued to deliver an essentially constant heat flux until
buckling occurred. The axial and circumferential temperature distributions were then re-
corded until the shell buckled. The temperature profiles obtained at buckling were consistent
for both types of shells, approximating a constant axial distribution and a typical bell-shaped
circumferential distribution. For the stainless-steel shells, buckling temperatures were around
300°F ( ~ 140°C), whereas for the thicker cold-rolled steel shells they were around 500°F
( ~260°C). In both series there was significant scatter in the experimental buckling temper-
atures.
In cylindrical shells with narrow heated strips, buckling occurred as a sudden localized
snap-through, accompanied by a distinctly audible sound. For the wider heated strips, lo-
calized buckling patterns appeared simultaneously at several locations within the heated strip.
In the thicker cold-rolled steel cylindrical shells of the second test series, having wider heated
strips, a different failure mode occurred with a barrel-shaped deformation leading to yielding
at the clamped supports. Agreement with Donnell theory analysis, using measured circum-
ferential temperature distributions, was only fair but was considered a confirmation of pre-
vious conclusions by other investigators. that the buckling temperature corresponding to
uniform axial compression represents a lower limit to the buckling of cylindrical shells
subjected to highly non-uniform circumferential heating.
Ross, Mayers and Jaworski [19.141] were surprised to find that their experimental buckling
temperatures were consistently higher than the predicted ones, in contradiction to the well-
known trend for mechanical loading of imperfect shells, that experimental results are sig-
nificantly below the classical critical stresses. To clarify these paradoxical results, another
Thermal Buckling 1613

experimental investigation on uniformly heated clamped cylindrical shells was undertaken at


Stanford University by Ross, Hoff and Horton [19.142].
Six stainless steel cylinders with (Rit) = 344 and five cold-rolled steel cylinders with
(R It) = 291, similar to the specimens tested in the earlier study, were heated axisymmetri-
cally by an array of 12 IR heat lamps, which were aligned parallel to the inner surface of
the cylindrical shell along circumferentially equidistant longitudinal generators. The ends
of the cylinders were longitudinally restrained and clamped. The test procedure was essen-
tially the same as that used in the previous tests. Again a step input, with respect to time,
of maximum thermal energy was applied, but here the heating was introduced uniformly,
with the assistance of a concentric cylindrical reflector, to the entire inner surface of the
shell.
The test rig was similar to that of the earlier test program, with some modifications.
A continuous representation of the axial-compressive load resisted by the cylindrical shell during
heating was obtained by recording the output of a highly sensitive, semiconductor, temperature-

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
compensated, strain-gage bridge which was mounted midway between the end supports, on the
central shaft of the test-rig structure. With the load-measuring system, it was possible to resolve
and record the axial-compressive force resisted by the test cylinder to an accuracy of 0.25 percent.
... Twenty-three Chromel-Alumel thermocouples, which were spot welded to the test specimen
at strategic locations, comprised the necessary sensors for determining the temperature distributions.
The thermocouple stations on the shell outer surface were chosen to provide the necessary data for
constructing smooth temperature profiles in both the longitudinal and circumferential directions.
The thermocouple outputs were cyclically scanned (scan period, 4.20 sec), attenuated through a
low-level preamplifier and then autographically recorded on an individual channel of a Sanborn
recorder. In addition to the scanned information, a continuous recording of the temperature existing
at top center (x = 0, y = 0) of the cylindrical shell was obtained during the entire test. Another
thermocouple was attached to the central shaft, adjacent to the strain gage bridge. Thus, a temper-
ature history of the central shaft, during any heating sequence, was also acquired. The temperature
of the two end-support clamping rings was monitored during particular tests. In this manner, data
pertaining to the thermal expansion of the clamping rings were obtained .
. . . A contact microphone, firmly attached to the test specimen, provided an audio-output signal
which was subsequently amplified and then simultaneously recorded with the load and temperature
histories.
The experimental data were compiled and graphically presented in the form of axial-
compressive load versus equivalent end-shortening curves for each cylinder tested. Typical
plots are shown in Figures 19.65a and b. The calculated values of equivalent end shortening,

I I

I
I

I
~
(,
;,
(,
It
h
/;
!J

0
o 10 20 30 40 so Go 10 eo 90 100 uo 120 130
5
End Sh01'feni119 o TL 1t 16 1n
(b)

Figure 19.65 Stanford University 1966 thermal buckling experiments on uniformly heated steel cylin-
drical shells-plots of axial-compressive load versus equivalent end shortening (from
Copyright Wiley [19.142]): (a) stainless steel-Specimen no. 6, (b) cold-rolled steel-Specimen no. 7
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1614 Thermal Buckling and CreE!p Buckling

based on the idealization of rigid end support, were individually corrected by an amount
equal to the sum of the elastic and thermal expansions experienced by the central support
shaft. The load versus end-shortening curves for which this correction was significant are
shown by a dashed line in the figures.
The buckling process was as follows:
Upon initiation of heating, a test cylinder experienced outward thermal expansion in the radial
direction. As a result, the thin circular cylindrical shell became barrel-shaped. The ensuing cur-
vature of the initially straight generators relieved the compressive thermal stress so that, ultimately,
the shell buckled near the clamping bands as a consequence of the severe restriction of both lateral
expansion and end rotation.
In all the tests, it was observed that buckling occurred as a localized "snap-through" accom-
panied by a distinctly audible sound. The test program indicated that "snap-through" type insta-
bility was more pronounced in the stainless-steel test specimens than in the cold-rolled specimens.
In addition, the stainless-steel cylinders buckled in some instances at locations not necessarily
adjacent to the clamped boundaries.
A "complete" buckle pattern (a deformat:~on pattern in which half-wave buckles encompassed
the entire circumference of the shell) in the circumferential direction consisted of between nine
and twelve half-wave buckles which were located within a single tier in the vicinity of one cross-
sectional plane. In cases where the deformation pattern was "incomplete", the buckles were gen-

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
erally arranged in a stepped-tier array with some buckles off the cross-sectional plane.
The buckles were initially oval shaped or elliptically defined in outline with a circumferential
half-wave length of approximately 1 ~h in. ( ~ 38 mm, i.e. the number of circumferential half-waves
n ~ 20). Continued heating after the "snap-through" process resulted in transition of the individual
buckles to the well-known postbuckled diamond shapes and an enlargement of dimensions so that
a "complete" deformation pattern comprised eleven or twelve buckles.
This, accompanied by simultaneous readjustment in axial load, was the usual secondary
buckling (discussed in Chapter 9; see for example Figures 9.3 or 9.32) .
. . . The effects of lateral outward thermal expansion on the load-supporting ability of the test
cylinders were particularly important. Because of the continually increasing deflection of the ini-
tially straight axial generators, the cylindrical shell refused to produce the anticipated axial-
compressive stress. This behavior was observed in the prebuckling range. Thus, even though the
response of certain cold-rolled steel cylinders prior to buckling was almost linear, the slopes of the
load vs. end-shortening curves realized from experiments were appreciably less than the slopes of
similar curves obtained for uniformly compressed, machine-tested cylindrical shells. Consequently,
the equivalent end-shortening associated with a given value of axial load in a thermal-buckling test
was greater than the analogous end-shortening value realized in a machine-buckling test.
... In comparison with machine testing, the thermally buckled test cylinders experienced average
critical temperatures approaching 59 percent (cold-rolled steel) and 65 percent (stainless steel) of
the classical value, whereas the critical loads evidenced in thermal testing were characteristically
low, i.e., 25 percent (cold-rolled steel) and 26 percent (stainless steel) of the classical critical value.
Ross, Hoff and Horton concluded that because the cylindrical shells barreled out during
heating, the axial compressive stress was somewhat relieved, allowing higher critical tem-
peratures than those predicted by linear buckling theory. Though this conclusion was not
accepted by some investigators (for example, Bushnell and Smith [19.146] indicated that for
cylinders of the geometry tested by Ross et al. the barreling effect was not significant and
suggested that the axial stress relief there might have been due to slippage at the boundaries),
the barreling effect and the resulting nonlinear relation between temperature rise and stress
are now generally accepted as an explanation of this thermal buckling paradox.
Since the investigators at Stanford University had encountered difficulties with their sup-
port systems of unknown stiffness, Smith at Lockheed Palo Alto [19.143] and [19.146] took
particular care to provide maximum rigidity to his test rig for thermal buckling of conical
shells. The support system of the test setup, shown in Figure 19.66 consisted of two thick
steel end plates connected by a thick central steel tube. "This assembly was made more

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Thermal Buckling 1615

Clamped

Figure 19.66 Lockheed Palo Alto 1964 thermal buckling tests on conical shells heated along axial
strips-schematic of conical shell in test rig, showing the rigid support system (from
[19.146])

rigid by the introduction of 8 half-inch thick steel webs welded at equal intervals to the
larger diameter plate and central tube. The shell was potted into grooves in the end fixtures
by the following procedure: the large diameter plate was heated to 600aF (~315°C) with an
acetylene torch. The conical shell was then placed in the fixture and allowed to come to
thermal equilibrium. Then 50-50 tin-lead solder heated to 600°F ( ~ 315aC) was poured into
the mounting groove." The solder solidified by natural cooling in 10 minutes and the plate
was fully cooled in 5 hours. The rig was then inverted and the potting was repeated for the
small end. "After solidification of the solder at the small end, the clamping ring indicated
in [Figure 19.66] was released, to allow the small end to slide on the central tube during
the subsequent 5-hour cooling process and thus relieve axial thermal axial stresses. After
cooling, the clamping ring was retightened to provide nominally clamped boundary condi-
tions."
The conical shells were heated along a center strip by an array of IR quartz tubular heaters.
The test procedure for the conical shells was similar to that employed at Stanford University
for cylindrical shells. The rigidity of the test setup, obtained by a combination of the heavy
4-in. (-102-mm) diameter, 1.25-in. (-32 mm) thick steel central tube, the two 2-in. (~50-
mm) thick steel end plates and the eight 0.5-in. thick connecting steel webs, is worth noting.
In passing, it should be pointed out that the paper of Bushnell and Smith [ 19 .146] probably
represents the first systematic comparison between extensive computations (with BOSOR 3)
of thermal buckling of shells and available experimental results.
In the mid-seventies two programs of buckling tests on cylindrical shells subjected to
combined heating and mechanical loads were carried out at the Technion in Haifa [9.178],
[19.147]-[19.149]). All the shells were heated along two opposite generators, but in the first
program, of 46 specimens, the mechanical preload was axial compression only, whereas in
the second one, of 35 specimens, the mechanical preloads were combinations of axial com-
pression and torque. The test specimens in both programs were all of the same geometry
and material (R = 191 mm, net length L = 670 mm and t = 0.635 mm, i.e. Rlt = 301 and
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1616 Thermal Buckling and Creep Buckling

LIR = 3.5). They were made of Alclad 2024-T3 sheet (with ± 10 percent thickness varia-
tions) rolled and riveted with a standard lap joint, which in the tests was located midway
between the heated generators.
The test setup shown in Figure 19.67 was employed for the first test program (with axial
preload only [19.148]). Essentially the same setup, but after it had been adapted for appli-
cation of torque (see Figure 19.69), was used for the second test program [9.178] and
[19.149]. In the basic configuration (Figure 19.67) the shell was mounted on the isolating
gasket rings (15), between the front (9) and rear (12) bases. The purpose of the rings (5),
made of pressed Durestos, was to provide heat insulation between the shell and the bases.
The shell was tightened around the edges by split clamps (16) while pressed against the
isolating gasket rings. The clamp details are shown in Figure 19.68. The clamps consisted
of four aluminum 2024 quadrants having small clearances between themselves and an in-
ternal diameter equal to the shell external diameter. The quadrants were fitted into V-slotted
split rings. The clamping effect was achieved by tightening the protruding bolt, (c) in figure
19.68. It was presumed that in this way the radial deflection w, the angle of rotation awl ax

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
and the circumferential displacement v were kept at zero throughout the heating. The clamps
also prevented a possible relative sliding between the compressed surfaces, which contributed
to maintain zero axial displacements. After clamping, the shell was compressed axially by
the hydraulic jack (2) to close any available clearances. Later the pressure was released,
permitting instrumentation calibration.
Because the heaters were installed above and below the shell, the heat output of the upper
heater had to be increased in order to obrain similar temperature distributions. A variable
autotransformer (a Variac) was therefore incorporated in the test setup (see [14] in Figure
19.69), to control the power supply to the lower heater. By experimental adjustment, the
temperature rise was kept similar in both top- and bottom-heated generators.
In the first program the temperature distribution was measured with thermocouples, the
displacements were measured with LVDT's and the onset of buckling was detected by two
microphones. The influence of the in-plane u-displacement boundary condition was also
studied by Frum and Baruch, who concluded that the u-displacement had a dominant effect
on the onset of buckling.
In the second test program [9.178], the cylindrical shell, (10) in Figure 19.69, was again
mounted between two bases, (9) and (12) . around which it was tightened by rings, similar
to Figure 19.68. The tight clamping prevented slipping between the shell and the bases when
a torque was applied, and as before restrained deflection and rotation of the shell at its

Figure 19.67 Technion experiments on cylindrical shells subjected to combined heating along two
generators and axial compression-test setup (from [19.148]). Legend: 1. frame, 2. hy-
draulic jack, 4. load cell, 9. front base, or shell clamping ring (movable edge), 10.
cylindrical shell, 11. upper generator line heater, 12. rear base, or shell clamping ring
(fixed edge), 15. isolating gasket ring, 16. clamp, 17. combination bearing, 18. central
shaft, 19. angle regulator
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Thermal Buckling 1617

Figure 19.68 Technion expeli ments on cylindrical shells subjected 10 combined heating along two
generators and axial compression- the split clamps (from [19. 148]): (a) steel ring, (b)
aluminum quadrant. (c) tightening boll

boundaries. The rear base ( I 2) was allached to the supp01ti ng frame, whi le the front base
(9) allowed axial displacement and rotation about the axis of the cylinder. Thus. axial com-
pression and torque could be applied through the front base, by a hydraulic jack (2) and a
torque application device (5)- ( 8).
Axial displacement was agai n measured here with an LVDT, but the normal deflections
of the heated generator were measured by means of a simple photographic technique de-
scribed below. When the axial thermal expansion had to be restrained during a test, the
required increasing axial force was supplied by the hydraulic jack, through the manual con-
trol of the Amsler universal testing machine, to which the jack was connected.
The circumferential rigid rotation of the front clamped edge of the cylinder about its axis
was measured by a thin cantilever, on which strain gages were bonded and calibrated to read
its end deflection. This measuring device was posi tioned so as to measure the tangential
displacement of the front base.
Temperatures were measured along the heated generators only. In most tests only two
thermocouples were used for measurement at the middle of the generators. ln a few tests
temperatures were also measured near the thermally isolated clamped edges.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 19.69 Technion experiments on buckling of cylindrical shells under combined axial preload,
heating along two generators and torque-lest setup (from [ 19.149)): Legend: I. frame,
2. hydraulic jack for axial compression. 3. amplifier and switch for heater, 4. load cell,
5. weight applying torque, 6. pulleys. 7. wheel, 8. torque arm, 9. shell clamping ling
(movable edge), 10. cylindrical shell. II. upper generator li ne heater, 12. shell clamping
ring (fixed edge), 13. restraint of fixed end rotation, 14. control variac for lower heater
power
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1618 Thermal Buckling and Creep Buckling

Since the measurement of normal defleciions at the heated generator was difficult because
of the high temperatures occurring there, a photographic technique was developed. The initial
gap between the IR heater and the shell was about 6 mm and reduced to much less during
a test. A camera was placed in front of the gap between the upper heater and the shell (see
Figure 19.70). A long millimetric grid was placed above the cylinder and photographed.
This picture was used as a reference scale, according to which the following pictures, which
were taken during the heating, could be calibrated. The second picture was taken immediately
at the beginning of the operation of the heater and used as a record of zero deflection. Then,
every few seconds as required, more pictures were taken, until buckling occurred. Any pho-
tograph taken during a heating test (see Figure 19.72) included four strips, which, counting
from the top, were:
1. A wide black strip-the blackened side-wall of the heater.
2. A thin bright line-the lower side of the heater, illuminated by the reflection from the
shell.
3. A black strip-the gap between the heater and the shell, whose width decreases as the
shell deflection increases.
4. A bright area, which is the shell, illuminated by the IR heater. Under this wide strip is

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
a black area, which is the continuation of the shell.
Typical test photographs for one shell are shown in Figure 19.73. The thermocouple was
also used as a reference point for measuring axial distance along the generator after the
reference grid was removed. It should be noted that every camera triggering shortened the
recorder input, where the temperatures were recorded versus time. Hence, every picture was
marked and correlated to the corresponding time and temperature in the test.
After each test the film was developed ;md the printing of all the pictures was performed
in the same conditions and on the same paper, to minimize printing distortions.
The data were read with a Nikon projector (model V-16) using a magnification ratio of
20:1. The reliability of the data was examined in a special test, by comparison with results
that were simultaneously read by five LVDT's. These LVDT's were positioned perpendicular
to the inside upper generator along the shell (see Figure 19.71, where the 250 X 350 mm
opening cut for accessibility is also shown) and measured its deflections. The comparison
of the LVDT measurements with those obtained from the photographic technique showed
differences of 3-7 percent for temperatures up to IOOC, when the LVTD's lost their linear
behavior.
The test procedure included mounting of the shell, brazing of thermocouples and tight
clamping. Then the desired torque and axial preload were applied and a reference grid
photograph was taken.

LINE HEATER

CAMERA
\
BLACKENED SIDE-........._~

~ CYLINDRICAL
SHELL

Figure 19.70 Technion experiments on buckling of cylindrical shells under combined axial preload,
heating along two generators and torque-photographing the gap o (from [19.149])

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Thennal Buckling 1619

Figure 19.71 Technion experi ments on buckling or cylindrical shells under combined axial preload.
heating along two generators and torque-special comparison test of LVDT measure-
mcms with photographic technique resuhs: positions of LVOTs (from [ 19.149]). Note
the 250 x 350 nm1 accessibility opening for installation or the five LVDTs

During heati ng tests, the upper and lower temperat ures were read and recorded on an XYY
recorder, where every picture of the upper generator was marked. W ith the Amsler testing
machi ne, axial compressio n was carefull y added manually, as required for prevemio n of
axial d isplacement. The compression and the ax ial displacement were both recorded on the
XYY recorder, and the rigid ro tation of the front edge of the cylinder was also recorded. The
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

tests were usually continued beyond the initiatio n of buckli ng.


Three test series were carried out in this program. In each series the same constant axial
preload was exerted. whereas different to rques were applied in o rder to develop temperature-
torque interaction curves. Each cylindrical shell was used for one test o nly.
The buckling modes included local buckling, appearing at a heated generator and mostly
near one of its edges, and torsional buckling, where the torsio nal waves were. as expected
more well defi ned the larger the torque and the smaller the axial preload. IL was found that
for small and medium torques, the initial buckling was local. In these cases the tests were
continued. For small torq ues additional local buckles appeared, but fo r medium torques, a
to rsio nal buckling mode appeared later, in the second, third o r fo urth buckli ng due to the
additional applied torq ue. For large torques. the first buckling was in a torsional mode and
the tests were no t continued further. Measurement of the u pper-generator deHection showed
a correlation between the expansion process and the buckl ing mode. Where only small or
medium torques were applied, the dcHcction grew grad ually from the clamped edges to the
center of the cylinder and the buckling was of the local mode near an edge of a heated
generator. In shells loaded by large torques a central indentation, which deepened gradually,
was regularly found and the buckl ing mode was then entirely torsional.
The radial pre buckling deHecti ons of the heated generator indicated a close and significant
relation between the deflection and the buckling behav ior. The de flected shape was inHuenced
by the load combination and affected the buckling modes. The application of torque caused
a dip at the midlength of the cylinder. On the other hand, heating caused expansion and
hence an outward deflection of the generato r. Fo r small torq ues the thermal expansion
straightened the generator and the torsional central dip disappeared. Because the clamped
edges restrained expansio n, the largest curvatures appeared there and caused local buckling.
For large torques the central dip became deeper relative to its surroundi ngs during heating.
Now the torque predominated and buckling was torsional. For medium torques the buckli ng
mode was local edge buckling when the buckl ing occurred after the central dip had di sap-
peared. but when the buckling occurred before that, the dip was in the center.
The condition of uni form heating along a line aimed at could no t be perfectly compl ied
with. T he initial heated line. due to its deflection and curvature, became a heated strip. whose
width and heat flux varied with the increasing temperature. However, as shown in [19. 147 ],
small variatio ns in the width of the strip d id not significantly affect the temperature distri-
buti ons.

