You are on page 1of 629

Buckling Experiments:

Experimental Methods in
Buckling of Thin-Walled
Structures
Basic Concepts, Columns, Beams
and Plates Volume 1

Buckling Experiments: Experimental Methods in Buckling of Thin-Walled Structures: Basic Concepts, Columns, Beams and
Plates – Volume 1. J. Singer, J. Arbocz and T. Weller Copyright © 1998 John Wiley & Sons, Inc.
Buckling Experiments:
Experimental Methods in
Buckling of Thin-Walled
Structures
Basic Concepts, Columns, Beams
and Plates Volume 1

J. Singer
Technion-Israel Institute of Technology, Israel

J. Arbocz
Technical University Delft, The Netherlands

T. Weller
Technion-Israel Institute of Technology, Israel

JOHN WILEY & SONS, INC.


1
This book is printed on acid-free paper. 
Copyright  1998 by John Wiley & Sons, Inc., New York. All rights reserved.
Published by John Wiley & Sons, Inc.
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.

Library of Congress Cataloging-in-Publication Data:

Singer, J.
Buckling experiments: experimental methods in buckling of thin
-walled structures/J. Singer, J. Arbocz, T. Weller.
p. cm.
Includes bibliographical references and index.
Contents: v. 1. Basic concepts, columns, beams, and plates.
ISBN 0-471-95661-9 (v. 1 : cloth)
1. Buckling (Mechanics) Experiments. I. Arbocz, Johann.
II. Weller, T. III. Title.
TA410.S57 1997
624.10 76 dc21 96-52326

Printed in the United States of America.


10 9 8 7 6 5 4 3 2
Contents

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


Preface xv
Abbreviated Contents of Vol. 2: Shells, Built-up Structures and xi
Additional Topics

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
References 11

2 Concepts of Elastic Stability 15


2.1 Physical Concepts Types of Observed Behavior and
Their Meaning 15
2.1.1 Instability of Columns 16
2.1.2 Instability of Plates 18
2.1.3 Instability of Columns with Compound Cross-Sections 21
2.1.4 Effect of Modal Coupling 25
2.1.5 Buckling of Frames 28
2.1.6 Lateral Buckling of Beams 32
2.1.7 Instability due to Patch Loading 36
2.1.8 Buckling of Beam-Columns 39
2.1.9 Buckling of Rings and Arches 41
2.1.10 Buckling of Shallow Arches 45
2.1.11 Buckling of Circular Cylindrical Shells 50
a. Axial Compression 53
b. Combined External Pressure and Axial Compression 57
c. Combined Torsion and Axial Compression 59
d. Combined Bending and Axial Compression 63
vi Contents

2.1.12 Buckling of Shells of Revolution 66


a. Externally Pressurized Shallow Spherical Caps 69
b. Toroidal Shell Segments under External Pressure
p D pe  72
c. Toroidal Segments under Axial Tension 77
d. Domed (torispherical) End-Closures under Internal
Pressure 78
2.1.13 Influence of Nonlinear Effects 80
a. Axially Compressed Cylindrical Shells 81
b. Bending of Cylinders Ovalization of the Cross-Section 84
c. Plastic Buckling 88
2.2 Mathematical Models for Perfect Structures 94
2.2.1 Static Versus Kinematic Approach 95
2.2.2 Approximate Solutions of Bifurcation Problems 101
a. The Rayleigh Ritz Method 102
b. Galerkin’s Method 106
2.2.3 Computational Tools for Bifurcation Problems 110
a. The BOSOR-4 Branched Complex Shell of
Revolution Code 111
b. Finite Element Formulation of Bifurcation Problems 121
References 124