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1620 Thermal Buckling and Creep Buckling

It was found that in general small torques did not significantly reduce the critical tem-
peratures. There appeared 10 be interaction effects due to the different nature of the simul-
taneous loadings and the thermal expansion that stiffened the shell for small torques. For
medium torq ues, a mode interaction (local edge buckling and torsional buckling) occurred,
which signi ficantly decreased the critical temperatures.
Practically no addi tional the1mal buckling experiments on aerospace shells were carried
out in the eighties and early nineties, though there was considerable theoretical and com-
putational activity (see for example [ 19.43) and [19. I 50]- ( 19. 152)). The scarcity of thermal
buckling tests on aerospace shells is exemplified by Thornton's 1996 review [ 19. 153], in
which the most recent shell thermal buckling tests mentioned date back to the early seventies.
Some thermal buckl ing tests on shells were, however. performed in other engineering fields
in the last decade, such as for nuclear reactor component design (see [ 19.88] and [ 19.106]).
The lull of a decade in aerospace thermal buckli ng experiments came to an end when the
preparations for the U.S. National Aerospace Plane (NASP) revived interest in such st udies.
However, actual experimental elevated temperatures investigations on shells and curved pan-
els conmtenced only in the mid-nineties, in connection with the U.S. High-Speed Research
Program (HSR, a supersonic transport class of vehicles) and the Reusable Launch Vehicle
Program (RLV, an earth-to-low earth orbit Shuttle-like class of vehicles), on which NASA
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

research activities have focused in the nineties.


T he research effort at the NASA Langley Research Center started by defining, designing
and developing special new test facilities for combined thermal and mechan ical loading of
shells and curved panels, with emphasis on composites (see for example [ 19.154]. This
Combined Loads Test System (COLTS), that has been developed at Langley, is a unique
stmctures research test facility that permits complex, combined loads testing of large aero-
space stn1ctures including thermal loading. Typical aerospace stmctures to be tested will
include panels from transport aircraft fuselages. full fuselage barrels. and panels from launch
vehicles. The COLTS consisL~ of two pressure box test machines (see Figure 19.77) and a
combined loads test machine (sec Figures 19.74 and 19.75). A small-scale combined loads
test machine has also been built at Langley for experiments on laboratory-scale curved panels
(see Figure 19.78).
The motivation for development of the NASA COLTS was presented in [ 19.154], [ 19.155]
and [ 19. 158]. Briefly, it was argued that for developing verified structures technologies that
could be employed in future airframe designs, a faci li ty was required in which different
configurations of full-scale and sub-scale panels, fuselage shells and wing boxes cou ld be
tested with combinations of mechanical, internal pressure and thermal loads. In other words,
a combined load test system with the flexibility to accommodate the whole spcctmm of these
structures and loadi ngs was needed. Titc COLTS system was designed, in cooperation with
U.S. industry, to answer these needs. A part of the system was designed and installed at
Langley by the Boeing Commercial Aircraft Group and another part by the Lock.lteed Aer-
onautical Systems Company.
The combined loads test machine (Figure 19.74) is the primary facility of COLTS. It can
apply combi ned mechanical loads, pneumatic pressure and thermal loadings to broad classes
of aerospace str uctures. It has the capability to apply a 2.700 kip (-12,000 kN) axial load,

•.
--- , ~---- -
.
6
4
HEAlER]/ lhKrno<ouplt cs Axiol j
SHEll / Releun,e Point

Figure 19.72 Technion experimems on buckling of cylindrical shells under C·ombined axial preload,
heating along two genei'ators and torque- a typical photograph taken during a heating
test (from r19.149])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Thermal Buckling 1621

lf'C) ----
0
43
66
82
100

liS~~
129
137
145
ISJ
158
BUCKLIN G

Figure 19.73 Technion experiments on buckling of cylindrical shells under combined axial preload,
heating along two generators and torque-typical upper-generator photographs during a
tests for temperaiUres 0-158°C (from [9.178] and I 19.149])

a 600 kip ( - 2,670 kN) vertical shear load, a 6000 ft-kip (- 8, 130 kN m) lorsion load (torque).
and a 20 psi (- 140 kN/m 2) pressure load. Specimens will be tested at temperatures up to
400°F ( - 204°C} and at cyclic, spectrum fatigue loading conditions, a broad variety of types
of tests, that will naturally include thermal buckling experiments. 'TYpical test spec imens
would be curved 120 in. (-3.05 m) long and 96 in. (- 2.45 m wide) panels having a radius
of 125 in. (- 3.20 m), as well as complete shells, up to 540 in. (-13.7 m) long and 180 in.
( - 4.57 m) in diameter. This NASA facility is now in full operation.
The two additional views of the combined loads test machine presented in Figure 19.75
(from [ 19. 155]) reveal more details. The end view (Figure 19.75a) shows that the loading
platen is suspended from a gantry that can traverse f01ward or backward to accommodate
test articles of different lengths. Compression and bending loads are appl ied to the test
articles by six 450 kip ( - 2000 kN) hydraulic actuators located parallel to the specimen
(along the x axis), side shear load is appl ied by two 300 kip ( - 1330 kN) actuators located
along the y axis, and a torsional moment is applied by means of two 300 kip ( - 1330 kN)
actuators located in the z direction.
The combined loads test machine is located in a concrete pit that is approximately 32 ft
deep. 47 fl wide and 72 ft long (9.8 m x 14.3 m x 21.9 m). This arrangement is to ensure
that pressurized structural testing can he perfotmed safely. --`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 19.74 The NASA Langley new combined loads test machine fo r combined mcchanic.a i and
thermal loading tests of full-scale and sub-scale panels, fuselage shell s and wing boxes.
which is the main faci lity of the NASA Combined Loads Tests System (COLTS) (cour-
tesy of NASA Langley Research Center)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1622 Thermal Buckling and Cree'P Buckling

Loading platen
d}ii -~{iipca~:;e /Gantry

25 feet

(a) 36 feet

Plan view Track


~~
9: ~~

~
l
i' ':l.
Axial load
"-Lead screw

actuators-"""'\--- .i'
•Mi
UL
~
I

--
~
·':
p
$~'
~~mil
I

i:!: ~
-b. ~:1tli1~,
" ti ~3 .. Oriv e motor

-v ~ v
~Load reaction Loading platen_/~ ; ) Ij I
~
, platen

(b)
'
i::l!iilliE' ,___....!..:=: :-. '
1
!111 L -:-11
!

55 1eet

Figure 19.75 The NASA Langley new combined loads test machine for combined mechanical and
thermal loading tests on broad classes of aerospace structures (from [19.155]): (a) end
view of the loading platen, (b) plan view

The plan view of the test machine (Figure 19.75b) illustrates the drive motor and lead
screw mechanism that is used to reconfigure the test machine to test structures of 10-, 15-,
25-, 30-, 42- and 45-ft (~3-13.7-m) lengths. The interior length of the concrete pit is also
shown in the figure. The length of the axial actuators is extended by means of tubular
extensions to connect the loading platen with the load reacting platen in order to test spec-
imens that are longer than 10 ft ( ~ 3 m).
As shown in Figure 19.74, for curved panels tests a D-box test fixture (described in
[19.155]) replaces the test shells. As emphasized there, one of the design requirements for
the D-box test fixture is to provide proper support conditions to a curved panel, such that
the stress state in the panel will be the same as for the corresponding cylindrical shell at a
given load condition. This test fixture, shown in Figure 19.76, must have adequate radial
stiffness to support the pressure load, but should have a small axial stiffness compared to
the test panel. The small axial stiffness of the D-box test fixture allows the test panel to
experience most of the applied axial load. The low axial stiffness is accomplished by con-
structing the fixture from an assembly of curved I-beams, a cross-section of which is shown
in the inset of Figure 19.76a. The I-beam sections are 8.0 in. (~203 mm) deep and 15 of
these sections are used to construct the D--box test fixture. The axial stiffness of the D-box
test fixtures is 5 percent of that of the D-box test fixture assembled with the curved stiffened
paneL The axial stiffness of the curved panel is assumed to be 1.1 X 106 lb!in. ( ~ 1.93 X
105 kN/m), which is representative of the stiffness of a typical fuselage shell.
This D-box test fixture is designed to test curved panels with 60-130 in. (~1.5-3.3 m)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

radii and 20 and 22 in. ( ~0.51 and 0.56 m) frame spacings. The panels are attached to the
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Thermal Buckling 1623

(a) l·b.. m usembly

Hingo lilting-,

. '), I , ~
\~, Cross bar .I h j
'--~·-
(b) O·box

Figure 19.76 The NASA Langley new combined loads test machine-D-box test fixture for testing
curved stiffened panels (from [19.155]): (a) overall configuration, (b) cross-sectional view

D-box test fixture with the hinge fittings shown in the figure. Thirteen of these hinge fittings
are provided between the 1-beams for this purpose. A cross-section of the D-box, presented
in Figure 19.76b, shows the details of the hinge fittings.
When the D-box assembly is internally pressurized, the assembly expands in a manner that forms
an axisymmetric shape, which causes the hinge supports to move inward. This deformation will
result in the test panel bending in a way that is not representative of the response of an internally
pressurized shell. To prevent this undesirable deformation, cross bars are mounted between the
hinge points as shown in the figure, such that the distance between the hinge points can be held
constant or adjusted as needed to induce the appropriate stress state in the test panel.

Having become fully operational, the combined loads test machine has been employed to
verify the analyses and calculations recently performed on metal and composite shells and
curved panels under combined mechanical loads (as for example those of [ 19.1 55] and
[19.158]). Thermal buckling tests will follow and will also be related to earlier theoretical
studies (such as [19.150]-19.152]).
One of the pressure-box test machines has been completed and some test programs have
been performed or are in progress (see [19.157] or [19.158]). Figure 19.77 shows this pres-
sure-box test machine for curved stiffened panels subjected to internal pressure and biaxial
tension. It was designed to ensure that the appropriate boundary conditions are imposed on
the panel, to ensure that the stress state developed in the panel simulates that of a fuselage
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

shell. This design requirement is of particular importance for investigation of curved panels
and for damage propagation studies, as well as in the case of thermal loading, and has been
complied with in the test machine.
The schematic of the pressure-box test machine (presented in Figure 19.77b) shows more
details of the loading system. Axial loads of up to 7000 lb/in. ( ~ 1230 kN /m) are applied
to the panel by two 225 kip ( ~ 1000 kN) hydraulic actuators connected to a curved steel
plate (an axial load plate), located at each end of the panel. Pressure of up to 20 psig ( ~ 138
kN I m2 ) is applied to the concave side of the test panel, using a high-pressure air supply
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1624 Thermal Buckling and Creep Buckling

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
(a)

(b)

Figure 19.77 NASA Langley pressure-box te$t machine for combined illlemal pressure and other me-
chanical loading of curved panels. which is another facility of the NASA Combined
Loads Test System (COLTS) (courtesy of NASA Langley Research Center): (a) view of
machine, (b) schematic

source and a pneumatic control system. Ci rcumferential or hoop loads that develop in the
skin of the panel arc reacted by two additional steel plates (the hoop load plates) and two
steel rods that are connected to each side of the panel. Hoop loads th at develop in the frames
of the panel are reacted by steel rods that are connected to each end of the frames. Each of
these hoop load rods includes a turnbuckle, the length of which can be adjusted. to ensure
that the proper hoop loads are introduced in the panel frames and skins. The reactions in
the hoop rods are measured by means of load cells built into the rods. A continuous rubber
seal permits the panel to float freely when pressurized, while minimizing air leakage.
A stiffened composite curved panel (wi th four stringers and three frames) was tested in
the pressure-box test machine under internal pressure. The experimental resu lts of the panel
response were compared with those from finite-element analyses and from a nonli near struc-
tural analysis with STAGS (2.53], and good agreement was found.
Another pressure-box test machine is being developed at NASA Langley to test full-scale
reusable-launch-vehicle cryogenic-tank panels (see [ 19.157)).
This machine is equipped with an internal prCS$urization system. a cryogenic cooling system. and
a heati ng system to simulate the mechanical and thennal loading conditions that arc representative
of a reusable-Jauncb-vebicle mission profi le. The cryogenic cooling system uses liquid helium and
liquid nitrogen tO simulate liquid hydrogen and liquid oxygen tank internal temperature$. A quanz
lamp heating system is used for heating the external surface of the test panels to simulate . ..
external surface temperatures during re-entry of the launch vehicle. The pressurizat ion system uses
gaseous helium and is designed to be controlled independently of the cooling system.

The axial and hoop loads are reacted in a similar manner to that in the other pressure-box
test machine, discussed above. Only the seal is replaced here by one made of reinforced
Teflon material to wi thstand the pressure at cryogenic temperatures.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Thermal Buckling 1625

Two heater systems are needed to generate the test conditions for the panel, simulating
the re-entry of the launch vehicle. One is an array of quartz lamps to heat the external
surface of the test panel to a maximum of 1000°F ( ~540°C), and the other is a heater coil
inside the cryogenic box to simulate the soak temperature condition of 250°F ( ~ 120°C) on
the interior surface of the cryogenic-insulated test panel.
The NASA Langley tests carried out to date (see for example [19.158]) relate primarily
to composite panels under combined mechanical loads. However, thermal buckling experi-
ments and other tests that include thermal loads are also in progress at Langley (see for
example [19.157]). Many new results and new techniques for thermal buckling tests can
therefore be expected in the near future.
Another experimental investigation on thermal and mechanical buckling and postbuckling
responses of smaller, selected curved composite panels, which has been carried out at the
NASA Langley Research Center [19.156], should be mentioned here since it represents a
good example of current laboratory-scale test techniques for thermal buckling.
The test specimens were 10 in. (~254 mm) long, 10 in. (~254 mm) wide curved panels,
having a 60 in. (~1.525 m) radius and (Rit) = 1500. They were fabricated from eight
unidirectional layers of IM7 /5260 graphite-bismaleimide preimpregnated tape and had a
[ ± 45/0/90]_ quasi-isotropic lamination. Initial geometric imperfections were measured on
both outer and inner surface of each panel and averaged to yield the contour of the midsur-
face geometric imperfections. The panels were instrumented with back-to-back pairs of strain
gages and thermocouples (42 strain gages and 18 thermocouples). Two additional strain gages
were attached to a piece of quartz and placed in the insulated heated enclosure near the test
panel in order to measure the apparent strain of the gages at the test temperature, with which
the other strain gage readings could be corrected for their thermal expansion.
The test setup is shown in Figure 19.78. The panels were supported with stainless steel
curved end fixtures along the curved ends to provide clamped boundary conditions and with
stainless steel knife-edge supports along the straight edges to simulate simple supports. Some
of the panels were tested at room temperature, and on these panels the shadow moire tech-
nique was employed to observe radial deformations. End shortening was measured by three
direct current differential transformers (DCDT's), located at three corners of the load frame
platens.
The measurement of the thermal response of the panels, required both a load frame to
restrict axial end shortening and an insulated, heated enclosure to provide a uniformly dis-
tributed temperature increase. In this test rig (of Figure 19.78) "primary heating is provided
by coil resistance heaters with a fan (not shown in the figure) at the back of the enclosure
to circulate the heated air. Additional heating of the curved end fixtures was provided by
heated platens that consist of eight cartridge heaters embedded in a 0. 75 in. ( ~ 19 mm) thick
piece of stainless steel." The heated platens were controlled to ensure the same temperature
increase in the stainless steel curved end fixture as that of the air in the furnace. Two layers
of ceramics provided thermal insulation between the heated platens and the load frame
platens. Because a single large piece of ceramic insulation would tend to warp out-of-plane
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

when heated, the insulation was cut into small blocks, which deflected from their fiat and
parallel shape by less than the machining tolerances of the rig.
In order to accommodate the relatively large furnace within the load frame platens, the
elevated temperature tests were conducted in a large 300 kip ( ~ 1330 kN) capacity hydraulic
testing machine. End shortening was measured in the high-temperature tests by a combina-
tion of linear variable differential transformers (LVDT's) located inside the heated furnace
and one DCDT outside the furnace. Since the LVDT's were not thermally compensated, they
only provided a relative measure of end shortening, after a constant elevated temperature
had been reached.
Temperature was governed using four separate control loops that controlled the tempera-
ture of the top heated platen, the bottom heated plate, the side cartridge heaters, and the air,
which was heated primarily by convection from the resistance heaters and from the air
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1626 Th ermal Buckling and Creep Buc kling

CMamle lns.Aabon
I Heated platen
CUIYOCI Ot1d lixt\Ke

\ Loa~frame pf!~
' ' Coil resistance heotor 1
Cartridge heater
(a ) lnwlated box Knife edge suppon

(b)

Figure 19.78 NASA Langley laboratory-scale cxpcnrncntal thermal and mechanical buckling ~tudie<
on selected cun·ed composite panel\ tc\1 \Ctup: (a) schematic (from [ 19.156[). (b) ''c"
of a panel in test rig (courtC<) of NASA Langlc) Research Center)

circulating fan. The temperature wa< increa\Cd ~lowly 10 eliminate transient thennal effecls;
heating a tesl panel from room temperature to 400"F (-204°C) took aboul two hour~.
Tite effects of various experimentally de1enni ned conditions on the load-strain response
of the panel were studied with a geometrically nonlinear fini le-element analysis. This s1udy
showed 1ha1 1hc panel thickness, the inilia l geomelrie imperfections and the experimenwll y
determ ined boundary cond itions signi licarH iy aiTected 1hc predic1ed buckl ing loads. whereas
the measured thermal gradients had o nly lillie eiTecl.
This laborato ry scale experiment indicated thm the expertise gained in the developmem
of the more sophisti cmed large thennal loading experimenlal setups. discussed in Seclion
19.2. cou ld be applied also to small . simpler 1e"'·

19.4 Creep Buckling


19.4. 1 The Concepts of Creep and Creep Buckling
Creep. a time-dependent state of deformation of the malerialthat occurs even when 1he loads
acting on lhe structure remain constant was introduced in Subsectio n 19. 1.2 as o ne of the
primary undesirable cffec1s of a high-tempcralure cnvironmem. In the fifties, creep and creep
buckling were considered very impo rlant fac1or' in lhc design of high-speed ai rcrafl and
were therefore extensively s1udicd (see for example [ 19. 11-[19.3). [ 19.39). 119.64- 119.66]
and [ 19. 159 )- f 19. 164]). But when lhe resulls of lhese Sludics were relaled to design cri teria
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
of high -~peed aircraf1, creep was found to be less imponanl than predicted, lead ing for
Copyright example
Wiley to 1he de-emphasizing of NACA creep and creep huckling programs at the end of
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Creep Buckling 1627

the fifties (see [19.8]). However, in the decades that followed, creep and creep buckling
reoccupied their central position in aerospace research. In particular, in the hot regions of
power plants, creep played an important part, but buckling due to creep was not very likely,
since the structural elements of engines are usually not thin-walled. On the other hand, creep
characteristics significantly affected the growth and consequent clearance problems of hot
rotating parts.
Creep and creep buckling have been important factors also for the structural elements in
many other engineering applications that involve high temperatures, like steam turbines,
certain chemical installations (for example cracking of oil or the synthesis of ammonia),
nuclear power plants and some types of civil engineering structures (see [19.165] and
[19.166]). In the latter, creep buckling has often been found to have a very significant effect
even at moderate temperatures. This has been prominent in concrete structures, since concrete
exhibits substantial creep already at room temperatures. Such moderate temperature creep
buckling experiments on spherical domes are discussed in Subsection 19.4.4. Creep buckling
failures at moderate temperatures may also occur in pressure vessels made of plastic mate-
rials, or in some ocean engineering applications (such as the titanium alloy spherical hull of
the Woods Hole Oceanographic Institute submersible ALVIN [19.167]).
The first recorded experiments relating to creep date back to the 1830's when the French
engineer Vicat tested iron wires loaded by different weights and measured their deformation
over a period of nearly three years, stopping his tests only when the most heavily loaded
wire broke [19.165]. But the history of the systematic study of the engineering aspects of
creep actually began with Andrade's experiments in the first decade of the twentieth century
[19.168]. He introduced the concepts of primary, secondary and tertiary creep in the case of
uniaxial creep tests with constant load (see Figure 19.7), which are still the basis of creep
analysis today.
The physical and metallographic aspects of creep, as well as its structural effects, have
been widely treated in many books and papers (see for example [19.66]-[19.70] and
[19.169]-[19.174]). For our purpose, however, it suffices to summarize the main structural
effects of creep (as outlined for example by Hoff in [19.64] and [19.66]).
The obvious structural effect of creep is a change in the shape of the elements of the
structure. When the stresses and the temperatures are not uniform, the creep deformations
vary from element to element, with the result that the structure is distorted. Because large
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

deformations may accumulate, their magnitude must be calculated to ensure that they remain
within allowable limits and do not interfere with the proper functioning of the structure.
A second effect of creep is the change it causes in the distribution of the stresses in the
structure. In the presence of creep, Hooke's linear stress-strain relation is replaced by a more
complex one, in which time is a parameter and the stress enters in a highly nonlinear func-
tional form. The resulting stress distribution therefore significantly differs from that predicted
by the classical theory of elasticity.
A third effect of creep is the occurrence of a tensile instability phenomenon, necking in
a tensile creep test, as a result of which the tensile specimen subjected to a constant load
lasts only for a finite length of time.
A fourth effect of creep is another kind of instability in the presence of compressive
stresses, creep buckling, which is discussed below.
As pointed out by Hotl [19.64J,
Results of creep tests are usually presented in diagrams in which the applied stress is plotted on
double logarithmic scales against the steady (secondary) creep rate and against the time for fracture
to occur. With most materials the lines connecting the test points on these scales are approximately
straight, yielding an empirical relationship
s= deldt =(a-lA)" (19.16)

where s is the steady creep rate, t is time, u is the tensile stress, A is a constant, and n is a number
equal to the slope of the straight line plot of stress versus creep rate. This type of plot as well as
Copyright Wileythe power law were probably first used by Norton [19.169].
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1628 Thermal Buckling and CreE'P Buckling

The power law is often referred to as the Norton-Bailey law. Some typical values of n and
A for some aluminum alloys at elevated temperatures are shown in Table 19.2, from test data
presented by Dorn and Teitz [19.171].
It should be recalled that the Norton-Bailey law, Eq. (19.16), represents only the secondary
(steady) creep rate for uniaxial stress. For multiaxial states of stress more general formula-
tions are necessary (see for example [19.66] and [19.173]). Furthermore, for a more complete
representation, primary creep as well as the initial elastic or plastic strain rate should be
added, which complicates the analysis. But often the steady creep behavior is the dominant
effect, and then one can disregard the transient condition to obtain approximate predictions.
Many constitutive relations other than the Norton-Bailey law of Eq. (19.16) have been pro-
posed and employed with similar success (see for example [19.173]). For more details the
reader may turn to the vast literature on creep, samples of which were referenced above.
However, the main concern in this section is the instability in the presence of compressive
stresses and creep-creep buckling. The importance of creep buckling is due to the fact that
the presence of creep can cause a structure to buckle at a lower load than would be expected
from an elastic analysis, or even an elastic-plastic one. Buckling in the presence of creep
does not occur at once, but only after a period of time. Therefore, the decisive criterion in
creep buckling is the critical time rather than the critical load. The designer has to ascertain
that the critical time of the structural element that is prone to creep buckling exceeds its
design life.
The phenomenon of creep buckling carr be demonstrated by consideration of a simply
supported column (shown in Figure 19.79). To start (following for example [19.39], [19.64]
and [16.22]), one observes that every column is imperfect and that initially its center line
deviates slightly from the straight line along which the compressive force P is acting. The
initial deviation w 0 multiplied by the force P represents a bending moment, under which the
initial curvature of the column increases in consequence of creep. The increased curvature
is equivalent to a larger deviation w, which in turn results in a larger bending moment and
in more rapid creep. The consequence of this behavior pattern is that with time the column
bends to such an extent that it no longer can perform its structural function. When the
exponent n of the creep law, Eq. (19.16), is large, as for instance in most metals, the behavior
is typically nonlinear: the curvature increases very slowly at the beginning of the loading,
but after some time the lateral deflections become noticeable and soon grow more rapidly,
leading to the collapse of the column.
To simplify the analysis it is usually assumed that both the initial and final deflected
shapes are half sine waves and that the column is an idealized !-column (whose cross-section
consists of two concentrated flanges of area A/2 and a weightless but shear-rigid web of
depth h, as shown in Figure 19.79). As a further simplification, the inertia forces are usually
neglected, an error found by Kempner and Hoff [19.175] to be indeed small.
The analysis is now presented along the lines of Hoff's study in [16.22] and the presen-
tation in Chapter 6 of [19.70]. If the idealized column has an initial deflection of w 0 the
strains in the two flanges can be expressed by

Table 19.2 Typical values of creep constants 11 and A for aluminum alloys
Material Temperature oF 212 300 400
Temperature oc 100 149 204
{Exponent 11 90 38.2 10
6061-T6 A in ksi 42 42 46
A in kN/mm2 x 10' 290 290 317
{Exponent 11 30 14 8.5
2024-T3 A in ksi 81 100 71
A in kN /mm2 x 10' 558 690 490
--`,`,`````,`,````,,``,``,,`

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Creep Buckling 1629

A/2
6)
tM A/2
G
f.--n~

(b)

Figure 19.79 Idealized 1-column with initial deviation w,i (a) simply supported column, (b) idealized
cross-section, two flanges of area A/2 and a weightless but shear-rigid web of depth h

E1 = E0 + (w,u - W 0 cJ(h/2)}
(19.17)
E, = E0 - (w," - Wru.J(h/2)
where E 0 is a uniform compressive strain due to the axial load, w is the total deflection of
the column, E, is the strain on the convex (tension) side and E, is the strain on the concave
(compression) side. Subtraction of the second equation from the first, in Eqs. (19.17), yields
(E, - EJI h = W,u - w0_., (19.18)
Equilibrium of the stresses with the external force and moment yields respectively
F = (CT, + CTJ(A/2) = -P0 }
19
and M = (CT, + CTJ(Ah/4) = -P0 w (19. )

where P 0 is the axial force and P0 w is the moment P 0 applies at a given section of the
column.
Creep buckling under a constant load is a slow process, and therefore the behavior will
be primarily governed by the secondary (steady) creep. Hence, it is expected that a good
approximation will be obtained by using the Norton-Bailey creep law of Eq. (19.16), which
implies neglecting the elastic contribution. Substitution of Eq. (19.16) in Eqs. (19.19) yields
(AA/2)(£) 1" + £)1 ") = -P0 }
(AA/4)(£) 1" - £)1 ") = -P0 (wlh) 0 9 -20)
By solving Eqs. (19.20) simultaneously fore and e, and substituting the result in the time
derivative of Eq. (19.18), one obtains the governing differential equation of the column

a'(2wlh)
----'----':c- (2U/h 2 )(-P0 /AA)"{[1 + (2wlh)]"- [1 - (2w/h)]"} = 0 (19.21)
at·a(xl L?

Note that since the initial deflection w 0 is independent of time, w0 = 0, which simplifies Eq.
(19.21). Furthermore, for convenience, the variables w and x in this equation have been
nondimensionalized to (2wl h) and (x/L) respectively.
In the linear case, n = 1, and Eq. (19 .21) reduces to

a'( 2 ~~if- + (4UP,/AAh )(2wlh)


at · a x
2
= 0 (19.22)

Equation (19 .22) has the solution


--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1630 Thermal Buckling and Cree'p Buckling

w = a 0 (h/2)[exp(4.UP1/AAh 2 7T:>)t] sin(1rx/L) (19.23)


which satisfies the initial condition
w0 (x,O) = (h/2)a 0 sin( 7TX!L) (19.24)
where (a 0 h/2) is the amplitude of the initial deflection, assumed to be a half sine wave. The
solution of Eq. (19.23) reveals that in the linear case, as the timet progresses the deflection
grows without bound, and therefore there is no critical condition at any finite time.
For the case when the index of the power law n = 3, and with the same initial condition
Eq. (19.24), the governing differential equation, Eq. (19.21), becomes

a'( 2 w~h) + (4UP~/A'AW)[3(2w/h) + (2w/h) 3 ] = 0 (19.25)


at · a(x L) 2

which is a nonlinear differential equation. Hoff showed in [16.22] that a reasonably accurate
solution to Eq. (19.25) could be obtained by assumption of the one-term approximation
(19.26)
where a 1(t) is to be determined. By substitution of Eq. (19.26) into Eq. (19.25) one arrives
(as in [16.22]) at the differential equation
(daJdt) = (3UPM7TWA'A')(4a 1 + aj) (19.27)
This equation can be solved for t through separation of variables:

t = (113)(7Th/L) 2 (AA/ P 0 ) 1 f.' { l![a 1(4 +aT)]} da 1 (19.28)

Since in general

(19.29)

Eq. (19.28) can be integrated to yield


t = (l/24)(7Th/L)"(AA/P0 )' log{(arfa 0 n(4 + a~)/(4 +aT)]} (19.30)
Equation (19.30) can be rearranged to yield an expression for a 1(t) that was to be determined
[a 1(t)/a 0 ] = exp(t/2k) · [I + (a 0 /2)"[1 - exp(t/k)]}- 112 (19.31)
where
(19.32)
From Eqs. (19.31) and (19.32) one sees that the amplitude of the deflection a 1(t) increases
with time and that it becomes infinite at a critical time tcr, when the denominator of Eq.
(19.31) vanishes. This critical time is therefore
t" = k log[1 + (4/a~)] (19.33)
An example of creep buckling of a column, calculated by Hoff [ 16.22] with this simplified
analysis, is shown in Figure 19.80. The modulus of elasticity of the column E = 6.5 X 106
psi ( ~45 kN /mm 2 ) and its deformations are governed by the steady creep law
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

£ = (u/ A)" = w- 12u' (19.16A)


i.e. n = 3, A = 104 psi (~69 N/mm 2 ). The slenderness ratio is 2Lih = 100 and the initial
deviation amplitude is a 0 = 0.01(h/2). The figure represents creep buckling under a com-
pressive load that is 75 percent of the Euler load. The conditions chosen are similar to those
prevailing with an aluminum alloy at 600°F ( ~ 315°C). The figure shows the development
of the deflection with time, and it can be observed that the column snaps through in about
41 seconds after the load application.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Creep Buckling 1631

100

80

60

40
I
vvI ,_.w
,_.0
,_~-

--
20

"'
<(

10 20 30 40 50
t, SEC

Figure 19.80 Creep buckling of a column-development of central deflection with time, calculated by
Hoff (from [16.22]); E = 6.5 X 10" psi (~45 kN/mrn 2 ), the creep law is i; = (a-/10 4 ) 3
/hr, the slenderness ratio 2 Ll h = 100, the initial deviation amplitude is a 0 = 0.01(h/2)
and PIP1 = 0.75

From the solutions for n = 1 and n = 3, Kraus (Chapter 6 of [19.70]) noted the following
characteristics of the creep buckling process (of columns):
1. Creep buckling is distinguished by deflections or velocities that increase beyond all
bounds.
2. Creep buckling can occur at a finite time only for a nonlinear creep law.
3. Creep buckling will occur at any axial compressive load, no matter how smalL Its
examination will show if the critical time is less or more than the design life of the
structure.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

4. Creep buckling will occur only if the column has initial imperfections, otherwise
tcr --+ oo, but in real structures there is always an initial imperfection.
5. The critical time depends strongly on the axial load, but less on the initial shape.
6. The small deflection theory employed is really not valid near tw as the deflections are
growing rapidly, but it roughly indicates the creep buckling behavior.
7. The Euler load for instantaneous buckling does not appear in Eq. (19.33) for tw since
by using the steady creep power law, Eq. (19.16), the elastic strains were neglected in
the analysis.
The above analysis is rather simplified, although it suffices to demonstrate creep buckling
behavior, and many more sophisticated studies have been carried out since. First the creep
law was modified to include the elastic deformations, replacing Eq. (19.16) with
i; = (iJIE) + (a/ll)" (19.34)
and later also primary creep was included (see for example [19.163] and [19.164]). In the
fifties, sixties and seventies, creep buckling was widely studied (see for example [19.160]-
[19.164] and [19.177]-[19.181]), including also creep buckling of plates and shells (see for
example [19.167] and [19.182]-[19.185]). Relevant computer codes were also developed
during that period. However, though less intensively, theoretical creep buckling studies con-
tinued into the eighties and nineties (see for example [19.186], [19.187] and [16.35]).

19.4.2 Early Creep Buckling Experiments on Columns


The extensive theoretical and numerical studies of creep buckling in the fifties and sixties
were accompanied by a considerable experimental research effort. Many column creep tests
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1632 Thermal Buckling and Creep Buckling

were carried out in that period at the NACA Langley Research Center (see for example
[19.188]), the Polytechnic Institute of Brooklyn (see [19.180] [19.189]), the Battelle Me-
morial Institute in Columbus, Ohio [19.190], Stanford University [19.191] and [19.192], the
Aeronautical Research Institute of Sweden (FFA), Stockholm [19.193] and at other labora-
tories.
Systematic creep buckling experiments on columns commenced in the early fifties, when
it was realized that the test data available for evaluation of different theories were very scarce
indeed. At NACA Langley it was then concluded that "a great deal of experimental work
is required before sound procedures can be established for the design of columns which may
be subjected to creep" [19.188]. As a result, a comprehensive test program on creep buckling
of aluminum alloy columns was initiated, partly at NACA Langley Research Center (see
[19.188]) and partly at Brooklyn Polytechnic Institute under NACA sponsorship (see
[19.189]).
In both test programs the specimens had the same rectangular cross-section, 0.5 X 0.375
in. (~12.7 X 9.5 mm), but in the Brooklyn tests the material was 2024-T4 and alliS columns
had an effective length of 12 in. ( ~ 305 mm), whereas in the Langley tests the aluminum
alloy was 7075-T6 and the 54 columns had different effective lengths 3.18-11.9 in. ( ~80.8-
302 mm). The essential parts of the apparatus in both cases were a lever type of loading
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

mechanism (dead weight) designed to maintain constant axial compressive loads in the col-
umn and an oven capable of producing a constant elevated temperature.
The Brooklyn Polytechnic setup [19.189] was the smaller and simpler of the two. Here
the axial compressive loads were applied to the column by means of a lever of !-section
with a mechanical advantage of 5. A loading extension projected from the lever into the
oven, where its hardened end surface applied a load to the column. The ends of the column
specimens were fitted with adjustable knife-edge fixtures. Because these fixtures were stiffer
than the column itself, the length of an equivalent column was slightly smaller than the
actual distance from knife-edge to knife-edge. But calculations showed that for the columns
tested the difference was negligible.
The knife-edged fulcrum of the loading lever pressed upward against a V-groove that was
rigidly fastened to the frame of the apparatus. At the other end of the lever, weights were
suspended from a fixture that rested with a knife-edge in a hardened V-groove attached to
the lever. At this loading end of the lever a special device, with a calibrated loading screw,
was provided to allow close control over the variations in the column load during the heating
process and over the rate of application of the load.
This was necessary since as the temperature of the column increased during the heating
process, its length also increased and pushed the lever upward. This upward push of the
column decreased the load on the cantilever, unless the loading screw was readjusted. For
this reason the load reading of the cantilever was followed carefully and the screw was used
to maintain a constant load on the column.
The oven, which consisted of an isolated box with 10 ring heaters, was designed to
produce constant temperatures in the range from 150-IOOOoF ( ~65-538°C), with a tolerance
of ± 3°F ( ~ ± 1.7°C) along the column. Iron-constantan thermocouples were employed to
measure the temperatures. In the preliminary tests five pairs of thermocouples were arranged
along the column to obtain a detailed survey of the temperature distribution, in order to
optimize the position of the baffles in the oven. The experience gained showed that the fan
in the oven was not necessary and that it was easy to maintain the desired uniformity (of
± 3°F) along the columns. Hence, in the final tests only three pairs of thermocouples were
attached to each specimen. The tests were carried out at a nominal temperature of 600oF
( ~ 3l5°C), generally considered to be above the permissible range for aluminum alloy but
intended here to emphasize the creep effects.
In the final tests (on 15 columns) the load was brought up rapidly to a maximum prede-
termined value, smaller than the experimentally determined instantaneous critical load at
600oF (~3l6°C), of 1099 lb (~4890 N). Two series of tests were carried out, one with an
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Creep Buckling 1633

axial compressive load of 905 lb ( ~4025 N) and one with a load of 980 lb ( ~4360 N), 82
percent and 89 percent of the instantaneous critical load, respectively. The load was main-
tained until the column buckled. The time to failure was measured with a stopwatch.
The results of the 15 tests are presented in Figure 19.81, where the critical time is plotted
in minutes versus the amplitude of the first harmonic of the initial deviation, for the two
axial compressive load values. A study of the figure shows that 9 out of the 15 test points
can be connected with simple smooth lines that are similar in shape to theoretical buckling-
time curves, like those presented in [19.160]. The remaining five points are not very far
removed from the curves, but their critical times differ sufficiently from the trend to deserve
some comments. Two of these five points, those representing specimens 14 and 15, deviate
from the upper curve intentionally. These tests were carried out in order to ascertain the
effect of the length of the time for which preload was applied. In test 14 the preload was
300 lb rather than the usual 100 lb, whereas in test 15 it was only 35 lb, resulting in shorter
and longer critical times, respectively. The digressions of the remaining three irregular points
(2, 6 and 7) may have been partly due to the higher harmonics of the initial deviation from
straightness (for a more detailed discussion see [19.189]).
In conclusion, it was stated that the tests showed the critical time to decrease with in-
creasing initial deviations from straightness, with increasing compressive load during the test
and with increasing values of the preload, provided all other parameters were kept constant.
The parallel test setup at NACA Langley (see Figure 19.82 and [19.188]) was a little more
sophisticated, though essentially similar to that used at the Polytechnic Institute of Brooklyn.