3 Postbuckling Behavior of Structures 131


3.1 Introduction 131
3.2 Asymptotic Imperfection Sensitivity Analysis 134
3.2.1 Initial Postbuckling Behavior of Columns 136
3.2.2 Initial Postbuckling Behavior of Plates 139
3.2.3 Initial Postbuckling Behavior of Shells 143
3.2.4 Experimental Verification 148
3.3 Direct Solutions of the Nonlinear Stability Problem 154
3.3.1 Elastic Postbuckling Behavior of Columns 154
3.3.2 Plastic Postbuckling Behavior of Columns 156
3.3.3 Postbuckling Behavior of Plates 160
a. Perfect Plates 161
b. Imperfect Plates 166
3.3.4 Postbuckling Behavior of Circular Cylindrical Shells 167
a. Perfect Shells 167
b. Imperfect Shells 170
3.3.5 Concluding Remarks 175
References 177

4 Elements of a Simple Buckling Test a Column Under


Axial Compression 181
4.1 Columns and Imperfections 181
4.2 Von Kármán’s Experiments 182
4.3 The Basic Elements of a Buckling Experiment 185
Contents vii

4.4 Demonstration Experiments 187


4.4.1 University College London Initial Postbuckling Experiments 187
4.4.2 Mechanical Models 189
4.5 Southwell’s Method 194
4.5.1 Derivation of Southwell Plot for a Column 194
4.5.2 Application to von Kármán’s Columns 195
4.6 Application of the Southwell Method to Columns, Beam
Columns and Frames 197
4.6.1 Lundquist Plot 197
4.6.2 Donnell’s Applications of the Southwell Plot 198
4.6.3 Applications to Frames and Lateral Buckling of Beams 203
4.6.4 Southwell’s Method as a Nondestructive Test Method 206
4.7 Remarks on the Applicability of the Southwell Plot 207
References 213

5 Modeling Theory and Practice 217


5.1 Mathematical and Physical Modeling 217
5.2 Dimensional Analysis 218
5.2.1 The Procedure in Dimensional Analysis 218
5.2.2 The Buckingham Pi Theorem 219
5.3 Similarity 220
5.3.1 The Concept of Similarity 220
5.3.2 Model Laws 221
5.4 Application to Statically Loaded Elastic Structures 223
5.4.1 Prescribed Loads 223
5.4.2 Displacements and Strains 226
5.5 Loading Beyond Proportional and Elastic Limits 228
5.6 Buckling Experiments 229
5.6.1 Similarity Considerations for Buckling 229
5.6.2 Choice of Materials for Buckling Experiments 230
5.6.3 Elasto-Plastic Buckling 232
5.6.4 Goodier and Thomson’s Experiments on Shear Panels 234
5.7 Scaling of Dynamically Loaded Structures 237
5.7.1 Free Vibrations 238
5.7.2 Impact of a Rigid Body on a Structure 238
5.7.3 Scale Model Testing for Impact Loading 241
5.7.4 Plates Subjected to Impulsive Normal Loading 251
5.7.5 Response of Structures to Blast Loading 254
5.8 Scaling of Composite Structures 259
5.8.1 Problems in Scaling of Laminated Composites 259
5.8.2 Scaling Rules for Laminated Beams and Plates 260
5.8.3 Scaling for Strength and Large Deflections of
Composites 260
5.8.4 Scaling of Composite Plates 268
5.8.5 Scaling of Composite Cylindrical Shells 270
viii Contents

5.9 Model Analysis in Structural Engineering 272


5.9.1 Model Analysis as a Design Tool 272
5.9.2 Model Analysis in Vibration Studies 273
5.9.3 Buckling Experiments on Models of a Composite Ship
Hull Structure 275
5.9.4 Design of Thames Barrier Gates 279
5.9.5 Photoelastic Models 281
5.10 Analogies 282
References 283