10 o,---

,.r--
~-
-~-- ------
- --=t=---=-±---
~~~-=~~
------r-----+~
~~---~--.--

07

=+=-~-
'/!3

o__'_\< --l~
!
---·--
~-
\
)(14 t\(12 I
-r---
~
I

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

)(II~ ~( 10
I
I 06

I
"'
'( ~~
I0
~/g

h'\ "-
;~/~8
\

r-
\ '

~ 02
----
~ 0 TESTS AT P-9BO LB

r
~
)( TESTS AT P= 905 LB

~
~I
ll__ j
oo2 .oo4 .oo6 .ooe
AMPLITUDE OF fiRST HARMONIC OF DEVIATION
.oJo .o12 .014
fROM STRAIGHTNESS, IN.

Figure 19.81 Early creep buckling experiments on columns at the Polytechnic Institute of Brooklyn-
critical times of columns versus the amplitude of the first harmonic of initial deviation
from straightness, for the two test series, one at P = 905 lb (82 percent of the instan-
taneous critical axial compressive load at 600°F) and one at P = 980 lb (89 percent of
it) (from [19.189])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1634 Thermal Buckling and Creep Buckling

Figure 19.82 Early column creep buckling experiments at NACA Langley Research Center-test
equipment (from [19.188))

The column was placed in fixtures supported on kni fe-edges arranged so that the axes of
rotation of the fixtures coincided with the ends of the specimen (see Figure 19.83). The
same equipment was used to support the specimen in both the static and the creep tests;
however, for detertnining the static strength , load was applied by the hydraulic testing ma-
chine, and for determ ining creep behavior. a dead-weight apparatus (a lever type of loading
mechanism, shown on the left in Figure 19.82) was used to apply constant load.
Note that in Figure 19.82 the furnace is in the test position, hiding the column and its
end fixtures. The electrical power supplied to each of the three horizontal banks of strip
heaters in the furnace could be controlled individually, as well as that supplied to the heaters
built into each end fixture. An automatic control device perm itted the temperature of the
columns tO be maintained within ±3°F (- ± I.7°C) at the test temperatures. The speci men
temperature was obtained with the aid of an automatic temperature recorder and iron-
constantan thermocouples that were fastened by spring clips at the midheight and near each
end of the columns.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
The column test specimens, with the same rectangular cross-section as those of the parallel
Brooklyn tests (0.5 X 0.375 in.), were machined from a 0.375 -in. thick 7075-T6 plate. The
magnitude of the initial central out-of-straightness was measured for each speci men by a
dial gage reading to 0.000 I in. Compressive strength speci mens were machined from the
same plate.
A special co lumn-alignment procedure was employed at the initiation of the tests. and the
column alignment was then checked periodically. In all the tests the furnace and fixtures
were initially stabilized at the test temperatures. The specimens were then placed in the
fixtures and exposed to the test temperature for 30 mjnutes prior to appl ication of load. For
the compressive stress-strai n tests and static tests of the colunms, the load was applied by a

Figure 19.83 Early column creep buckling experiments at NACA


Langley Research Center-sectional view of a col-
umn end fixture (from [19.188])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Creep Buckling 1635

hydraulic testing machine at a rate of average strain of 0.002/minute. The results of these
tests were the basis for the choice of the fixed compressive load in the creep buckling tests.
In the column creep tests, after the specimen had been exposed to the test temperature
the fixed dead load was applied (by the lever system) at a rate requiring approximately 20
seconds for the total load to bear. The load was maintained on the specimen until collapse
occurred, and the lifetime was measured from the beginning of loading and ranged, in gen-
eral, from several minutes to a few hours.
The results of the rather systematic column creep buckling tests were analyzed in different
ways, with the aim to obtain some design recommendations. For example, in Figure 19.84
the results of the creep tests are presented in terms of stress plotted against lifetime on a
logarithmic scale-a method similar to that used in analyzing material creep or fatigue data.
Over the range of lifetime covered by the test results, the curves are assumed to be straight
lines. The symbols on the vertical axis indicate the magnitude of the tangent-modulus
stresses, and application of the tangent-modulus stress is assumed to imply that the column
lifetime is zero. It is evident from the grouping of the limited amount of test data that a plot
of this nature permits a reasonable prediction of the column lifetime. The figure also indicates
that a small variation in stress produces a greater change in the lifetime of long columns
than of short columns.
This presentation, and some of the other ones in the report, provided some useful design
information for prediction of column lifetimes.
Further study and extensive tests at the Polytechnic Institute of Brooklyn [19.180] were
intended to provide a practical procedure for determination of collapse or critical time of
aluminum alloy columns subjected to creep, while also evaluating the more sophisticated
theory of [19.163] and [19.164]. The experimental setup and procedures for these tests were
improved versions of the earlier Brooklyn Polytechnic test techniques. The method developed

Test dolo

L
50x 103 p

-
-30----a....._r--_ 300 'F
40
a
-5o a
30

20 --
<> ---=--
-70
10 =to

0 .I 10 0 .I 10
i'f.
psi
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

16 xtd 8xld
soo'F soo'F

12

8
6

4
l~ 50
70
' 100
a

<>

0 I 10

I, hr

Figure 19.84 Early column creep buckling experiments at NACA Langley Research Center-lifetime
curves for 7075-T6 aluminum alloy columns (from [19.188])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1636 Thermal Buckling and Cre•~P Buckling

for prediction of the collapse time gave good agreement with the experimental results ob-
tained at Brooklyn Polytechnic, as well as with those obtained earlier at the Battelle Me-
morial Institute [19.190]. The experimental evaluation of the Hoff-Fraeijs de Veubeke theory
(of [19.163], and [19.164]) was, however. not conclusive, which motivated another experi-
mental creep buckling investigation by Madsen and Lempriere at Stanford University
[19.192].
The Stanford test program was designed to emphasize the influence of plasticity on creep
buckling and thus to accentuate the difference between the simple theory of creep collapse
and the more rigorous Hoff-Fraeijs de Veubeke theory. A buckling machine was designed at
Stanford, drawing on the experience of the different versions of the Brooklyn Polytechnic
test setups [19.189] and [19.180].
The machine was designed and constructed to perform the following task: dead-weight
loading, up to 8,000 lb ( ~ 35,600 N), of pin-ended columns, 4-16 in. ( ~ 102-406 mm long),
heated up to 500°F ( ~260°C) with a temperature distribution uniform to within ± 1op
( ~ ± O.SOC). Instrumentation for recording both the lateral deflections of the midpoint of the
columns and the load history was considered necessary to permit an evaluation of the ec-
centricity of the columns in the hot state at the start of the buckling tests.
The buckling machine is shown schematically in Figure 19.85 and described in detail in
[19.192]. "In principle it consists of a double-lever weighing system (of ratio 50:1) which
uses stellite lathe-tool bits for knife-edge bearings, and a loading lever which is adjustable
by a screw and wheel. In order to simplify the column alignment procedure the loading
knife-edges (representing pinned ends) are fixed to the machine so that only small hardened
steel pads with a shallow V have to be attached to the column. This also reduces some of
the rigid end-effect introduced by the usual end-piece arrangement."
The main weighing lever was designed with a view to keeping all three pivot axes in a
plane and giving the main fulcrum adequate torsional rigidity. This was achieved by splitting
the fulcrum bearing into two knife-edges, one at each end of a cranked cross beam (shown
on the right side of the figure).
The lower weighing lever is essentially hanging from the main lever since it requires a downward
reaction at its fulcrum. This is compensated for by a large balance weight on the main lever which
is adjusted so as to just keep the fulcrum in contact in the unloaded state.
The link between the two levers is made through knife-edges, one of which is permitted freedom
to rotate in the plane of the edges by a cross pivot. This consists of a semi-cylindrical bar fastened
to the plate carrying the lower knife edge .md bearing on the plate to which the link rods are

St'.relo' Adjustment for


Leveling Lever System ® Denotes knife-edge-pad bearings

Support in line vith Load


Point to minimize Side Motion

Main Lever
Support Cranked
to keep Bearings
and Load Point
Level

Hight Side

Figure 19.85 Stanford University column creep buckling experiments~arrangement of buckling ma-
chine (from [19.192])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Creep Buckling 1637

connected. The arrangement allows for such imperfections as alignment errors between the planes
of the two levers or unsymmetric construction of the link, and prevents the application of a torque
to the main lever. The lower lever is provided with two small micro-switches which indicate the
horizontal position, a rigid stop just beyond this position so that arbitrary loads could be applied,
and a water dashpot to prevent oscillations. (The lever system proved to have considerable elastic-
ity.)

The reaction lever carries a knife-edge at the loading end to form the upper pin-end of
the column, and is supported by a 1.25-in. ( ~32-mm) diameter shaft and ball race. It has a
cutaway in the tensile flange for a load link that uses strain gages. Several sizes of links are
used, each for maximum sensitivity in a given load range. The link and lever flange are both
serrated to give adequate rigidity. The strain gages form a complete bridge and are directly

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
connected to an amplifier and recorder.
The lever system of Figure 19.85 is a typical example of a precise dead-weight system,
as required in creep buckling tests, and therefore warrants the detailed description presented.
A special deflection measurement device was designed for recording of the true lateral
displacement of the center of the columns relative to the load line. The heating system for
the column tests consisted of a standard design split cylindrical oven, which (together with
additional small heaters on the end pieces of the columns) provided a uniform temperature
distribution, within l oF ( ~O.SCC), over the length of an 11 in. ( ~ 280 mm) column.
It was decided to measure the creep law and stress-strain curves of the material in pure
compression. A special piston-cylinder type fixture that could be used in the buckling ma-
chine was therefore built, the details of which are presented in [19.192].
The specimens had the same cross-section as the columns tested earlier at NACA Langley
and Brooklyn Polytechnic Institute, were made of 2024-T4 aluminum alloy and were fully
annealed. Three groups of columns were made, of lengths 11, 6.35 and 5.5 in. ( ~280, 161
and 140 mm), respectively. They were tested in creep under axial loads, representing 38.7,
19.4 and 14.6 percent of the corresponding instantaneous (Euler) loads, respectively.
The experiments, the measurements and the data reduction were carried out with great
care. However, the experimental scatter still exceeded the small differences between the
simple creep theory and the more rigorous Hoff-Fraeijs de Veubeke theory, and hence the
evaluation of the improvements due to the latter was again not conclusive. The results and
their analysis point out the difficulties of creep buckling experiments and indicate the broad
material spectrum necessary for better correlation with theory.
Another typical experimental study on creep buckling of aluminum alloy columns was
carried out at the Aeronautical Research Institute of Sweden (FFA) in the mid-sixties
[19.193]. The experimental setup consisted of a loading device, fairly similar to the usual
universal testing machine static compression fixture, placed in one of the creep testing ma-
chines at FFA. To obtain simple supports, the columns were attached to a pair of cylindrical
rollers, fitted with adjusting screws for positioning of the column (see Figure 19.86); a rather
simple type of end fittings, which, however, proved to be quite satisfactory. With aid of the
strain gages, located as shown in the figure, the eccentricity of the specimens could be well
defined.
The experiments included 21 creep buckling tests on 2024-T columns with rectangular
cross section and 4 tests on similar-size T-section columns made of the same material. The
columns had different slenderness ratios (,\ = 37-90). The creep buckling tests were per-
formed with a constant axial load and at a temperature of 250°C ( ~480°F). In some of the
tests the emphasis was on a very carefully adjusted initial eccentricity.

19.4.3 Creep Buckling Tests on Plates


In the study of creep buckling of plates it should be recalled that the buckling stress of a
perfectly elastic flat rectangular plate is not the maximal value of the stress that the plate
can support (which is significantly larger because of the stable postbuckling behavior of the
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1638 Thermal Buckling and Cre•~P Buckling

Figure 19.86 Creep buckling tests on aluminum alloy columns at the


Aeronautical Research Institute of Sweden (FFA)-
c:ylindrical roller-type end fittings employed for alignment
of the specimens (from [19.193])

plate, discussed in Chapters 2 and 8), but only the value at which large, visible deflections
develop in it. Similarly, a fiat rectangular plate, whose material creeps, does not collapse and
become useless when the critical time is reached (as occurs in creep buckling of columns
or most shells). If application of the load continues, the load is redistributed in such a manner
that the compression increases near the edges and decreases near the middle of the plate (the
redistribution that is the basis of von K~trman's effective width, discussed in Chapter 8).
After this redistribution of load, the plate can support it for a long period of time. The
significance of the critical time in rectangular fiat plates subject to creep differs therefore
from that in columns. For plates, the approach of the critical time indicates only the occur-
rence of large visible deflections. It is, however, still important, since beyond the critical
time the lateral deflections grow at an increasing rate, which eventually may render the plate
useless for engineering purposes.
Since the consequences of creep buckiing in plates are less severe than in columns or
shells, the study of the stability of plates in the presence of creep was less intensive (see for
example [19.198]-[19.205]), and very few early tests can be found in the literature (for
example [19.194]-[19.197] and [19.201]). Different approaches have been employed for the
analytical prediction of creep buckling times of plates: a statical approach [19.198], a dy-
namical approach [19.202], a critical strain approach [19.201], an approach based on a
variational theorem for creep, which is an analog of Reissner's variational theorem in elas-
ticity, [19.199] and [19.200], and an approach based on the existence of initial deviations
from perfect flatness [19.203]-[19.205], in which an idealized sandwich model of the plate
(an extension of the idealized !-section idea used in columns) was employed.
The earliest creep buckling experiments on plates were apparently carried out at NACA
Langley in the mid-fifties [19.194]-[ 19.197]). The test specimens were aluminum alloy plates
and skin-stringer panels, and their unloaded edges were supported in V-grooves. As pointed
out in Chapter 8, such supports arc a reasonable rough approximation to simple supports
only for small buckling deformations. Hence, in later tests investigating the postbuckling
behavior in the presence of creep, with irs accompanying large lateral deformations, other
methods that provide more accurate simple supports were employed, similar to those used
in the clastic postbuckling experiments discussed in Chapter 8.
The experiments carried out by Mathauser and Deveikis on machined 2024-T3 aluminum
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

alloy plates [19.196] are a good example of the early Langley plate creep buckling tests.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Creep Buckling 1639

The purpose of the investigation was to obtain experimental data on both compressive
strength and creep lifetime at elevated temperatures. Forty-eight 20-in. ( - 508-nun) long,
0.0625 in. (- 1.6-mm) thick plates, of width-thickness ratios (blr) = 20, 30, 45 and 60, were
tested for creep buckling at 400"F, 450°F and 500°F (- 204°C, 232°C and 260°C). Compres-
sive strength tests were also can·ied out on 45 similar plates at temperatures ranging from
room temperature to 600°F (- 316°C), and add itional smaller 2024-T3 specimens for deter-
mining the compressive stress-strain properties of the material were also machined from the
same sheet and tested.
T he plate test setup, shown in Figure 19.87. includes the V-groove fixture for the unloaded
edges, the lever of the dead- load apparatus for the creep tests. upper and lower heated rams
and instrumentation for measurement of the plate shorteni ng. The V-groove support was
selected because at the time it was considered satisfactory and room temperature strength
tests with this type of support had demonstrated good correlation with experimental plate
strengths obtained from stiffened panel tests. Shorteni ng of the plates, in both the compres-
sive strength and creep tests. was determined by measurement of the relative motion between
the upper and lower loading surfaces. Th is relative motion was transferred by rods to linear
variable differential transformers (LVTD's) and recorded autographically (see Figure 19.87).
In the compressive strength tests unit shortening was recorded against load; in the creep tests
it was recorded against time. The furnace is not shown in the figure. During the tests the
temperature of the specimen was obtained at five equally spaced stations on the longitudi nal
centerli ne of the plate, with the aid of an automatic temperature recorder and iion-constantan
thermocouples fastened by spring clips. T he temperature of the plates was maintained within
± s•F (- ± 2.8°C) of the test temperature.
Before the tests proper, the plate specimens were aligned in the V-groove supports and
the furnace closed before heating began. Approx imately one hour wa.- required to heat the
specimen to the test temperature. For both the compressive strength and creep tests, the
specimens were exposed to the test temperature for one half-hour prior to application of

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 19.87 Early plate creep buckling experimentS a1 NACA


Langley Research Center- test setup, showing V-
groovc fixture, the lever of the dead-load apparatus,
upper and lower heated rams and inslrumenla-
tion for measurement of plate shorlcning (from
( 19.196J)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1640 Thermal Buckling and Cre1ep Buckling

load, with the exception of some specimens tested at 400oF ( ~204°C) to determine strength
after longer exposure times.
In the creep tests the applied load was maintained on the specimen until collapse occurred,
provided that failure occurred within six hours. Collapse was associated with a sudden drop
of the weight cage on the arm of the dead·load apparatus (not shown in Figure 19.87). When
collapse did not occur in this period, the specimen was unloaded and cooled to room tem-
perature. The creep test was resumed the next day for an additional six-hour period; however,
for this cycle, the half-hour exposure period was omitted. This procedure was continued until
creep collapse occurred.
The most significant information obtained from the creep tests is the lifetime. A typical
set of such results is the lifetimes obtained at 400°F (~204°C), shown in Figure 19.88. The
scatter is seen to be fairly small, and was similar also for the tests at the other temperatures.
For the creep buckling tests on the 2024-T3 aluminum alloy plates of [19.96], the plate-
creep lifetimes were predicted from the compressive creep behavior of the material. Com-
parison between the experimental and predicted plate-creep lifetimes yielded satisfactory
correlation for all the values of width-thickness ratio (bIt) and temperatures tested. Thus, it
appeared that the compressive creep behavior of the material might suffice for estimating
creep lifetimes of plates made of that material. However, as in similar empirical methods,
extensive experimental plate-creep lifetime data for other materials and geometries are
needed to validate the generality of the method.
Mathauser's other early plate creep buckling tests at Langley dealt with aluminum-alloy
multi-web box beams subjected to bending at elevated temperatures [19.194], and then with
aluminum-alloy skin-stringer panels under compression at elevated temperature [ 19 .195]. In
the former tests four nominally identical 2024-T3 multiweb box beams (see Figure 19.89b)
were loaded in bending (as shown in Figure 19.89a) in a furnace, two heated continuously
at 375°F and 425oF ( ~ 190°C and 2l8°C) and two intermittently at similar temperatures. The
creep test failures were tensile in three of the tests (in which the lifetime could be satisfac-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

L(o~I--~~~~-LL~~--L--L-L~~~
o,~- 10 so 100
Time, r, hr

Figure 19.88 Early plate creep buckling experiments at NACA Langley Research Center-creep life-
time results for 2024-T3 aluminum-alloy plates tested at 400oF ( ~204°C); on the vertical
axis, the symbols with the tick marks indicate the maximum compressive strength of the
plates after half-hour exposure to test temperature (from [19.196])
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Creep Buckling 1641

'lt 48
1ft, 96
J-'1
(a)

~-in. diam. rivets, ~-in pitch (see ncle)

i xi x~ angles, 755-T6

1
Linear toper
@webs, 24S-T3
2-! tip to
3!root
__l
~~==ml!F=z=~~,

(b)

Figure 19.89 Early creep bending experiments on multi-web box beams carried out at NACA Langley
Research Center (from [19.194]): (a) view of box beam showing support locations and
points of application of loads, (b) typical cross-section of box beam

torily predicted) and compressive by buckling in the fourth beam specimen (in which the
creep buckling failure could not be predicted at the time).
In the skin-stringer panel compression tests [19.195] 22 panels were tested for creep
buckling at 400aF ( ~204°C). The panels were of identical cross-section (shown in Figure
19.90) were made of 2024-T3 sheet and 2024-T4 Z-section extrusions, had bays of width
thickness ratio (bit) = 50 and four different lengths yielding nominal slenderness ratios
(LIp) = 18, 36, 64 and 92. All specimens were tested flat-ended without side support. The
lifetimes obtained from the panel creep tests (see Figure 19.91) show slightly more scatter
than the corresponding results from creep tests of plates (Figure 19.88). This could be ex-
pected because the initial out-of-straightness of the stringers, comparable with that of col-
umns, might have had a significant effect on the panel lifetime.
The compressive and tensile creep tests of 7075-T6 and 2024-T3 aluminum alloy sheet,
carried out at Langley by Heimerl and Farquhar in the late fifties [19.197], are included in
the list of early plate creep buckling tests, though they were not proper creep buckling
investigations but rather material creep tests. They are mentioned because the NASA TN
summarizing the tests presents a detailed discussion of V-grooves used in compression spec-
imens and provides other important hints for creep tests that could benefit an investigator
planning plate creep buckling experiments.
A decade later, Levi and Hoff carried out another series of plate creep buckling experi-
ments at Stanford University [19.204]. The primary aim of the tests was to verify their plate
creep buckling and postbuckling analyses. For reliable comparisons, the experimental bound-

.34!
1 2.00 I 2.00 ~

.14jtt 7-t-.040 l
i . .~7~.22
AN 470-3-3 rivets,~ -ln. pitch
16

Figure 19.90 Early creep buckling experiments on uniaxially compressed, aluminum alloy skin-stringer
panels carried out at NACA Langley Research Center-panel cross section (from
[19.195])
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1642 Thermal Buckling and Cre,ep Buckling

1 Srmbal
18 0

36 0

64 f---
92 .
0

~
~
~ I~

~ v""' ~ 0 pL
~
~ "~.
~ !-.......... 0
-~"' ~
........ "'--.., -......
~ ~
~
------......

92
-
I I I II I I I I I I I
,. 10

\.·"'
Figure 19.91 Early creep buckling experiments on skin-stringer panels can·ied out at NACA Langley
Research Center-lifetime curves for stiffened panels tested at 400°F ( ~204°C); the solid
symbols on the vertical axis indicate the maximum stress obtained in the compressive
strength tests for each slenderness ratio at 400oF (~204°C) (from [19.195])

ary conditions had to simulate closely those of the mathematical model employed, which
were simple supports. The Stanford investigators therefore tried several types of simple
supports for the uniaxially compressed rectangular plates.
One of these supports consisted of several hundred flat Inconel fingers that supported the
two long edges, (somewhat similar to the Cambridge fingers, discussed in Chapter 8; see
Figure 8.33), while the load was applied along the shorter edges by two rigid platens with
a V-groove. Experiments were carried out with these boundary conditions and the critical
time was successfully determined in them. The fingers permitted the plate to move in its
own plane but prevented any lateral motion, as required in the mathematical model. However,
when the deformations in the specimen became very large, wrinkles began to develop along
the long edges of the plate, between the points supported by adjacent fingers, making this
type of support unsuitable for postbuckling studies, since for appropriate postbuckling tests
the edges of the plate had to be kept perfectly within the plane of plates.
Hence, eventually the type of support chosen was two pairs of stainless steel knife edges,
bolted to a rigid fixture in such a manner that each pair constrained one long edge of the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Creep Buckling 1643

plate to remain in the plane of the plate, as shown in Figure 19.92. This type of support is
not ideal, but it is a reasonable practical approximation to simple supports also in the post-
buckling region and was employed in some of the plate tests discussed in Chapter 8 (see
for example Figures 8.18b, 8.43c and 8.70). It must be borne in mind, however, that this
knife-edge system will tend to develop friction along the long edges as the buckling waves
become very large.
In the mid-seventies Hoff [19.207] presented an assessment of creep buckling of plates
and shells and, with the aid of correlation with experiments, concluded that the available
methods of analysis are satisfactory and can be used by designers.
For plates, the comparison with experiment used the data obtained by Hoff and Benoit at
Stanford University [19.208] on square tubes with (bit) = 75, tested in compression at a
temperature of 600°F ( ~ 315°C). The 33 test specimens, of three different lengths, were made
by bending this 2024-T3 aluminum alloy sheet, of thickness 0.020 in. ( ~0.5 mm), into square
boxes. The use of this type of plate specimen was decided upon when two different sets of
edge supports designed, built and tested earlier at Stanford (discussed above, or see [19.204])
were found to develop undesirable restraints after the buckling displacement had reached
large values. Ideally, in the square tubes buckles can form alternately in the inward and
outward directions around the tube and the edges can act as frictionless hinges. Unfortunately,
the ideal simple-support condition also requires that the initial deviations from flatness follow
the same pattern, which did not always occur. Hence, the square tube is also not a perfect
simulation of simple-support boundary conditions.
In Figure 19.93 the relationship between stress and critical time for the plates tested is
compared with predictions for a creep exponent n = 6. The relatively small scatter indicates
the practicality of the analysis (presented in [19.203]-[19.206]). The similar comparison for
shells is discussed in the next subsection.
Creep buckling of plates did probably not present any serious problems to designers in
the last two decades, and therefore apparently no new plate creep buckling tests have been
reported.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

rigid support
fu:ture

Figure 19.92 Stanford University plate creep buckling experiments on axially compressed 2024-T3
aluminum alloy plates, with (bit) = 75 and aspect ratio 8. Support of the plates (from
[19.204] and [19.205]): (a) plan view showing the arrangement of the stainless steel
knife edges, (b) the lower and upper platens with the 3t deep fixing slots for the speci-
mens
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1644 Thermal Buckling and Cre.ep Buckling

o L ·11 in.
"' Bin.
• 6in.
I
1~~---L--Ll-L4_1L;L~7~8~91LQ____~I--~J_J4-LI~6J7J8~h1D'
Amplifiecl critical time faa,! Ia,- a)

Figure 19.93 Stanford University creep buckling experiments on axially compressed 2024-T3 alumi-
num alloy square tubes-comparison of predicted relationship between stress and critical
time with experimental values (from [19.207])

19.4.4 Creep Buckling Tests on Shells


The well-known unstable postbuckling behavior of most shells, which may result in violent
elastic or plastic failures, also has a strong influence on the behavior of shells in the presence
of creep. Creep buckling of shells is therefore a significant practical problem and has been
subject to extensive theoretical and numerical studies (see for example Section 6.3 of Chapter
6 and Chapter 8 of [19.70], and [19.167]. [19.182], [19.209]-[19.213] and [19.216] as well
as many experimental investigations (see for example [19.201], [19.213]-[19.215], [19.217]-
[19.229]).
The first systematic experimental study of shell creep buckling was probably that con-
ducted at the Polytechnic Institute of Brooklyn, under NACA Langley sponsorship, in the
late fifties (see [19.217]). The investigation dealt with thin-walled cylindrical shells loaded
by end moments and included tests on 102 shells made of 5052-0 aluminum alloy. The test
specimens included 54 unreinforced shells, 17 with longitudinal stiffeners, 15 with circum-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

ferential stiffeners and 16 with both longitudinal and circumferential stiffeners. The shells
were subjected to pure moment loading, mostly at an elevated temperature of 500°F
( ~204°C) and some at room temperature.
The primary objective of the experiments was to determine the influence of changes in
the ratio of skin thickness to cylinder diameter, and the effect of the number and placement
of reinforcements on time to failure in creep buckling. The effect of cylinder length was also
briefly investigated. The geometries of the specimens therefore covered the following ranges:
(Rit) = 125-250, (L/R) = 4.75-6. The stiffeners were of rectangular cross section (except
some 2024-T4 rings of circular cross-section) and were riveted to the shell. They were fairly
heavy, with an area ratio of about 0.4-1.2, and not very closely spaced.
The main parts of the test apparatus were a lever loading mechanism designed to maintain
a constant bending moment on the cylinder (see Figure 19.94), a rigid support of the cylinder
and an oven capable of producing a constant elevated temperature in the cylinder throughout
the range of 150-1000° ( ~65-540°C) within a tolerance of ± 3°F ( ~ ± 1.7°C). The rear wall
of the oven can be seen on the right-hand side of Figure 19.94.
The pure bending moment applied to the cylinder was the weight suspended from the apex of the
loading triangle multiplied by the distance from the apex to the pivot in the middle of the base of
the triangle [see Figure 19.94] plus the weight of the loading triangle multiplied by the distance
of its center of gravity from the pivot.
... Temperatures were measured at 20 points on the specimen. Iron-constantan thermocouples
were fastened to the specimen with heat-resistant tape in such a manner as to insure that the
junctions remained in close contact with the surface at all times. The thermocouples were wired
to a selector switch connected to a potentiometer, [and the temperatures at the 20 points were read
consecutively].
Of these points surveyed, four were at each extreme end of the cylinder, four at one-
quarter length, four halfway between the ends and four at three-quarter length. In each of
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Creep Buckling 1645

Figure 19.94 Early creep bending and buckling tests on thin circular cylindrical shells at the Poly-
technic Institute of Brooklyn- test setup: loading lever and counterweight (courtesy of
NASA Langley Research Center)

the planes one thetmocouple was located at the top, one at the bottom, and one on each end
of the horizontal diameter of the cylinder. The control points for automatic temperature
regulation were coincident with the survey poi nts halfway between the ends.
The de formations measured were (see Figure 19.95):

"I. Rotation of the end plane of the cylinder 0


2. Vertical displacement of the point of interseclion of the axis of the cylinder with end
plane l)
3. Axial displacement of the point of intersection of the axis of the cylinder with the end
plane f3
4. Change in vertical diameter in a plane perpendicular to the axis of the cyli nder and
bisecting the cylinder I:!.D
For measurement of the first three of these deflections, the loadi ng linkage was used
to transmit the motion of points in the end plane of the cy linder to the outside of the
oven where measurements could be made at room temperature."

MeasuremenL~ were made by using cantilever-beam transducers, the cylinder deflections were
then obtained from the geometry of the system (Figure 19.95). More details of these mea-
surements are presented in [ 19.2 17].

I .•.
-~
•.l -. .... .... ..
t-J.,

..

Figure 19-95 Early creep bending and buckling tests on thin circular cylindrical shells at the Poly-
technic Institute of Brooklyn- schematic drawing of equipment to measure deHections
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley (from ( 19.217))


Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1646 Thermal Buckling and CreE~P Buckling

"Measurements made in the first few experiments indicated that the magnitude of f3 was
small compared with that of 8 and 8, and in subsequent tests only the latter two quantities
were measured."
The change in diameter t.D in the elevated temperature tests was measured by means of
a high-temperature linear variable differential transformer.
The stationary transformer coil was suspended from the upper surface inside the cylinder. A long
slender rod was attached to the transformer core which was placed inside the coil with the end of
the rod resting on the bottom surface of the cylinder, so that the assembly could move freely in a
vertical direction. Relative motion between the core and coil as a result of the cylinder flattening
changed the output of the transformer. The change in output voltage was calibrated as a measure
of change in diameter.
. . . Many difficulties were encountered in measuring diameter change of cylinders at elevated
temperatures. Calibration of the transformer was awkward and stabilization of the recorder proved
troublesome. Because of these difficulties, the change in diameter was measured in comparatively
few experiments and the accuracy of the measurement was not up to the standard of that of the
other deformation measurements.
A typical creep buckling failure of a shell reinforced by both internal stringers and external
rings (type F) is shown in Figure 19.96. Shells of this type yielded consistent results, as did
practically all the other types. But agreement with predictions by a simplified theory was
good only at comparatively low values of the load, and poor at high values of the load. The
wide range and scope of the experiments, however, provided useful shell creep buckling
information to the designers at that time.
Some years later, Samuelson conducted, at FFA Stockholm, another systematic investi-
gation of creep buckling of circular cylindrical shells, which in this case were subjected to
axial compression [19.214], [19.218], [19.219] and [19.222].
The test specimens, shown in Figure 19.97, were small shells made of a Swedish aluminum
alloy with properties similar to those of 2024-T, with (Rit) ~ 32, 56, 75 and 110 for type
A, (Rit) ~56, 75, 110 and 150 for type Band (LIR) ~ 2.4-3.2 for all the test shells. They
were fabricated from two long tubes; first the inner surface was completely machined to
specifications, and then the outer surface was finished, after a wooden mandrel had been
fitted inside the shell to support the thin wall during the machining process. The material
properties in tension and compression were obtained from specimens cut from the same
tubes.
The general arrangement of the compressive creep testing unit used is depicted in Figure
19.98. It consisted of a loading frame (2), an electric furnace (3) and a lever using dead
weights (6) for load application. The furnace (3) could be lifted vertically for easy accessi-
bility to the loading frame. The furnace temperature was maintained at a constant level by
means of a temperature control unit governed by a thermocouple. With proper adjustment
of electric power, the temperature variations in the furnace would usually not exceed ± 0.5°C
(~ ± 1oF). The temperature distribution along the shell was continuously monitored by means
of three thermocouples, yielding measured differences within ± 1oc ( ~ ± 1.8°F).
The loading frame is illustrated in more detail in Figure 19.99. The test shells (1) were
placed between the two end plates (2 and 3), in which slots were machined to facilitate
location of the outer surface of the cylinders. Thus, clamped edge conditions were simulated.
The total axial deformation of the shells was measured by means of steel rods (4) and two
dial gages (5) placed diametrically opposite with respect to the test specimen. The dial gages,
which could read to 0.001 mm, were calibrated against electrical resistance strain gages and
measured the axial deformations within less than 1 percent accuracy.
An observation hole was provided in the furnace, as well as suitable illumination, per-
mitting monitoring of the buckling behavior and photographing the buckling patterns.
In a typical experiment the test shell was first loaded repeatedly at room temperature to
the same stress as was to be applied in the creep test, in order to allow it to settle and
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Creep Buckling 1647

(a)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(b)

Figure 19.96 Early creep bending and buckling tests on thin circular cylindrical shells at the Poly-
technic Institute of Brooklyn-typical failure of a shell reinfoneed by both stringers and
rings, experiment71 : (a) extemal view of failed shell (from [19.2 17)). (b) internal view
of same failed shell (courtesy of NASA Langley Research Center)

I
- H
!-"- , Jo.,.
I
I
i
I
I I
L
I - t-.!!. 170
I
l i
I r 30
1

i -'
9l8 J 8
(a) (b)
Figure 19.97 Creep buckling experiments on small aluminum alloy circular cylindrical shells subjected
to axial compression carried out at the Aeronautical Research lnstilule of Sweden
(FFA}-test specimens (from [19.2 18]): (a) type A, (b) type B

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1648 Thermal Buckling and CreE'P Buckling

Figure 19.98 FFA creep buckling experiments on small aluminum alloy circular cylindrical shells
subjected to axial compression-compressive creep test setup (from [19.218]): 1. test
specimen, 2. loading frame, 3. electric furnace, 4. lever, 5. adjusting screw and wheel,
6. weights, 7. knife edges

t
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 19.99 FFA creep buckling experiments on small aluminum alloy


circular cylindrical shells subjected to axial compression-
loading frame of creep test setup (from [19.218]): 1. test
specimen, 2. lower end plate, 3. upper end plate, 4. steel
rods for measuring the displacements of the end plates, 5.
Copyright Wiley
Provided by IHS Markit under license with WILEY dial gages
Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Creep Buckling 1649

deform plastically to accommodate local imperfections. The furnace was then heated while
in the upper position to about 20°C above the desired creep temperature, and subsequently
lowered to surround the loading frame.
When, after about 1.5 hours, the temperature in the specimen had reached steady-state
conditions, at the creep test temperature of 225°C ( ~440°F), the load was applied through
lowering the lever (no. 4 in Figure 19.98), by means of the adjusting screw (no. 5 in Figure
19.98). Then the chronometer was started and the initial deformation was recorded, which
marked the beginning of the creep buckling test.
The first test series [19.218] included 31 shells of types A and B (see Figure 19.97a and
b). Later, additional tests were conducted on 10 similar shells made from another batch of
the same aluminum alloy, which implied different creep constants, as measured in material
tests (see [19.219] and [19.222]). The test setup and furnace temperature in this second series
were the same as in the first one, but the focus in the later tests was on the influence of
significantly different stress levels on creep buckling.
Figure 19.100 shows some typical creep buckling curves for thin-walled and thicker-walled
cylindrical shells of the first test series. The end shortening versus time is presented in the
figure for three thin specimens (nos. 3, 4, 5 of wall thickness t = 0.42, 0.41, and 0.43 mm,
respectively) and three thicker-walled specimens (nos. 25, 26, 27 of wall thickness t = 1.39
mm for all three). All six shells of type A were of length 150 mm and were subjected to
the same stress of 12 kg/mm 2 ( ~ 118 x 106 pa), but the time is shown in minutes in Figure
19.100a and in hours in Figure 19.100b.

0.5
,--~~~-·T
I I
- --- - ·-f-~·
I I I
I
X

0.3
j 3
\)
! 4
I.> d-12.0
-
'

~
0.2

0.1 ----- - --- --- f - - - 1---· -~


1-

10 20 30 40 SO 60 70 t,min 90

(a)

1.2

~
~ 0·12.0

0.8

0.6
2~
~~
~
..... ~ ~
~
0.2 ~~

0
0 7 l.hr 9

(b)
Figure 19.100 FFA creep buckling experiments on small aluminum alloy circular cylindrical shells
subjected to axial compression-typical creep buckling curves, end shortening (IlL/ L)
versus time for type A shells (from [19.218] and courtesy of Dr. L.A. Samuelson): (a)
specimens 3,4,5, with (Rit) =Ill, 115, 110, respectively, (L/R)
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
3.2, and the time =
in minutes, (b) specimens 25, 26, 27, with (Rit) = 33 for all three, (LIR) 3.2, and =
Copyright Wiley the time in hours
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1650 Thermal Buckling and Creep Buckling

These creep buckling curves, and the similar ones for the other specimens (see [19.218])
reveal a close resemblance 10 an ordinary tensile creep curve (like Figure 19.7). with the
tertiary, accelerated creep being followed here, by instead of rupture. buckling. The secondary
creep is also, in the case of creep buckling, the mai n factor governing the critical time. For
buckling the accelerated creep period is of interest. It is usually preceded by a small buckle,
which grows rapidly induci ng a higher rate of end shortening that results partly from the
shortening of the shell due to the curvature of the buckles and pattly from the high stresses
in the buckled region.
At first only the postbuckling deformations were studied, but it was soon reali zed that