6 Columns, Beams and Frameworks 289


6.1 Buckling and Postbuckling of Columns 289
6.1.1 Column Curves and “Secondary” Effects in Column
Experiments 289
6.1.2 Column Testing 294
6.1.3 Test Procedures 297
a. Preparation of Specimens 299
b. Initial Dimensions 299
c. Aligning the Column Specimen 299
d. Instrumentation 299
e. Testing 300
f. Presentation of Test Data 302
g. Evaluation of Test Results 303
6.1.4 Columns in Offshore Structures 303
6.1.5 End-Fitting Effects in Column Tests 304
6.2 Crippling Strength 309
6.2.1 Crippling Failure 309
6.2.2 Gerard’s Method for Calculation of Crippling
Stresses 310
6.2.3 Crippling Strength Tests 311
6.2.4 Crinkly Collapse 314
6.2.5 Thin-Walled Cold-Formed and Welded Columns 315
6.3 Torsional-Flexural and Distortional Buckling 320
6.3.1 Torsional Buckling 320
6.3.2 Torsional-Flexural Buckling Tests 320
6.3.3 Distortional Buckling 326
6.4 Lateral Buckling of Beams 328
6.4.1 Lateral instability of beams 328
6.4.2 Prandtl’s Lateral Buckling Experiments 329
6.4.3 Other Early Lateral Buckling Tests 330
6.4.4 Recent Lateral Buckling Investigations 333
6.5 Interactive Buckling in Columns and Beams 344
6.5.1 Mode Interaction and Early Studies 344
6.5.2 Interactive Buckling Experiments 345
6.6 Beam-Columns 356
6.6.1 Beam-Columns as Structural Elements 356
6.6.2 Recent Experiments on Tubular Beam-Columns 357
Contents ix

6.7 Buckling of Frameworks 367


6.7.1 Frame instability 367
6.7.2 Tests on Model Frames 369
6.7.3 Behavior of Connections 371
6.7.4 Seismic Loads on Multi-Story Frames 377
6.7.5 Space Structures 392
References 397

7 Arches and Rings 409


7.1 Background 409
7.2 Shallow Arches 410
7.2.1 Arches Under Concentrated Loads 410
a. Circular Arch 420
b. Sinusoidal Arch 422
7.2.2 Arches Under Uniform Pressure Loading 427
7.2.3 Additional Empirical Investigations 434
7.3 Rings and High Rise Arches 434
7.3.1 Rings Contact Buckling 434
7.3.2 High Rise Arches 439
7.4 Lateral Buckling of Arches 440
7.4.1 Theoretical Background 440
7.4.2 Experimental Studies 443
References 450

8 Plate Buckling 453


8.1 Buckling and Postbuckling of Plates 453
8.1.1 Historical Background 453
8.1.2 Effective Width 455
8.1.3 Postbuckling Behavior and “Secondary Buckling” 459
8.1.4 Influence of Geometric Imperfections 464
8.1.5 Influence of Residual Stresses 465
8.2 Experiments on Axially Compressed Plates 470
8.2.1 The US Bureau of Standards Test Setup 470
8.2.2 Needle and Roller Bearings and Knife Edges for
Simple Supports 473
8.2.3 The ETH Zurich and US Navy DTMB Plate Buckling Tests 479
8.2.4 The Cambridge University “Finger” Supports 484
8.2.5 Other Examples of Simple and Clamped Supports 491
8.2.6 Loading Systems 498
8.2.7 Large Test Rigs 503
8.2.8 Special Loading Systems for Annular Plates 505
8.2.9 Deflection Measurement 508
8.2.10 Controlled (Deliberate) Initial Deflections 512
8.3 Determination of Critical Load and Southwell’s Method
in Plates 516
8.3.1 Definition of the Buckling Load in Plates 516
8.3.2 Southwell’s Method in Plates 520
x Contents