small defonnations could be detected before and at the onset of buckl ing. This led to a
decision to carry out visual inspections of the shells during the whole test. These inspections
showed that wh ile no buck les appeared duri ng steady creep, at the beginning of tertiary
creep, when the end shortening started to accelerate, a small perturbation or buckle occurred,
which would grow with an increasi ng rate till the buckli ng coll apse took place, as pointed
out above. Attempts to photograph this growth of the initial buckle were not successful,
except one series of 16 mm movie camera pictures. which qualitatively confirmed the process
but provided no funher quantitative data.
Typical creep buckling pallems observed are presented in Figure 19. 10 1, three examples
from the first test series (Figure 19.101a. b, c) and two from the second series (Figure
19.10ld, e). There are some differences between the creep buckli ng defonnations and those
observed in the elastic (or plastic) short-time buckl ing tests on simi lar shells that accompa-
nied the creep buckling tests (and are presented in [ 19.218]). First, the number of circum-
ferential buckles is less, by one or two, in the case of creep buckl ing. Since in elastic stability
the number of circumferential waves is essentially a function of the geometry only, the
number of circumferential waves in creep buckl ing must, therefore, also be a function of the
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

(ol {b ) (c)

( d) (e)
Figure 19.101 FFA creep buckl ing expe ri mentS on small aluminum alloy circular cylindrical shells
subjected to axial compression- typical creep buckling pauerns, for specimens with
(LI R) = 3.2 and applied stress of 12 kg/mm' (from [19.218]): (a) shell with (Rit) "'
I 15. (b) shell with (Rit)"' 60, (c) shell with (R il) = 33; for specimens with (L/R) =
2.2 and different stresses (from 119.2 19]): (d) shell with (Rit ) = 115 and applied stress
of 8 kg/mm2, (e) shell with (Rit) = 11 5 and applied stress of 4 kg/mm2
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Creep Buckling 1651

creep behavior of the shell. Secondly, in all the creep buckling tests the buckles developed
in the neighborhood of the boundaries, indicating that the boundary conditions have a greater
influence in creep buckling than in elastic buckling of similar shells.
In Figure 19.101 one can clearly see the influence of the wall thickness, or (Rit) ratio,
on the creep buckling patterns. The completely nonsymmetric mode for the thin-walled
specimen with (RI t) = 115 (see Figure 19.10la) changes to a partly axisymmetric mode for
the shell with (Rit) = 60 (in Figure 19.10lb) and becomes completely axisymmetric when
(Rit) = 33 (in Figure 19.10lc). The trend towards an axisymmetric pattern for thicker walls
is similar to that in short-time buckling, but for creep buckling it is more pronounced. The
two remaining pictures in the figure illustrate the effect of the applied stress in thin-walled
specimens with (RI t) = 115. Reducing the applied stress, from 8 kg/mm2 (in Figure 19.10ld)
to 4 kg/mm 2 (in Figure 19.10le), affects the buckling pattern in a similar manner as would
thickening the walls, i.e. strengthening the trend towards axisymmetry.
Samuelson also carried out creep buckling experiments on similar small aluminum alloy
shells subjected to bending (see [ 19.222]). A separate vertical loading frame was built for
the bending tests that fitted into the furnace used for axial compression. The test specimens
were manufactured with special flanges to permit application of both tensile and compressive
forces. The test setup consisted essentially of two heavy steel tubes, into which the test shell
was placed, the lower one attached to the foundation, and the upper one fitted with another
stiff tube to which the bending moment was applied and whose rotation was measured during
the test.
The creep buckling time of a cylinder subjected to bending was found to be much higher
than that for an identical shell subjected to axial compression of the same magnitude as the
maximum bending stress. This correlates well with similar observations for elastic bending
instability discussed in Subsection 9.7.2 of Chapter 9, where the primary reason for the
higher experimental buckling stresses in bending was given as the lower practical imperfec-
tion sensitivity in this case.
In the mid-sixties another series of creep buckling tests on very thin cylindrical shells
subjected to axial compression was performed by Carlson et al. at Stanford University
[19.224]. The specimens were 11 small electroformed nickel cylinders of an average wall
thickness t = 0.0032 in. ( ~0.08 mm) and an average (R It) = 384 and (L/ R) = 4.5. The
shells were tested at a temperature of 650°F ( ~ 340°C) under an axial compressive stress of
7.0-9.0 kips (~48-62 X 103 kN/m'). These very thin-walled specimens all failed by buck-
ling into nonsymmetric two-tier diamond patterns, similar to those usually obtained in elastic
buckling of thin circular cylindrical shells under axial compression.
A further series of creep buckling tests on aluminum alloy cylindrical shells subjected to
axial compression was conducted at the U.S. Army Materials and Mechanics Research Center
in Watertown, Massachusetts in the late sixties (see [19.223]).
The specimens were machined from one lot (to ensure the same creep constants) of 2024-
T4 2 in. (~51 mm) outside diameter aluminum alloy extruded tubing. Both inner and outer
diameters were machined to obtain cylindrical shells with (Rit) = 50 and 90 and (L/R) =
2 or 3. By careful alignment of the lathe and by the use of support jigs during manufacture,
the variation in wall thickness was kept down in most specimens to less than ± 0.001 in.
( ~ ± 0.025 mm), i.e. less than 5 percent of the wall thickness in the thicker shells and less
than 10 percent in the thinner ones, which indicates the difficulty of achieving uniform wall
thickness in small specimens. The test shells were annealed to the "0" condition prior to
testing.
The creep test, and the accompanying short-time tests, were performed in a precision
pneumatic testing machine (shown schematically in Figure 19.102), equipped with a tubular
furnace and a saturable reactor temperature controller. Specimen temperature, on the outer
and inner surfaces, was sensed by Chrome!/ Alumel thermocouples, whose junctions were
held in contact with the specimen by light springs (see Figure 19.102).
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1652 Thermal Buckling and Creep Buckling

Loaded T/C Bead


Fixed Ram Spherical not shown}
Compression Head

Stop

Figure 19.102 U.S. Army Watertown Research Center creep buckling tests on axially compressed
aluminum alloy shells-test setup, schematic (from [19.223])

The testing machine was used in the horizontal position to minimize temperature gradients
due to the "chimney effect," which would be present in a vertical installation unless com-
pensated for. Preliminary measurements on a cylindrical shell, which was specially instru-
mented with many thermocouples, indicated a maximum temperature variation of ± 1oF
( ~ ± 0.6°C). These preliminary tests were also utilized to optimize the spring-clip configu-
ration for the thermocouples and to adjust the temperature-control system for the creep test
temperature of 500°F ( ~260°C).
End shortening was measured by concentric tube compressometers with LVDT's that were
coupled to the compression rams (see Figure 19.102) and was autographically recorded. The
pneumatic testing machine could be programmed to apply a load at a preset controlled rate
and hold this load over long periods of time with less than 0.1 percent variation. This testing
machine was therefore suitable for creep testing.
Thirty specimens were tested for creep buckling at 500°F ( ~ 260°C) and applied stress
levels between 8,000 and 5,500 psi (~5,5000 and 38,000 k:N/m2 ), 20 with (Rit) ~ 90 and
10 with (Rit) ~50. Both axisymmetric and asymmetric buckling modes were observed. The
thicker-wall (Rit) ~ 50 cylinders all creep buckled axisymmetrically, whereas the thinner-
wall (R/t) ~ 90 cylinders buckled asymmetrically at stress levels above 7,500 psi (~5,200
k:N/m 2 ), either axisymmetrically or asymmetrically at applied stresses of 7,500 and 7,000
psi ( ~52,000 and 48,000 kN /m2 ), and axisymmetrically at stress levels below 7,000 psi
( ~48,000 kN /m 2 ). These results correlate well with those obtained in Samuelson's tests on
axially compressed shells, discussed above.
A few years later, Samuelson conducted, at FFA Stockholm, creep buckling tests on cy-
lindrical shells subjected to external pressure (see [19.214]). Considerable attention was given
to the initial imperfection in these experiments.
A special loading device was designed in which the external pressure was applied in a
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

narrow, annular chamber between the test specimen and the outer pressure vessel (see Figure
19.103). Thus, the inside of the cylinder was accessible for measurements and observations
during the creep test. For measurement of the radial imperfections and deformations, a dif-
ferential transducer was attached to a central measuring shaft, which could be rotated with
a constant angular velocity. The radial deformation or out-of-roundness could thus be re-
corded at any time during the test.
An electrical resistance furnace was employed to provide the test temperature of 225°C.
A number of thermocouples were attached to the testing equipment to provide a complete
mapping of the temperature distribution. The temperature control unit of the furnace main-
tained the temperature within ± 1oc.
The test specimens were again machined from aluminum alloy tubes (made of a Swedish
alloy with properties similar to 2024-T). The shells were similar to the ones used in previous
tests, with (Rit) = 79 and (L/R) = 2.1, but had heavy integral end rings (see Figure 19.104).
When a test specimen had been mounted in the loading device, the pressure to be applied
in the creep buckling test was first applied at room temperature in order to test the seals.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Creep Buckling 1653

Pr• uur•
\upp\y--1 •=---i

Pnuure
c: hQ mbt r _""1---4-...:::!~
- Tron\durtr

AotOhl'lg Shott
Furnoct

Figure 19.103 FFA creep buckling experiments on alumonum alloy circular cylindrical shells subjected
to external pressure- test setup with radial defonnation scanner (from [19.214J)

After that. the furnace was activated. and the testing temperature was reached after about
three hours. Before load application, the initial radial deflections were recorded. and sub-
sequently measurements were taken 111 regu lar intervals. The transducer used to measure the
radial deformations was calibrated at room temperature only. and therefore the absolute
values of the measured displacements at the elevated temperature could contain errors of the
order of 5- 10 percent. Seven creep tests were carried out. Five of the specimens had delib-
erately machined initial defects of nearly hannonic shape. One of them. test specimen DI-
NAH. is shown in Figure 19. 104 after buckling due to external pressure. The specimens with
predefonned imperfections of nearly harmonic shape were found to have a much shorter
critical time.
Other external pressure creep buckling tests on aluminum ci rcular cyli ndrical and truncated
conical shells were carried out in the early sixties at the Southwest Research Institute in San
Antonio. Texa.~ (see [19.227)).
Creep buckling of shells occurs in a broad variety of engineering applications. A recent
example is a problem that arose in deteriorated sewer pipes, which were repaired by instal-
lation of a tight-fi tting thin-walled polymeric lining inside the cracked or deformed host pipe.
Due to external groundwater pressure. creep buckling could eventually occur in these linings.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Figure 19.104 FFA creep buckling experiments on aluminum alloy circular cylindrical shells subjected
to external pre<>ure- typical test specimen DINAH after creep buckling (counesy of
Dr. L.A. Samuelson)
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1654 Thermal Buckling and Creep Buckling

which was studied theoretically and experimentally by Boot and Welch at the University of
Bradford in England in the mid-nineties [ l9.228]. Since the constrained creep buckling of
these repair linings is a highly nonlinear problem, a series of tests was conducted on typical
linings to evaluate their response to the typical external pressure loadings. The tests reported
(in [19.228]) were only short-term tests, not proper creep buckling tests, but they defined
the response and behavior of the linings, which facilitated better creep life predictions. Ex-
tension of the tests to longer times, i.e. performing actual creep buckling tests, which in this
case would have to be accelerated ones (since the expected lifetime of the linings is of the
order of decades), would no doubt lead to more reliable creep life predictions and is therefore
warranted.
Another example of creep buckling of polymeric composite cylindrical shells was inves-
tigated by Rikards and Teters at the Latvian Academy of Sciences, Riga, in the early sev-
enties [19.226]. The study focused on the changes of the axisymmetric buckling mode to a
nonsymmetric one, essentially an extension of the investigation of this transition in nickel
shells, carried out by Hoff and his students at Stanford University in the same period (see
[19.215], [19.213] and [19.225]). The experimental part of the Riga investigation included
40 fairly thick polyethylene (a nonlinear viscoelastic material) shells of (R/t) = 8-25 and
=

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
(L/ R) 4, loaded in axial compression. The initial imperfections were measured, as well
as the radial deflections with time after load application. It was found that the buckling
mode, at the critical time, differed significantly from the shape of the initial imperfection.
Contrary to similar metal shells, the polymer shells, though rather thick-walled, were never
observed to buckle in an axisymmetric mode.
A different engineering application in which creep buckling may be encountered is certain
types of nuclear reactors. An example of such creep instability in a spherical shell was
studied by Penny and Marriott at the University of Liverpool, in the late sixties. The structure
investigated was a hemispherical shell subjected to an inward radial load applied through a
solid boss.
The experimental study employed hemispherical shells, of 3 in. ( ~ 76 mm) radius and
(Rit) = 65, made from aluminum alloy in a half-hard state, taken from a single rolled billet
(to ensure identical creep constants). Extreme care was exercised in the machining of the
specimens because of their imperfection sensitivity, and the required accuracy was achieved
by the use of lathe attachments that permitted tool-following guided by precisely made
templates.
The shell static testing procedure consisted essentially of loading the shell through the
screw-driven cross-head of a universal testing machine. Measurements of the load through
a load cell and the cross-head movement were simultaneously recorded to produce a load/
displacement curve, with due allowance for the load cell flexibility. Throughout the loading
the shell was maintained at a temperature of 200°C ( ~ 390°F) to within + 1oc.
The heating arrangement consisted simply of a length of heating tape randomly placed
inside the shell in a loose roll, with the shell being thoroughly insulated on its outside by
layers of asbestos wool. Control of temperature was achieved by use of a platinum resistance
thermometer as a sensor, connected to a proportional controller (silicon controlled rectifier
type).
The same heating and temperature control method was used for the shell creep tests. The
loading this time was through dead weights applied directly to the shell boss. The loading
rate at the beginning of each test was carefully controlled by movement of the machine
cross-head. The movement of the boss relative to the shell base was continuously monitored
from the output of an LVTD, whose core was directly connected to the shell boss. Two shell
creep tests were conducted at load levels of 250-290 lb ( ~ 1,110-1,290 N).
Boss displacement curves for these two tests are shown in Figure 19.105. As can be seen,
the experimental results for the response of the shells correlate very well with prediction by
a relatively simple theoretical method proposed in [19.220], in particular for the lower load
level of 250 lb. Physically, the experiments and theory demonstrated that a metal spherical
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Creep Buckling 1655

boss
displacement
3
( 10- inch•s) 10
r---~~~~-+--~--~~~+--

I!!'Xperimental result
- - - theore-tical prediction
0~--~--~---L--~~----~--~--~----"
20 40 60 80 100 120 140 160
time (hours)

Figure 19.105 University of Liverpool creep buckling tests of boss-loaded spherical shells-boss dis-
placement of boss-loaded shells during creep at 200oc (from [19.229])

shell in the creep range would snap through in much the same way as when the shell was
loaded statically.
Another example of creep buckling in spherical shells, which applies to common civil
engineering structures, is the experimental investigation of the buckling and creep buckling
of concrete spherical domes subjected to external pressure, carried out by Vandepitte and
his co-workers at the State University Ghent, Belgium, in the seventies and early eighties
[19.166], [19.230]-[19.232]. As pointed out in Subsection 19.4.1, creep buckling has been
very prominent in concrete structures since concrete exhibits substantial creep already at
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

room temperatures. This motivated the extensive, 15-year-long, Ghent study on spherical
model domes made of microconcrete.
The geometry of the test specimens is shown in Figure 19.106. The spherical caps were
fairly large and thin, 1.9 min diameter and nominal (R/t) ~ 347. Their geometry was chosen
so that buckling occurred before the microconcrete would be crushed by excessive com-
pression when a liquid applied inward radial pressure to the domes. It took the Ghent re-
searchers nearly a year to develop sound microconcrete spherical shells of 7 mm thickness,
(R/t) ~ 350 and with a base diameter of 1.9 m. The casting method, which employed a
very rigid steel mold and involved a lengthy process of 21 days, is described in detail in
[19.230].
The edge of each model dome was cast into a groove in a steel ring (see detail A in the
figure) designed to withstand the thrust exerted by the radially loaded shell. More than half
of the domes tested were prestressed by means of four 0.5-in. (-13-mm) circumferential
steel strands, of the kind usually used for prestressing concrete structures (see Figure I 9.1 07),
and tensioned around the ring. The purpose of the prestressing was to compensate for the
elastic extension of the steel ring due to the thrust, as well as for the contraction of the
concrete caused by the liquid pressure, so that a membrane state of stress was achieved
approximately (in the rapid loading tests).

A- Thrust ring

Figure 19.106

Copyright Wiley
4
Ghent University creep buckling tests of microconcrete spherical domes subjected to
uniform external pressure-geometry of test specimens (from [19.166])
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1656 Thermal Buckling and Creep Buckling

Figure 19.107 Ghent University creep buckl ing tests of microconcrete spherical domes subjected tO
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

uniform radial pressure-typical spherical cap during the prestressing of its thrust ring
(counesy of Professor D. Yandepiuc)

The modulus of elasticity and the creep strains of microconcrete were measured on thin
bars, which were cast on the same day and from the same batch as the corresponding model
dome. The bars were loaded so that the uniaxial compressive stress equaled the biaxial
compressive stress in the cap as calcu lated wi th membrane theory. For most of the model
domes subjected to pennanent loadi ng, the changes of length of unloaded bars, otherwise
kept under the same conditions as the loaded bars and due to changes of temperature and
moisture, were also recorded.
The test program included 75 spherical caps, 34 of which were tested rapidly and 41
subjected to long-term loading, which eventually resulted in creep buckling failure. In the
rapid tests the external radial liquid pressure was increased unt il failure of the dome ensued,
after a few hours. In most of the rapid tests the domes had unprestressed rings, and in only
one-third of these tests were the rings prestressed. The long-term experiments were preceded
by creep buckling calculations with BOSOR 5 [16.149]. About half of the models were
tested with unprestressed rings and about half with prestressed rings.
All the Ghent model domes failed in the same manner for both rapid and long-tenn
loadi ng: an almost circular, sometimes slightly elliptic part of the shell wall , broken into a
number of sectors or pieces of another shape, was punched out by the pressure towards the
center of the sphere (see Figure 19.108 and [ 19.231]).
T his may be interpreted as follows: along the perimeter of one of the buckling dents. perhaps of
the dccpe.~t dimple, the brittle unreinforced concrete cracked at the convex s ide of the dome and
was crushed by excessive local bending at the concave side, while it was crushed along a number
of radiuses of the dent at the convex side, and the punched hole was a ftXtzen materialization of
the buckl ing dent. After each collapse the fragments of the punched out disk could be pieced
together. The disk was always nearly circular; the ratio of its greatest to its smallest dimension
seldom exceeded 1.15. Its size was also remarkably constant and the average of the mean of both
diameters was 360 mm, which is equal to 2.76 VRi.

For the rapid tests the BOSOR 4 [2.92] calculation pred icted oblong buckling waves, with