8.3.3 Pivotal Plots for Plates 528


8.3.4 More Recent Applications of Southwell Plots and
Recommendations 531
8.3.5 Summary of Direct Methods for Determination of Buckling
Loads in Plates 533
8.4 Experiments on Shear Panels 538
8.4.1 Buckling and Postbuckling of Shear Panels 538
8.4.2 Experiments on Plates Subjected to Shear Picture Frames 542
8.4.3 Strength Tests on Plate Girders Under Shear 546
8.4.4 Technion Repeated Buckling Tests on Shear Panels 552
8.4.5 Aerospace Industrial Test Setups 558
8.5 Web Crippling 561
8.5.1 Web Crippling Due to Concentrated or Patch Loads 561
8.5.2 Web Crippling Tests 564
8.6 Biaxial Loading 570
8.6.1 Plates Under Multiple Loading 570
8.6.2 Biaxial In-Plane Compression Tests 570
8.7 Guidelines to Modern Plate Buckling Experiments 577
8.7.1 Guidelines or Ideas for Future Tests 577
8.7.2 Noteworthy Details in Some Modern Plate Tests 582
8.7.3 Imperial College London High Stiffness Test Machine 588
References 591

Author Index to Vol. 1 603

Subject Index to Vol. 1 611


Abbreviated Table of
Contents - Vol. 2

Vol. 2: Shells, Built-Up Structures and Additional Topics


Preface
9 Shell Buckling Experiments
9.1 Introduction
9.2 Buckling and Postbuckling Behaviour of Axially
Compressed Cylindrical Shells
9.3 Model Fabrication for Isotropic Shells
9.4 Test Setups for Cylindrical Shells Under Axial Compression
9.5 Recording of Buckling and Postbuckling Behaviour
9.6 Southwell’s Method for Shells
9.7 Cylindrical
Shells Under External Pressure, Bending or Torsion
9.8 Combined Loading
9.9 Conical Shells
9.10 Spherical Shells
9.11 Toroidal Shells, Torispherical Shells, Buckling Under
Internal Pressure
9.12 Shells Subjected to Transverse Shear Loads

10 Initial Imperfections
10.1 Introduction
10.2 Early Incomplete Imperfection Surveys
10.3 Early Complete Imperfection Surveys
10.4 The Awakening of Imperfection Measurement Awareness
10.5 Complete Imperfection Surveys on Large or Full Scale
Cylindrical Shells
10.6 Imperfection Surveys on Large Shells of Revolution
10.7 Recent Laboratory Scale Imperfection Measurement
Systems
10.8 Evaluation of Imperfection Data
xii Abbreviated Table of Contents - Vol. 2

10.9 Characteristic Initial Imperfection Distributions


10.10 Imperfection Data Banks
10.11 Probabilistic Design Methods
10.12 Residual Stresses
10.13 Imperfection Measurements and Data Banks in Columns
and Plates
10.14 Concluding Remarks

11 Boundary Conditions and Loading Conditions


11.1 Boundary Conditions in Column Buckling
11.2 Boundary Conditions in Plate Buckling
11.3 Boundary
Conditions in Buckling of Circular Cylindrical Shells
11.4 Concluding Remarks

12 Stiffened Plates
12.1 Built-Up Structures, Local and General Instability
12.2 Buckling and Postbuckling Strength of Stiffened Plates
12.3 Experiments
on Stiffened Plates Subjected to Axial Compression
12.4 Sandwich Plates

13 Stiffened Shells
13.1 Global and Local Buckling of Stiffened Shells
13.2 Model Fabrication for Stiffened Shells
13.3 Experiments on Stiffened Cylindrical Shells Subject to
Axial Compression
13.4 Experiments on Stiffened Cylindrical Shells Under
External Pressure, Bending and Torsion
13.5 Stiffened Conical and Spherical Shells
13.6 Experiments on Stiffened Curved Panels
13.7 Special Stiffened Shells

14 Composite Structures
14.1 Background
14.2 Flat Panels
14.3 Wing Box Structures
14.4 Curved Panels and Shells

15 Nondestructive Buckling Tests


15.1 Nondestructive Methods for Buckling Tests
15.2 Vibration Correlation Techniques (VCT)
15.3 Static Nondestructive Methods
Abbreviated Table of Contents - Vol. 2 xiii

16 Plastic Buckling Experiments


16.1 Plastic Buckling Phenomena
16.2 Plastic Buckling Experiments
16.3 Combined Loading Tests in Plastic Buckling
16.4 Southwell’s Method in the Plastic Range
16.5 Some General Remarks on Plastic Buckling