their greatest dimension in the meridional and their smallest dimension in the circumferential
direction. On the average, the actual size of the punched out disks was only 6 1 percent of
the size the BOSOR program predicted in the meridional di rection, and 139 percent of the
theoretical size in the circumferential direction. According to the calculations the buckling
waves should appear close to the edge of the dome. In most cases a hole was indeed punched
into the cap in the vicinity of its edge (sec for example Figures 19.108a, b).

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
Creep Buckling 1657

(a)

(b)

Fii(U"" 19.108 Ghent Unhersity creep buckling tests of microconcrete spherical domes subjected to
uniform radial pre~\ure-typical failure~ in the 'Phcrical caps (counesy or Professor
D. Vandepllle): (a) a specimen photographed while buckling failure occurred. (b) a
specimen after failure. Here one sees the circular hole punched out by the pressure.
The black bubble was fom1ed by the pla,tic film under some pressure resulting from
the water undcmcath

Theoret ically the deformation under long-term loading of a perfectly spherical dome is
axisymmetric, until an asymmetric pattern of dimples and bulges suddenly appears under
the bifurcation load. The Ghent test shells were not perfectly spherical. but their shape was
nearly axisymmetric. Nevenhclc~s. their deformation was already asymmetric soon after the
loadi ng was applied. In the tests of long duration the defonnation of the domes was mea.sured
after ccnain intervals, and in most cases. but not in all. a number of full waves could be
detected in the circumferential direction. The wave paucm was never regular and the am-
plitude of the radial displacement was different from wave to wave. In most creep tests the
curvature due tO the deformation was notably greater in one wave than everywhere else.
There a disk would eventuu lly be punched out. and in many cases it could be predicted
where buckling wou ld occur. The critical area became almost flat, and in I 3 cases curvature
of the shell even changed loc:tl ly. sometimes several weeks before failure.
Long-term radial pressure, which resulted in creep buckling. invariably caused the model
domes tO fail in the same way a~ rapid loading. namely by punching out a circular or slightly
elliptical disk, nonnally ncar the edge of the shell. The disks were somewhat larger than in
the case of rapidly increasing load. The mean diameter of the disks was. on the average,
about equal tO 2.94 VRi.
The typical failure of the microconcrete domes. by the punching out of a circular or
sli ght ly oval disk (see Figure 19. 108). observed in the Ghent tests is worth noting. Also of
interest is the applicabi lity of the Southwel l plot to the rapid tests of the microconcrete
--`,`,`````,`,````,,``,

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1658 Thermal Buckling and Creep Buckling

spherical domes (see [19.232]), which provides further corroboration of the applicability of
the Southwell method to shells, discussed in Section 9.6 of Chapter 9.
The carefully executed creep buckling tests of Vandepitte and his co-workers at Ghent
represent a good example of well-documented shell creep buckling experiments, in particular
of concrete shells, that should be consulted when one embarks on a program of creep buck-
ling tests.
To conclude this section, it should be pointed out that accurate creep buckling tests are
fraught with difficulties and require considerable skill. The data accumulated to date seem
to be in many cases insufficient for reliable design rules. Hence, though many theoretical
and numerical studies were conducted in the nineties (see for example [16.34], [16.35] and
[19.233]), there is a definite need for further creep buckling tests, in particular for plates and
shells.

References
19.1 Hoff, N.J., The Thermal Barrier-Structures, Transactions of the American Society of Mechanical
Engineers, 77, (5), 1955, 759-763.

--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---
19.2 Hoff, N.J., ed., High Temperature Effects in Aircraft Structures, Pergamon Press, London, New
York, 1958.
19.3 Hoff, N.J., High Temperature Effects in Aircraft Structures, Applied Mechanics Reviews, 8,
(11), 1955,453-456.
19.4 Bisplinghoff, R.L., Some Structural and Aeroelastic Considerations of High Speed Flight, Jour-
nal of the Aeronautical Sciences, 23, (4), 1956, 289-329.
19.5 Heldenfels, R.R., Frontiers of Flight Structural Design, in: Aeronautics and Astronautics, Pro-
ceedings of the Durand Centennial Conference held at Stanford University, August 5-8, 1959,
N.J. Hoff, and W.G. Vincenti, eds., Pergamon Press, New York, 1960, 29-51.
19.6 Heldenfels, R.R., Structural Prospects for Hypersonic Air Vehicles, in: Aerospace Proceedings
1966, Proceedings of the 5th International Congress of the Aeronautical Sciences (ICAS), Lon-
don, J. Bradbrooke, J. Bruce, and R.R. Dexter, eds. Macmillan, London, 561-583.
19.7 Becker, J.V., The X-15 Program in Retrospect, Raumfahrtforschung, March-April 1969, 45-53.
19.8 Heldenfels, R.R., Historical Perspectives on Thermostructural Research at the NACA Langley
Aeronautical Laboratory from 1948 to 1958, NASA TM 83266, Feb. 1982.
19.9 Thornton, E.A., Thermal Structures: Four Decades of Progress, Journal of Aircraft, 29, (3),
May-June 1992, 485-498.
19.10 Stillwell, W.H., X-15 Research Results with a Selected Bibliography, NASA SP-60, 1965.
19.11 Dotts, R.L., Curry, D.M. and Tillian, D.J., Orbiter Thermal Protection System, in: Space Shuttle
Technical Conference, NASA Conference Publication 2342, Part 2, 1985, 1062-1081.
19.12 Truitt, R.W., Fundamentals of Aerodynamic Heating, Ronald Press, New York, 1960.
19.13 Anderson, J.D., Jr., Hypersonic and High Temperature Gas Dynamics, McGraw-Hill, New York,
1989.
19.14 Hankey, W.L., Re-Entry Aerodynamics, AIAA Education Series, AIAA New York, 1988.
19.15 Anderson, J.D., Jr., Aerothermodynamks: A Tutorial Discussion, in: Thermal Structures and
Materials for High-Speed Flight, E. A. Thornton, ed., Progress in Astronautics and Aeronautics,
vol. 140, AIAA, Washington, D.C., 1992, 3-57.
19.16 Ried, R.C., Orbiter Entry Aerothermodynamics, in: Space Shuttle Technical Conference, NASA
Conference Publication 2342, Part 2, 1985, 1051-1061.
19.17 Martellucci, A. and Harris, T.B., Assessment of Key Aerothermal Issues for the Structural
Design of High Speed Vehicles, in: Thermal Structures and Materials for High-Speed Flight,
E.A. Thornton, ed., Progress in Astronautics and Aeronautics, vol. 140, AIAA, Washington,
D.C., 1992, 59-91.
19.18 Boley, B.A. and Weiner, J.H., Theory of Thermal Stresses, John Wiley & Sons, New York,
1960.
19.19 Johns, D.J., Thermal Stress Analysis, Pergamon Press, Oxford, 1965.
19.20 Parkus, H., Thermoelasticity, 2d ed., Springer, Vienna. 1976.
19.21 Nowacki, W., Thermoelasticity, Pergamon Press, Oxford, 1962.
19.22 Hetnarski, R.B., ed., Thermal Stresses I, North-Holland, Amsterdam, 1986.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1659

19.23 Hetnarski, R.B., ed., Thermal Stresses II, North-Holland, Amsterdam, 1987.
19.24 Hetnarski, R.B., ed., Thermal Stresses III, North-Holland, Amsterdam, 1989.
19.25 Ziegler, F. and Rammerstorfer, F.G., Thermoelastic Stability, in: Thermal Stresses III, Hetnarski,
R.B., ed., North-Holland, Amsterdam, 1989, chap. 2.
19.26 Duwez, P., Materials for High Temperature Aircraft Structures, in: High Temperature Effects in
Aircraft Structures, N.J. Hoff, ed., Pergamon Press, London, 1958, 58-79.
19.27 Bushong, R.M., Graphite as an Aerospace Material, Aerospace Engineering, 22, (1), Jan. 1963,
40-45.
19.28 Cooper, T.D. and Srp, 0.0., Refractory Metals and Their Protection, Aerospace Engineering,
22, (1), Jan. 1963, 46-55.
19.29 Ramke, W.O. and Latva, J.D., Refractory Ceramics and Intermetallic Compounds, Aerospace
Engineering, 22, (1), Jan. 1963, 76-84.
19.30 Scala, E., Materials Aspects of the Re-Entry Problem, in: High Temperature Structures and
Materials, Proceedings of 3rd Symposium on Naval Structural Mechanics, A.M. Freudenthal,
B.A. Boley, and H. Liebowitz, eds., Pergamon Press, Oxford, 1964, 249-257.
19.31 Lisagor, W.B., Advanced Metallics for High-Temperature Airframe Structures, in: Thermal
Structures and Materials for High-Speed Flight, E.A. Thornton, ed., Progress in Astronautics
and Aeronautics, vol. 140 AIAA, Washington, D.C., 1992, 161-179.
19.32 Brown, A.S., High-Temperature Materials Warm up for Debut, Aerospace America, 31, (3),
March 1993, 18-21,27.
19.33 Jackson, L.R., Dixon, S.C., Tenney, D.R., Carter, A.L. and Stephans, J.R., Hypersonic Structures
and Materials: A Progress Report, Aerospace America, 25, (10), Oct. 1987, 24-30.
19.34 Tenney, D.R., Lisagor, W.B. and Dixon, S.C., Materials and Structures for Hypersonic Vehicles,
in: !CAS Proceedings 1988, Proceedings of the 16th Congress of the International Council of
the Aeronautical Sciences, Jerusalem, August-September 1988, AIAA, Washington, D.C., 1988,
1, 398-415.
19.35 Williams, J.C., Materials Requirements for Aircraft Engines, in: Aerospace Thermal Structures
and Materials for a New Era, E.A. Thornton, ed., Progress in Astronautics and Aeronautics,
vol. 168, AIAA, Washington, D.C., 1955, 359-383.
19.36 Heldenfels, R.R. and Roberts, W.M., Experimental and Theoretical Determination of Thermal
Stresses in a Flat Plate, NACA TN 2769, 1952.
19.37 Gossard, M.L., Seide, P. and Roberts, W.M., Thermal Buckling of Plates, NACA TN 2771,
1952.
19.38 Hoff, N.J., Thermal Buckling of Supersonic Wing Panels, Journal of the Aeronautical Sciences,
23, (11), Nov. 1956, 1019-1029.
19.39 Hoff, N.J., Buckling at High Temperature, Journal of the Royal Aeronautical Society, 61, Nov.
1957, 756-775.
19.40 Klosner, J.M. and Forray, M.J., Buckling of Simply Supported Plates under Arbitrary Sym-
metrical Temperature Distributions, Journal of the Aeronautical Sciences, 25, (3), March 1958,
181-184.
19.41 Hoff, N.J., Buckling of Thin Cylindrical Shell under Hoop Stresses Varying in Axial Direction,
Journal of Applied Mechanics, 24, (3), 1957, 405-412.
19.42 Abir, D. and Nardo, S.V., Thermal Buckling of Circular Cylindrical Shells under Circumfer-
ential Temperature Gradients, Journal of the Aero/Space Sciences, 26, 1959, 803-808.
19.43 Thornton, E.A., Thermal Buckling of Plates and Shells, Applied Mechanics Review, 46, (10),
Oct. 1993, 485-506.
19.44 Vosteen, L.F. and Fuller, K.E., Behavior of a Cantilever Plate under Rapid-Heating Conditions,
NACA RM L55E20c, July 1955.
19.45 Hoff, N.J., Approximate Analysis of the Reduction in Torsional Rigidity and of Torsional
Buckling of Solid Wings under Thermal Stresses, Journal of the Aeronautical Sciences, 23, (6),
June 1956, 603, 604.
19.46 Budiansky, B. and Mayers, J., Influence of Aerodynamic Heating on the Effective Torsional
Stiffness of Thin Wings, Journal of the Aeronautical Sciences, 23, (12), Dec. 1956, 1081-
1093,1108.
19.47 Singer, J. and Hoff, N.J., Effect of the Change in Thermal Stresses Due to Large Deflections
on the Torsional Rigidity of Wings, Journal of the Aeronautical Sciences, 24, (4), April 1957,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

310.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1660 Thermal Buckling and Creep Buckling

19.48 Singer, J., Thermal Buckling of Solid Wings of Arbitrary Aspect Ratio, Journal of the Aero/
Space Sciences, 25, (9), Sept. 1958, 573-581.
19.49 Singer, J., The Effect of Amplitude on the Torsional Vibration of Solid Wings Subjected to
Aerodynamic Heating, Journal of the Aeronautical Sciences, 24, (8), August 1957, 620-622.
19.50 Heldenfels, R.R. and Vosteen, L.F., Approximate Analysis of Effects of Large Deflections and
Initial Twist on Torsional Stiffness of a Cantilever Plate Subjected to Thermal Stresses, NACA
Report 1361, 1958.
19.51 Miura, K., On Torsional Rigidity and Torsional Vibration of Aerodynamically Heated Wings
Having a Small Amount of Pretwist, Aeronautical Research Institute, University of Tokyo,
Report 354, May 1960.
19.52 Mansfield, E.H., The Influence of Aerodynamic Heating on the Flexural Rigidity of a Thin
Wing, Royal Aircraft Establishment Famborough, Structures Report 229, Sept. 1957.
19.53 Garrick, I.E., A Survey of Aerothermoelasticity, Aerospace Engineering, 22, (1), Jan. 1963,
140-147.
19.54 Garrick, I.E., Perspectives in Aeroelasticity, 5th Theodore von Kanruin Memorial Lecture, Pro-
ceedings 14th Israel Annual Conference on Aviation and Astronautics, March 1972, Israel
Journal of Technology, 10, (112), 1972, 1-22.
19.55 Dixon, S.C., Griffith, G.E. and Bohon, H.L., Experimental Investigation at Mach Number 3.0
of the Effects of Thermal Stress and Buckling on the Flutter of Four-Bay Aluminum Alloy
Panels with Length-Width Ratios of 10, NASA TN D-921, Oct. 1961.
19.56 Guy, L.D. and Bohon, H.L., Flutter of Aerodynamically Heated Aluminum-Alloy and Stainless-
Steel Panels with Length-Width Ratios of 10 at Mach No. of 3.0, NASA TN D-1353, July
1962.
19.57 Schaeffer, H.G. and Heard, W.L., Supersonic Flutter of a Thermally Stressed Flat Panel with
Uniform Edge Loads, NASA TN D-3077, October 1965.
19.58 Doggett, R.V., Jr., Ricketts, R.H., Noll, T.E. and Malone, J.B, NASP Aeroservo-thermoelasticity
Studies, NASA TM-104058, April 1991.
19.59 Heeg, J., Zeiler, T.A., Pototzky, A.S., Spain, C.V. and Engelund, W.C., Aero-thermoelastic
Analysis of a NASP Demonstrator Model, NASA TM-109007, October 1993.
19.60 Nydick, 1., Friedmann, P.P. and Zhong, X., Hypersonic Panel Flutter Studies on Curved Panels,
in: Technical Papers, 36th AIAA/ ASM.E/ ASCE/ AHS/ ASC Structures, Structural Dynamics
and Materials Conference, New Orleans, April 1995, Part 5, 2995-3011.
19.61 Xue, D.Y. and Mei, C., Finite Element Nonlinear Panel Flutter with Arbitrary Temperatures in
Supersonic Flow, AIAA Journal, 31, (1), Jan. 1993, 154-162.
19.62 Abbas, J.F., Ibrahim, R.A. and Gibson, R.F., Nonlinear Flutter of Orthotropic Composite Panel
under Aerodynamic Heating, AIAA Journal, 31, (8), August 1993, 1478-1488.
19.63 Zhou, R.C., Mei, C. and Huang, J.-K., Suppression of Nonlinear Panel Flutter at Supersonic
Speeds and Elevated Temperatures, AIAA Journal, 34, (2), Feb. 1996, 347-354.
19.64 Hoff, N.J., Rapid Creep in Structures, Journal of the Aeronautical Sciences, 22, (10), Oct. 1955,
661-672.
19.65 Hoff, N.J., A Survey of the Theories of Creep Buckling, in: Proceedings of the Third U.S.
National Congress of Applied Mechanics, ASME, New York, 1958, 29-49.
19.66 Hoff, N.J., Stress Distribution in the Presence of Creep, in: High Temperature Effects in Aircraft
Structures, N.J. Hoff, ed., Pergamon Press, London, New York, 1958, 248-266.
19.67 Hoff, N.J., ed., Creep in Structures, Proceedings of IUTAM Colloquium, Stanford University,
July 1960, Springer-Verlag, Berlin, 1962.
19.68 Rabotnov, Y.N., Creep Problems in Structural Members, North-Holland, Amsterdam, 1969.
19.69 Hull, J., ed., Creep in Structures 1970, Proceedings of IUTAM Symposium, Gothenburg, Swe-
den, August 1970, Springer-Verlag, Berlin, 1972.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

19.70 Kraus, H., Creep Analysis, John Wiley & Sons, New York, 1980.
19.71 Kelly, H.N. and Blosser, M.L., Active Cooling from the Sixties to NASP, in: Current Technology
for Thermal Protection Systems, NASA Conference Publication 3157, 1992, 189-249, and
NASA TM 109079, July 1994.
19.72 Shideler, J.L., Webb, G.L. and Pittman, C.M., Verification Tests of Durable Thermal Protection
System Concepts, Journal of Spacecraft and Rockets, 22, (6), Nov.-Dec. 1985, 598-604.
19.73 Curry, D.M., Thermal Protection Systems Manned Spacecraft Fight Experience, in: Current
Technology for Thermal Protection Systems, NASA Conference Publication 3157, 1992, 43-
61.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1661

19.74 Shideler, J.L., Predicted and Tested Performance of Durable TPS, in: Current Technology for
Thermal Protection Systems, NASA Conference Publication 3157, 1992, 97-129.
19.75 Camarda, C.J. and Glass, D.E., Thermostructural Applications of Heat Pipes for Cooling Lead-
ing Edges of High-Speed Aerospace Vehicles, in: Current Technology for Thermal Protection
Systems, NASA Conference Publication 3157, 1992, 291-318.
19.76 Goldstein, H., Reusable Thermal Protection System Development-A Prospective, in: Current
Technology for Thermal Protection Systems, NASA Conference Publication 3157, 1992, 1-18.
19.77 Lustig, S., A History of High-Temperature Testing at WPAFB, in: Structural Testing Technology
at High Temperature, Proceedings of SEM Conference, Dayton, Ohio, Nov. 1991, Society for
Experimental Mechanics, Bethel, Conn., 1991, 29-34.
19.78 Taylor, J., Experimental Methods in High Temperature Structural Research, in: High Temper-
ature Effects in Aircraft Structures, N.J. Hoff, ed., Pergamon Press, London, 1958, 313-322.
19.79 Horton, W.H., Laboratory Simulation of Kinetic Heating, Aircraft Engineering, 26, (303), May
1954, 138-144.
19.80 Taylor, J., Research Equipment for Investigation of Aircraft Structure at Elevated Temperatures,
Aircraft Engineering, 29, (342), August 1957, 228-232.
19.81 Horton, W.H., The Influence of Kinetic Heating on the Design and Testing of Aircraft Struc-
tures, in: Proceedings of the Conference on High-Speed Aeronautics, Polytechnic Institute of
Brooklyn, New York, 1955, 233-270.
19.82 Horton, W.H., Experimental Methods in Kinetic Heat Tests, Aircraft Engineering, 29, (342),
August 1957, 232-240.
19.83 Heldenfels, R.R., The Effect of Nonuniform Temperature Distributions on the Stresses and
Distortions of Stiffened-Shell Structures, NACA TN 2240, 1950.
19.84 Heldenfels, R.R., A Numerical Method for the Stress Analysis of Stiffened-Shell Structures
under Nonuniform Temperature Distributions, NACA Rep. 1043, 1951.
19.85 Heldenfels, R.R. and Rosecrans, R, Preliminary Results of Supersonic-Jet Tests of Simplified
Wing Structures, NACA RM L53E26a, July 1953.
19.86 Heldenfels, R.R., Rosecrans, R., and Griffith, G.E., Test of an Aerodynamically Heated Multiweb
Wing Structure (MW-1) in a Free-Jet at Mach Number Two, NACA RM L53E27, July 1953.
19.87 Heimerl, G.J. and Roberts. W.M., Determination of Plate Compressive Strengths at Elevated
Temperatures, NACA Report 960, 1950.
19.88 Combescure, A. and Brochard, J., Recent Advances on Thermal Buckling, New Results Ob-
tained at CEA, in: Buckling of Shell Structures, on Land, in the Sea and in the Air, J.F. Jullien,
ed., Elsevier Applied Science, London, New York, 1991, 381-390.
19.89 DeAngelis, V.M., A Historical Overview of High-Temperature Structural Testing at the NASA
Dryden Flight Research Facility, in: Structural Testing Technology at High Temperature, Pro-
ceedings of the Seventh High Temperature Measurement Conference, Dayton, Ohio, Nov. 1991,
Society for Experimental Mechanics, Bethel, Conn., 1991, 35-37.
19.90 Jenkins, J.M. and Sefic, W.J., Experimental Investigation of Thermal-Buckling Characteristics
of Flanged, Thin-Shell Leading Edges, NASA TN D-3243, Jan. 1966.
19.91 Fields, R.A. and Vano, A., Evaluation of an Infrared Heating Simulation of a Mach 4.63 Flight
on an X-15 Horizontal Stabilizer, NASA TN D-5403, Sept. 1969.
19.92 Fields, R.A., A Study of the Accuracy of a Flight-Heating Simulation and Its Effect on Load
Measurement, NASA TN D-5741, 1970.
19.93 Fields, R.A., Olinger, F.V. and Monaghan, R.C., Experimental Investigation of Mach 3 Cruise
Heating Simulations on a Representative Wing Structure for Flight-Loads Measurement, NASA
TN D-6749, March 1972.
19.94 Fields, R.A., Reardon, L.F. and Siegel, W.H., Loading Tests of a Wing Structure for a Hyper-
sonic Aircraft, NASA TP 1596, Jan. 1980.
19.95 Jenkins, J.M., A Comparison of Experimental and Calculated Thin-Shell Leading-Edge Buck-
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

ling Due to Thermal Stress, NASA TM 100416, 1988.


19.96 Thompson, R.C. and Richards, W.L., Thermal-Structural Panel Buckling Tests, NASA TM
104243, Dec. 1991.
19.97 DeAngelis, V.M. and Fields, R.A., Techniques for Hot Structures Testing, in: Thermal Structures
and Materials for High-Speed Flight, E.A. Thornton, ed., Progress in Astronautics and Aero-
nautics, vol. 140, AIAA, Washington, D.C., 1992, 255-277. also NASA TM 101727, Nov.
1990.
19.98 Fields, R.A., Flight Vehicle Thermal Testing with Infrared Lamps, NASA TM 4336, Jan. 1992.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1662 Thermal Buckling and Creep Buckling

19.99 Monaghan, R.C. and Fields, R.A., Experiments to Study Strain-Gage Load Calibrations on a
Wing Structure at Elevated Temperatures, NASA TN D-7390, August 1973.
19.100 Pope, A. and Goin, K.L., High-Speed Wind Tunnel Testing, John Wiley & Sons, New York,
1965.
19.101 Maus, J.R., Aerodynamic and Aerothermal Facilities 2, Part 1: Short-Duration, High Enthalpy
Facilities, in: von Karman Institute (VKI), Methodology of Hypersonic Testing, 1993.
19.102 Sikora, T.P. and Leger, K.B., Simulation of High Heat Flux Levels with Graphite Heating and
Arc Lamps, in: Structural Testing Technology at High Temperatures, Proceedings of the Seventh
High Temperature Measurement Conference, Dayton, Ohio, Nov. 1991, Society for Experi-
mental Mechanics, Bethel, Conn., 1991, 3-8.
19.103 Blosser, M.L., Boundary Conditions for Aerospace Thermal Structural Tests, in: Aerospace
Thermal Structures and Materials for a New Era, E. A. Thornton, ed., Progress in Astronautics
and Aeronautics, vol. 168, AIAA, Washington, D.C., 1995, 119-144.
19.104 USAF WLFDD, Flier on Aerospace Structures Test Facilities, Wright-Patterson Air Force Base,
Dayton, Ohio 1996.
19.105 Davies, J. and Simpson, P., Induction Heating Handbook, McGraw-Hill, London, 1979.
19.106 Brochard, J., Combescure, A., Locatelli, T. and Tomassian, R., Experimental and Numerical
Thermal Buckling Studies on Cylinders, presented at SMIRT 10, International Conference on
Structural Mechanics in Reactor Technology, Anaheim, Calif., August 1989; also Report CEA-
CONF-10013, 1989.
19.107 Harpur, N.F., Concorde Structural Development, Journal of Aircraft, 5, (2), 1968, 176-183.
19.108 Boggs, B.C., The History of Static Test and Air Force Structures Testing, Wright-Patterson Air
Force Base, AFFDL-TR-70-3071, Dayton, Ohio, June 1979.
19.109 Sefic, W.J., NASA Dryden Flight Loads Research Facilities, NASA TM 81368, Dec. 1981.
19.110 Hanson, H.A. and Casey, J.J., High-Temperature Test Technology, FluiDyne Engineering Cor-
poration, Minneapolis, Minn., AFWAL-TR-86-3105, Feb. 1987.
19.111 Richards, W.L., A New Correction Technique for Strain-Gage Measurements Acquired in Tran-
sient-Temperature Environments, NASA Technical Paper 3593, March 1996.
19.112 Hussong, F. and Pappas, J., Instrumentation of the Lightly Loaded Splice Subcomponent, in:
Proceedings of the Structural Testing Technology at High Temperature Conference, Ojai, Calif.,
Nov. 8-10, 1993.
19.113 Heldenfels, R.R., Some Design Implications of the Effects of Aerodynamic Heating, NACA
RM L55F22, July 1955.
19.114 Mendelson, A. and Hirschberg, M., Aaalysis of Elastic Thermal Stresses in Thin Plates with
Spanwise and Chordwise Variations of Temperature and Thickness, NACA TN 3778, 1956.
19.115 Singer, J., An1iker, M. and Lederman, S., Thermal Stresses and Thermal Buckling, Chapter 1,
U.S. Air Force, Wright Aeronautical Development Center, WADC TR 57-69, 1957.
19.116 Przemieniecki, J.S., Thermal Stresses in Rectangular Plates, Aeronautical Quarterly, 10, 1959,
65-78.
19.117 Rama Rao, K. and Johns, D.J., Some Thermal Stress Analyses for Rectangular Plates, Journal
of the Aeronautical Society of India, 1.3, (4), 1961, 99-104.
19.118 Hoff, N.J., The Structural Effects of Aerodynamic Heating, in: Proceedings of the 3rd General
Assembly of AGARD of NATO, London, Sept. 7-10, 1953.
19.119 Zender, G.W. and Pride, R.A., The Combinations of Thermal and Load Stresses for the Onset
of Permanent Buckling of Plates, NACA TN 4053, June 1957.
19.120 Miura, K., Thermal Buckling of Rectangular Plates, Aeronautical Research Institute, University
of Tokyo, Report 353, 1960.
19.121 Rammerstorfer, F.G., Increase of the First Natural Frequency and Buckling Load of Plates by
Optimal Fields of Initial Stresses, Acta Mechanica, 27, 1977, 217-238.
19.122 Tauchert, T.R., Thermally Induced Flexure, Buckling, and Vibration of Plates, Applied Me-
chanics Review, 44, 1991, 347-360.
19.123 Thornton, E.A., Coyle, M.F. and McLeod, R.N., Experimental Study of Plate Buckling Induced
by Spatial Temperature Gradients, Journal of Thermal Stresses, 17, 1994, 191-212.
19.124 Kotanchik, J.N., Johnson, A.E., Jr. and Ross, R.D., Rapid Radiant-Heating Tests of Multiweb
Beams, NACA TN 3474, Sept. 1955.
19.125 Teare, W.P. and Fields, R.A., Buckling Analysis and Test Correlation of High Temperature
Structural Panels, in: Thermal Structures and Materials for High Speed Flight, E.A. Thornton,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1663

ed., Progress in Astronautics and Aeronautics, vol. 140, AIAA, Washington, D.C., 1992, 337-
352.
19.126 Polesky, S.P. and Shideler, J.L., Inverse Analysis for Structural Boundary Condition Character-
ization of a Panel Test Fixture, in: Aerospace Thennal Structures and Materials for a New Era,
E.A. Thornton, ed., Progress in Astronautics and Aeronautics, vol. 168, AIAA, Washington,
D.C., 1995, 145-152.
19.127 Johns, D.J., Houghton, D.S. and Webber, J.P.H., Buckling Due to Thermal Stress of Cylindrical
Shells Subjected to Axial Temperature Distributions, The College of Aeronautics, Cranfield,
England, Report 147, May 1961.
19.128 Anderson, M.S., Combinations of Temperature and Axial Compression Required for Buckling
of a Ring-Stiffened Cylinder, NASA TN D-1224, 1962.
19.129 Anderson, M.S., Thermal Buckling of Cylinders, in: Collected Papers on Instability of Shell
Structures, NASA TN D-1510, Dec. 1962, 255-266.
19.130 Anderson, M.S. and Card, M.F., Buckling of Ring-Stiffened Cylinders under a Pure Bending
Moment and a Nonuniform Temperature Distribution, NASA TN D-1513, 1962.
19.131 Libai, A. and Durban, D., Buckling of Cylindrical Shells Subjected to Nonuniform Axial Loads,
ASME Journal of Applied Mechanics, 44, 1977, 714-720.
19.132 Teng, G.-G. and Rotter, J.M., A Study of Buckling in Column-Supported Cylinders, in: Contact
Loading and Local Effects in Thin- Walled Plated and Shell Structures, V. Krupka, and M.
Drdacky, eds., Springer-Verlag, Berlin, 1991, 52-61.
19.133 Forray, M.J. and Newman, M., The Post-Buckling Analysis of Heated Rectangular Plates, Jour-
nal of the Aerospace Sciences, 29, (10), Oct. 1962, 1262.
19.134 Newman, M. and Forray, M.J., Axisymmetric Large Deflections of Circular Plates Subjected
to Thermal and Mechanical Loads, Journal of the Aerospace Sciences, 29, (9), Sept. 1962,
1060-1066.
19.135 Webber, J.P.H. and Houghton, D.S., Thermal Buckling of a Free Circular Plate, College of
Aeronautics, Cranfield, England, Note 105, August 1960.
19.136 Pride, R.A., Hall, J.B. and Anderson, M.S., Effects of Rapid Heating on Strength of Airframe
Components, NACA TN 4051, June 1957.
19.137 Richards, W.L. and Thompson, R.C., Titanium Honeycomb Panel Testing, in: Structural Testing
Technology at High Temperature, Proceedings of Society of Experimental Mechanics Confer-
ence, Dayton, Ohio, Nov. 4-6, 1991, 116-132.
19.138 Hudson, L.D. and Thompson, R.C., Single-Strain-Gage Force/Stiffness Buckling Prediction
Techniques on a Hat-Stiffened Panel, NASA TM-101733, 1991.
19.139 Shideler, J.L., Fields, R.A., Reardon, L.F. and Gong, L., Thermal and Structural Tests of Rene
41 Honeycomb Integral-Tank Concept for Future Space Transportation Systems, NASA TP
3145, May 1992.
19.140 Hill, D.W., Buckling of a Thin Circular Cylindrical Shell Heated along an Axial Strip, Stanford
University, Calif., Report SUDAER 88, Dec. 1959.
19.141 Ross, B., Mayers, J. and Jaworski, A., Buckling of Thin Cylindrical Shells Heated along an
Axial Strip, Experimental Mechanics, 5, (8), August 1965, 247-256.
19.142 Ross, B., Hoff, N.J. and Horton, W.H., The Buckling Behavior of Uniformly Heated Thin
Circular Cylindrical Shells, Experimental Mechanics, 6, 1966, 529-537.
19.143 Smith, S., An Experimental Investigation of the Thermal Buckling of Conical Shells, Thesis,
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Department of Aeronautics and Astronautics, Stanford University, Calif., May 1964.


19.144 Holmes, A., Measurement of Thermal Stresses in Ring-Stiffened Cylinders, Lockheed Missiles
and Space Co., Palo Alto, Calif., Report LMSC/Yl-69-66-1, Dec. 1966.
19.145 Holmes, A., An Experimental Investigation of Cylinder Buckling under Axial Load and Thermal
Stress Due to Asymmetrical Heating, Lockheed Missiles and Space Co., Palo Alto, Calif.,
Report LMSC/Y1-69-66-2, Dec. 1966.
19.146 Bushnell, D. and Smith, S., Stress and Buckling of Nonuniformly Heated Cylindrical and
Conical Shells, AIAA Journal, 9, (12), Dec. 1971, pp. 2314-2321.
19.147 Frum, Y. and Baruch, M., Experimental Analysis of the Heat Transfer in a Cylindrical Shell
Heated along a Generator, Experimental Mechanics, 15, (7), July 1975, 265-270.
19.148 Frum, Y. and Baruch, M., Buckling of Cylindrical Shells Heated along Two Opposite Generators
Combined with Axial Compression, Experimental Mechanics, 16, (4), April 1976, 133-139.
19.149 Ari-Gur, J., Baruch, M. and Singer, J., Buckling of Cylindrical Shells under Combined Axial
Preload, Non-Uniform Heating and Torque, TAE Report 315, Department of Aeronautical En-
gineering, Technion, Haifa, Israel, Dec. 1977.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1664 Thermal Buckling and Creep Buckling

19.150 Librescu, L. and Lin, W., Thermomechanical Postbuckling of Plates and Shells Incorporating
Non-Classical Effects, in: Thermal Stresses IV, R.B. Hetnarsky, ed., Elsevier, Amsterdam, New
York, 1996, 379-452.
19.151 Weller, T. and Patlashenko, I., PostbudJing of Infinite Length Cylindrical Panels under Com-
bined Thermal and Pressure Loading, International Journal of Solids and Structures, 30, (12),
1993, 1649-1662.
19.152 Johnson, T.F. and Card, M.F., Effects of Stiffening and Mechanical Load on Thermal Buckling
of Stiffened Cylindrical Shells, AIAA Paper 95-1317, in: Technical Papers of AIAA/ ASME/
ASCE/ AHS I ASC 36th Structures, Stmctural Dynamics and Materials Conference, New Or-
leans, La., April 10-13, 1995, Part 2, 1380-1388.
19.153 Thornton, E.A., Experimental Methods for High-Temperature Aerospace Structures, in: Thermal
Stresses IV, R.B. Hetnarski, ed., Elsevier Science B.V., Amsterdam, 1996, 1-89.
19.154 Ambur, D.R., Rouse, M., Starnes, J.H., Jr. and Shuart, M., Facilities for Combined Loads
Testing of Aircraft Structures to Satisfy Structural Technology Development Requirements, in:
Proceedings of the 5th Annual Advanced Composites Technology Conference, Seattle, Wash.,
August 22-26, 1994.
19.155 Ambur, D.R., Cerro, J.A. and Dickson, J.N., D-Box Fixture for Testing Stiffened Panels in
Compression and Pressure, Journal of Aircraft, 32, (6), Nov.-Dec. 1995, 1382-1389.
19.156 Breivik, N.L., Hyer, M.W. and Starnes, J.H., Jr., Thermal and Mechanical Buckling and Post-
buckling Response of Selected Curved Composite Panels, presented at the AIAA/ ASME/
ASCE/ AHS/ ASC 38th Structures, Structural Dynamics, and Materials Conference, Kissimmee,
Fla., April 7-10, 1997, AIAA Paper No. 97-1247.
19.157 Ambur, D.R., Sikora, J., Maguire, J.F. and Winn, P.M., Development of a Pressure Box to
Evaluate Reusable-Launch-Vehicle Cryogenic-Tank Panels, presented at the AIAA/ ASME/
ASCE/ AHS/ ASC 37th Structures, Structural Dynamics, and Materials Conference, Salt Lake
City, Utah, April 15-17, 1996, AIAA Paper No. 96-1640.
19.158 Ambur, D.R. and Rouse, M., Design and Evaluation of Composite Fuselage Panels Subjected
to Combined Loading Conditions, presented at the AIAA/ ASME/ ASCE/ AHS/ ASC 38th
Structures, Structural Dynamics, and Materials Conference, Kissimmee, Fla., April 7-10, 1997,
AIAA Paper No. 97-1303.
19.159 Hoff, N.J., Necking and Rupture under Tensile Loads, Journal of Applied Mechanics, 20, (1),
March 1953, 105-108.
19.160 Kempner, J. and Patel, S.A., Creep Buckling of Columns, NACA TN 3138, Jan. 1954; also
P.I.B.A.L. Report 205, Polytechnic Institute of Brooklyn, Nov. 1952.
19.161 Odquist, F.K.G., Recent Advances in Theories of Creep of Engineering Materials, Applied
Mechanics Review, 7, (12), Dec. 1954, 517-519.
19.162 Libove, C., Creep Buckling of Columas, Journal of the Aeronautical Sciences, 19, (7), July
1952, 459-467.
19.163 Hoff, N.J., Creep Buckling, Aeronautical Quarterly, 7, (1), Feb. 1956, 1-20.
19.164 Fraeijs de Veubeke, B., Creep Buckling in: High Temperature Effects in Aircraft Structures,
Pergamon Press, London, 1958, 267-287.
19.165 Finnie, I., Reflections on the Past and Future of Creep, in: Creep in Structures, Proceedings of
IUTAM Symposium, Gothenburg, Sweden, August 1970, J. Hult, ed., Springer-Verlag, Berlin,
1972.
19.166 Vandepitte, D. and Lagae, G., Buckling of Spherical Domes Made of Microconcrete and Creep
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

Buckling of Such Domes under Long-Term Loading, Inelastic Behaviour of Plates and Shells,
Proceedings of IUTAM Symposium, Rio de Janeiro, 1985, L. Bevilacqua, R. Feij6o, and R.
Valid, eds., Springer-Verlag, Berlin, Heidelberg, 1986, 291-311.
19.167 Jones, N. and Xirouchakis, P.C., The Creep Buckling of Shells, in: Creep in Structures, Pro-
ceedings of the 3rd IUTAM Symposium, Leicester, U.K., Sept. 1980, A.R.S. Ponter, and D.R.
Hayhurst, eds., Springer-Verlag, Berlin, Heidelberg, 1981, 308-330.
19.168 Andrade, E.N. da C., On the Viscous Flows of Metals and Allied Phenomena, Proceedings of
the Royal Society, Series A, 84, (A567), June 1910, 1-12.
19.169 Norton, F.H., The Creep of Steel at High Temperatures, McGraw-Hill, New York, 1929.
19.170 Bailey, R.W., The Utilisation of Creep Test Data in Engineering Design, Proceedings of the
Institution of Mechanical Engineers, Bl, 1935, 131-149.
19.171 Dorn, J.E. and Tietz, T.E., Creep and Stress-Rupture Investigations on Some Aluminum Alloy
Sheet Metals, Proceedings of the American Society for Testing Materials (ASTM), 49, 1949,
815-831.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1665

19.172 Hult, J., Creep in Engineering Structures, Blaisdell, Waltham, Mass., 1966.
19.173 Boyle, J.T. and Spence, J., Stress Analysis for Creep, Butterworths, London, Boston, 1983.
19.174 Ponter, A.R.S. and Hayhurst, D.R., eds., Creep in Structures, Proceedings of the 3rd IUTAM
Symposium, Leicester, U.K., Sept. 1980, Springer-Verlag, Berlin, Heidelberg, 1981.
19.175 Kempner, J. and Hoff, N.J., Behaviour of a Linear Visco-Elastic Column, Appendix II to Struc-
tural Problems of Future Aircraft, by N.J. Hoff, in: Proceedings of the 3rd Anglo-American
Aeronautical Conference, Brighton, England, Sept. 1951, Royal Aeronautical Society, London,
1952, 110-113.
19.176 Libove, C., Creep-Buckling Analysis of Rectangular-Section Columns, NACA TN 2956, June
1953.
19.177 Kempner, J., Creep Bending and Buckling of Linearly Viscoelastic Columns, NACA TN 3136,
Jan. 1954.
19.178 Kempner, J., Creep Bending and Buckling of Nonlinearly Viscoelastic Columns, NACA TN
3137, Jan. 1954.
19.179 Hoff, N.J., A Survey of the Theories of Creep Buckling, in: Proceedings of the Third U.S.
National Congress of Applied Mechanics, ASME, New York, 1958, 29-49.
19.180 Chapman, J.C., Erickson, B. and Hoff, N.J., A Theoretical and Experimental Investigation of
Creep Buckling, International Journal of Mechanical Sciences, 1, (2/3), 1960, 145-174.
19.181 Hoff, N.J. and Levi, I.M., Short Cuts in Creep Buckling Analysis, International Journal of
Solids and Structures, 8, 1972, 1103-1114.
19.182 Hoff, N.J., Creep Buckling of Plates and Shells, in: Applied Mechanics, Proceedings of 13th
International Congress of Theoretical and Applied Mechanics, Moscow, August 1972, Springer-
Verlag, Berlin, 1973, 435-451.
19.183 Hoff, N.J., Jahsman, W.E. and Nachbar, W., A Study of Creep Collapse of a Long Circular
Cylindrical Shell under Uniform External Pressure, Journal of the Aerospace Sciences, 26, Oct.
1959, 663-669.
19.184 Gerard, G., Theory of Creep Buckling of Perfect Plates and Shells, Journal of the Aerospace
Sciences, 29, Sept. 1962, 1087-1090.
19.185 Hult, J., Creep Buckling and Instability, in: Creep of Engineering Materials and Structures, G.
Bernasconi and G. Piatti, eds., Applied Science Publishers, London, 1978, 133-145.
19.186 Samuelson, L.A., Creep Collapse of Structures, in: ASME PVP-Vol. 84, 1984, 137-151.
19.187 Minahen, T.M. and Knauss, W.G., Creep Buckling of Viscoelastic Structures, International
Journal of Solids and Structures, 30, (8), 1993, 1075-1092.
19.188 Mathauser, E.E. and Brooks, W.A., An Investigation of the Creep Lifetime of 7075-T6 Alu-
minium-Alloy Columns, NACA TN 3204, July 1954.
19.189 Patel, S.A., Bloom, M., Erickson, B., Chwick, A. and Hoff, N.J., Development of Equipment
and of Experimental Techniques for Column Creep Tests, NACA TN 3493, Sept. 1955; also
PIBAL Report 239, Dec. 1953.
19.190 Carlson, R.L. and Breindel, W.W., On the Mechanics of Column Creep, in: Creep in Structures,
Proceedings of IUTAM Colloquium Stanford, Calif., 1960, N.J. Hoff, ed., Springer-Verlag,
Berlin, 1962, 272-290.
19.191 Lempriere, B.M., Comparison of Ranges of Applicability of Predictions of Creep Buckling
Time, in: Creep in Structures, Proceedings of IUTAM Colloquium Stanford, Calif., 1960, N.J.
Hoff, ed., Springer-Verlag, Berlin, 1962, 291-306.
19.192 Madsen, W.A. and Lempriere, B.M., Creep Buckling Experiments, Stanford University, De-
partment of Aeronautics and Astronautics, SUDAER 130, July 1962.
19.193 Samuelson, A., Creep Deformation and Buckling of a Column with an Arbitrary Cross Section,
Aeronautical Research Institute of Sweden (FFA), Report 107, 1967.
19.194 Mathauser, E.E., Investigation of Static Strength and Creep Behavior of an Aluminum-Alloy
Multiweb Box Beam at Elevated Temperatures, NACA TN 3310, Nov. 1954.
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

19.195 Mathauser, E.E. and Deveikis, W.O., Investigation of the Compressive Strength and Creep
Lifetime of 2024-T Aluminum Alloy Skin-Stringer Panels at Elevated Temperatures, NACA
TN 3647, May 1956.
19.196 Mathauser, E.E. and Deveikis, W.O., Investigation of the Compressive Strength and Creep
Lifetime of 2024-T3 Aluminum-Alloy Plates at Elevated Temperatures, NACA Report 1308,
1957.
19.197 Heimerl, G.J. and Farquhar, J., Compressive and Tensile Creep of 7075-T6 and 2024-T3 Alu-
minum-Alloy Sheet, NASA TN D-160, December 1959.
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
1666 Thermal Buckling and Creep Buckling

19.198 Rabotnov, G.N. and Shesterikov, S.A., Creep Stability of Columns and Plates, Journal of Me-
chanics and Physics of Solids, 6, 1957, 27-34.
19.199 Sanders, J.L., Jr., McComb, H.G., Jr. and Schlechte, F.R., A Variational Theorem for Creep
with Applications to Plates and Columns, NACA Report 1342, 1958.
19.200 McComb, H.G., Jr., Analysis of the Creep Behavior of a Square Plate Loaded in Edge Com-
pression, NACA TN 4398, September 1958.
19.201 Gerard, G. and Gilbert, A.C., A Critical Strain Approach to Creep Buckling of Plates and
Shells, Journal of the Aero/Space Sciences, 25, (7), July 1958, 429-434, 458.
19.202 Jahsmann, W.E. and Field, F.A., Creep Instability of Initially Flat Plates, in: Proceedings of the
Joint International Conference on Creep, Institution of Mechanical Engineers, London, 1963,
2-43.
19.203 Hoff, N.J., Creep Buckling of Rectangular Plates under Uniaxial Compression, in: Engineering
Plasticity, Sir John Baker Anniversary Volume, J. Heyman, and F.A. Leckie, eds., Cambridge
University Press, 1968, 257-276.
19.204 Levi, I.M. and Hoff, N.J., Creep Budling and Postbuckling of Plates, Stanford University,
Department of Aeronautics and Astronautics, SUDAAR Report 355, August 1968.
19.205 Levi, I.M. and Hoff, N.J., The Postcritica1 Behavior of Compressed Plates That Creep,lngenieur
Archiv, 38, (4/5), 1969, 329-342.
19.206 Hoff, N.J., Berke, L., Honikman, T.C. and Levi, I.M., Creep-Buckling of Flat Rectangular Plates
when the Creep Exponent Ranges from 3 to 7, in: Advances in Creep Design, the A.E. Johnson
Memorial Volume, A.E. Smith, and A.M. Nicolson, eds., Applied Sciences, London, 1971,
421-441.
19.207 Hoff, N.J., Theory and Experiment in the Creep Buckling of Plates and Shells, in: Buckling of
Structures, Proceedings of IUTAM Symposium, Harvard University, Cambridge, Mass., June
--`,`,`````,`,````,,``,``,,`-`-`,,`,,`,`,,`---

1974, B. Budiansky, ed., Springer-Verlag, Berlin 1976, 67-77.


19.208 Hoff, N.J. and Benoit, M., An Experimental Check of the Theory of the Creep Buckling of
Plates, in: Mechanics of Deformable Solids and Structures, the Y.N. Rabotnov Anniversary
Volume, U.S.S.R. Academy of Sciences, Moscow Mashinostroenie, 1975,491-494 (in Russian).
19.209 Hoff, N.J., Jahsman, W.E. and Nachbar, W., A Study of Creep Collapse of a Long Circular
Cylindrical Shell under Uniform External Pressure, Journal of the Aero/Space Sciences, 26,
Oct. 1959, 371-377.
19.210 Samuelson, L.A., A Theoretical Investigation of Creep Deformation and Buckling of a Circular
Cylindrical Shell under Axial Compression and Internal Pressure, The Aeronautical Research
Institute of Sweden (FFA) Report 100, 1964.
19.211 Pittner, E.V. and Hoff, N.J., Creep Buckling of Simply Supported Moderately Thin Circular
Cylindrical Shells, Acta Mechanica, 8, 1969, 116-125.
19.212 Samuelson, L.A., Creep Buckling of a Circular Cylindrical Shell, AIAA Journal, 1, (1), Jan.
1969, 42-49.
19.213 Levi, I.M. and Hoff, N., Interaction between Axisymmetric and Nonsymmetric Creep Buckling
of Circular Cylindrical Shells in Axial Compression, in: Creep in Structures, Proceedings of
IUTAM Symposium, Gothenburg, Sweden, August 1970, J. Hult, ed., Springer-Verlag, Berlin,
1972, 403-423.
19.214 Samuelson, L.A., Creep Buckling of Imperfect Circular Cylindrical Shells under Non-Uniform
External Loads, Aeronautical Institute of Sweden (FFA) Report HF-1327:4, August 1970.
19.215 Hoff, N.J., On the Transition from Axisymmetric to Multilobed Creep Buckling, in: Contri-
butions to the Theory of Aircraft Structures, A. van der Neut, ed., Delft University Press, 1972,
297-303.
19.216 Obrecht, H., Creep Buckling and Postbuckling of Circular Cylindrical Shells under Axial Com-
pression, International Journal of Solids and Structures, 13, 1977, 337-355.
19.217 Erickson, B., French, F.W., Patel, S.A., Hoff, N.J. and Kempner, J., Creep Bending and Buckling
of Thin Circular Cylindrical Shells, NASA TN D-429, June 1960.
19.218 Samuelson, L.A., An Experimental Investigation of Creep Buckling of Circular Cylindrical
Shells Subject to Axial Compression, Aeronautical Research Institute of Sweden (FFA) Report
98, 1964.
19.219 Samuelson, L.A, Creep Deformation and Buckling of a Circular Cylindrical Shell under Axial
Compression, Aeronautical Research Institute of Sweden (FFA) Report 108, 1967.
19.220 Kuznetsov, A.P. and Iungerman, N.M., Experimental Investigation of Shell Stability under Creep
Conditions, Prikl. Mekh. I Tekh. Fiz., (4), 1965, 128-131 (in Russian).
Copyright Wiley
Provided by IHS Markit under license with WILEY Licensee=McDermott Inc - India/8215328006, User=G, Boopathi
No reproduction or networking permitted without license from IHS Not for Resale, 02/13/2019 01:32:58 MST
References 1667

19.221 Diamant, E.S., Axisymmetric Creep in Cylindrical Shells, AIAA Journal, 5, Oct. 1967, 1870-
1876.
19.222 Samuelson, L.A., Experimental Investigation of Creep Buckling of Circular Cylindrical Shells
under Axial Compression and Bending, ASME Journal of Engineering for Industry, 90, 1968,
589-595.
19.223 Papimo, R. and Goldman, R., Experimental Creep Buckling of Aluminum Cylinders in Axial
Compression, Experimental Mechanics, 9, August 1969, 356-365.
19.224 Carlson, R., Schneider, B. and Berke, L., An Experimental Study of the Creep Buckling of
Circular Cylindrical Shells under an Axially Applied Compression, Stanford University, De-
partment of Aeronautics and Astronautics, SUDAER Report 248, 1965.
19.225 Benoit, M., Sendelbeck, R.L. and Hoff, N.J., An Experimental Study of the Creep Buckling of
Moderately Thin-Walled Circular Cylindrical Shells Subjected to Axial Compression, Stanford
University, Department of Aeronautics and Astronautics, SUDAAR Report 415, Sept. 1970.
19.226 Rikards, R.B. and Teters, G.A., Nonsymmetric Creep Buckling of Cylindrical Shells under Axial
Compression and External Pressure, in: Buckling of Structures, Proceedings of IUTAM Sym-
posium, Harvard University, Cambridge, Mass., June 1974, B. Budiansky, ed., Springer-Ve

You might also like