17 Influence of Holes, Cutouts and Damaged Structures


17.1 Effect of Holes and Cutouts on Plates and Shells
17.2 Experiments on Plates with Holes and Cutouts
17.3 Experiments on Shells with Holes and Cutouts
17.4 Stability and Strength of Damaged or Dented Shells

18 Buckling Under Dynamic Loads and Special Problems


18.1 Dynamic Buckling Phenomena
18.2 Impact Induced Buckling Experiments
18.3 Propagating Buckles

19 Thermal Buckling and Creep Buckling


19.1 Introduction
19.2 High Temperature Testing
19.3 Thermal Buckling
19.4 Creep Buckling

20 Some Comments on Measurements


20.1 Introduction
20.2 Strain
20.3 Displacement Sensors
20.4 Optical Methods
20.5 Data Acqusition Systems
20.6 Additional Sensing Devices
20.7 Summary

Author Index (for Vol. 1 and Vol. 2)


Subject Index (for Vol. 1 and Vol. 2)
Preface

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 experiment and theory, giving a detailed
and critical review of the many important buckling experiments 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 experiments, which are often described in great detail, with some comments
focusing on questions 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 accumulated experience would be useful to
the less experienced experimenter.
The wise experimenter should approach his 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 chapters. 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 texts 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
xvi Preface

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.
One of the guide lines (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 Kármán, 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 stiffened 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 detailed consideration, as they are still the main source of experience (or
data bank) to which a young experimenter should turn to.
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 there important research activities transcend
national boundaries and specific disciplines. They expose buckling experimentalists
to the vistas of benefits to be gained from the experience accumulated throughout
the many laboratories all over the world, as well as clarifying 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 appro-
priately acknowledged throughout the book. In some sections, it was felt fitting to
quote verbatim from some papers, reports and correspondence, 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,
Preface xvii

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. Kyriakides (University of Texas), A. Libai
(Technion), N.W. Murray (Monash University, Melbourne), H. Öry (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
Aerospace Structures Research Fund, the Jordan and Irene Tark Aerospace Struc-
tures Research Fund and the Caltech Sherman Fairchild Distinguished Scholars
Fund for their generous support.
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 understanding we could not have completed this book.

Josef Singer
Johaum Arbocz
Tanchum Weller
1
Introduction

1.1 Experiments as Essential Links in Structural


Mechanics
Stress analysis, structural analysis, or structural mechanics is the engineering disci-
pline, the purpose of which is the determination and improvement of the strength
and stiffness of structures and machines. There are two approaches towards this
goal: theoretical mechanics, focusing in recent years increasingly on numerical
mechanics, and experimental mechanics, with a mushrooming capacity of data
acquisition and reduction. The two approaches are intrinsically complementary,
though this is sometimes forgotten. In 1950 Hetenyi [1.1] presents the close relation
between theory and experiment as: “Experimental stress analysis strives to achieve
these aims (of determination and improvement of strength) by experimental means.
In doing so it does not remain, however, a mere counterpart of theoretical methods
of stress analysis but encompasses those, utilizing all the conclusions reached by
theoretical considerations, and goes far beyond them in maintaining direct contact
with the true physical characteristics of the problems under considerations.”
A decade or so later Drucker [1.2], [1.3] pointed out, however, “that all too
often, experimental work in applied mechanics is thought of only as a check on
existing theory or as a convenient substitute for analysis. This is a valid but a rather
inferior function of experiment. The greater and essential contribution is to guide
the development of theory by providing the fundamental basis for an understanding
of the real world.” He then concluded “that a researcher who remained in any
field would have to participate in both theory and experiment in order to remain
productive”.
Drucker concluded his 1962 General Lecture [1.2] with a warning that there is “a
strong steady drift of far too large a fraction of the best students” to theory only, and
that . . . “Unless appreciable numbers of the most qualified students aim at combined
experimental and theoretical research, the storehouse of physical information will
be depleted by the tremendous emphasis on analysis and theory, and the theorist
will be reduced to playing useless games. Experiment is essential, it is vital, and it
is creative. Over the years, experiment alone provides the basis for the refinement
and extension of existing theory and the development of new theory.”

Buckling Experiments: Experimental Methods in Buckling of Thin-Walled Structures: Basic Concepts, Columns, Beams and
Plates – Volume 1. J. Singer, J. Arbocz and T. Weller Copyright © 1998 John Wiley & Sons, Inc.
2 Introduction

An example in the field of elastic stability of what Drucker referred to as “playing


useless games” was presented by Koiter, in his 1985 Prandtl Lecture [1.4], where
he discussed the physical significance of instability due to non-conservative, purely
configuration-dependent, external loads. Koiter reminds the reader that . . . “In the past
decades much attention has been paid to stability problems of elastic structures under
the action of non-conservative purely configuration-dependent loads, e.g. so-called
follower forces whose directions follow deflections of the structure . . .”. In evaluating
the “physical significance or rather insignificance of such follower forces,” Koiter
quotes remarks by Herrmann in the latter’s 1967 review article [1.5]: “It is a peculiar
feature of stability problems of elastic systems subjected to (non-conservative)
follower forces that their analysis arose not out of a desire or need to consider a
system which presented itself in engineering practice or in the research laboratory,
but rather because the fictitiously applied follower forces acting on a given system
were arbitrarily prescribed to depend in a certain manner on the deformation. The
motivation of much if not most of the work mentioned in this survey appears to
have been sheer curiosity in determining the sometimes unexpected behavior of an
imagined system, rather than an explanation of observed phenomena.”
Koiter then emphasizes that: “The italicized emphasis of ‘fictitiously applied
follower forces’ and ‘imagined system’ is the present writer’s.” Engineers are
indeed rightly concerned about the complete lack of a physical justification of the
concept of follower forces in the analysis of stability problems by many authors.
Not a single experimental verification of this concept is to be found in the exten-
sive literature. The invalid example by Willems [1.6] was soon exposed [1.7], and
one wonders at the optimism expressed by some authors, e.g. [1.8], that “Beck’s
result (for follower forces) can probably be experimentally verified”. Koiter then
continues that: “The domain of the theory of elastic stability is conventionally
restricted to the stability of equilibrium of elastic structures under the action of
static, purely configuration-dependent external loads” and concludes that: “Since
no physical example of non-conservative, follower type of purely configuration-
dependent external loads is available, we are entitled to restrict our attention to the
stability of elastic equilibrium under the action of static conservative loads”. One
may reflect, that if emphasis would have been placed on accompanying experi-
mental research, the large volume of rather academic studies might have perhaps
been directed into more fruitful avenues.
In his 1967 Murrey Lecture [1.3], Drucker struck a more optimistic note than in
1962, and pointed out that experimental work was beginning again to take a more
important place in solid and structural mechanics research. Unfortunately this trend
was reversed again in the last decade, and Drucker’s 1962 warning [1.2], which
was quoted earlier, is today once more very appropriate.
Drucker then emphasized again the essentially complementary nature of experi-
ment and theory, stating: “Theory awaits experiment and experiment awaits theory in
a wide variety of fields. Often the two must go hand in hand if any significant progress
is to be made.” Then he emphasizes the basic similarity between a good experiment
and a good theory and states that: “The most important thought process that goes
into the planning of an experiment is exactly the same as in the development of a
The Role of Experiments in Structural Stability 3

theory. Success in either requires identification of the essential variables and param-
eters along with an appropriate limitation of the objective of theory or experiment.”
. . . “The choice of environment for static or dynamic problems of elastic response,
of flow, or of fracture in all but the best-known examples, involves all the elements
of thought which enter into the development of a theory.” For example, questions
like “how does the response of a part under examination affect the environment or
boundary conditions?” require both precise theory and careful experiments.
Hence theoretical and experimental mechanics have to progress hand in hand.
And as the rapid advances in computational methods and tools have enormously
broadened the horizons of theory, so have the triumphs of microelectronics and
ever more efficient computers brought about a virtual revolution in instrumentation,
introducing so-called intelligent instruments, which have multiplied our measure-
ment capabilities and accuracies (see for example [1.9] or [1.10]). However, as in
theory so in experiment, it is the basic thought process that precedes the actual
study, which guides and harnesses these capabilities.

1.2 The Role of Experiments in Structural Stability


Structural stability research in the 19th century was primarily experimental, or more
precisely, empirical. Near the turn of the century theoretical studies took the lead
and continued to do so in the 20th century, unfortunately in many cases without
correlation with experiments. Koiter painted this state of affairs in the Opening
Lecture of the 1974 IUTAM Symposium on Buckling of Structures [1.11]: “To put
it mildly, buckling theory and experiments have not always co-existed in harmony”.
One should remember that though judiciously chosen mathematical models may
predict the expected physical behavior, it is up to careful experiments to verify this
predicted behavior and validate the calculations. Furthermore, the experiments may
bring out elements of behavior of real structures, which have not been considered
in our, by necessity, simplified models. This second role of experimentation is
often overlooked.
The pattern of research in structural stability for many years has been one of
extensive theoretical studies combined at the most with corroborating experiments.
As pointed out by Chilver [1.12], this has been very useful in the study of essen-
tially neutral equilibrium problems of elastic stability. But in cases of extreme
instability, theory has only been a guide to practical behavior, and much of our
present useful design knowledge is based on careful experiments. The important
problems of stability, and in particular, postbuckling behavior, are not always
amenable to complete analysis, and accurate analyses may be rather difficult and
the computations very cumbersome. For example, Zandonini, in a 1983 review of
the stability of steel compression members [1.13], states that “since the problem
of determining the ultimate strength of a column (in the presence of geometrical
imperfections and residual stresses) can only be solved analytically in a very limited
number of cases, the experimental approach and inelastic second order numerical
analysis have become fundamental tools”. Note that if this is the state-of-the-art
4 Introduction

in the case of columns, the importance of experimental studies for other more
complex structures becomes evident.
And indeed, more prominence has been given in recent years to experimental
studies in buckling research and their interaction with theory, as is apparent, for
example, in reviews of the state-of-the-art of shell stability (like [1.14] [1.18]).
The trend in structural stability has been towards more awareness of the potential
of experimental studies and a beginning of more cooperation between theoretical
and experimental research. An interesting example of this awakening awareness
is the acknowledgement by two eminent theoreticians, Budiansky and Hutchinson,
in their 1979 survey of buckling [1.19], that with respect to practical design opti-
mization problems dominated by buckling behavior “theory lags experiment”. They
pointed out that a remarkable series of tests conducted in the mid-forties at NACA
Langley ([1.20] and [1.21]) provided an experimental optimization, or minimum
weight designs, for stiffened flat panels. In these tests over 150 2024-ST aluminum
alloy zee- and hat-stiffened panels, having systematically varied configurations,
were tested for ultimate compressive strength m Figure 1.1 (from [1.21]) shows

50

z-stiffened panels
40

Hat-stiffened panels

30
σm, ksi

20
ts
tw

10

ts

0
0 .2 .4 .6
Pi , ksi
L/ √C

Figure 1.1 Experimental optimization, or minimum weight designs, for stiffened flat panels.
Ultimate compressive strength tests of over 150 2024-ST aluminum alloy zee-
and hat-stiffened panels, carried out at NACA Langley in the mid-forties, yielded
envelope curves for the two types of stiffeners (from [1.21])
Motivation for Experiments 5

a comparison of the envelope curves of the ultimate compressive strengths


p of
minimum weight designs for the two types of stiffeners, where Pi L/ C is the
appropriate structural index, with Pi the load-per-unit-width of panel, L the panel
length and C an end-fixity coefficient. The most striking feature of these 50-year-
old experimental results is that they show (at least for equal sheet and stiffener
thicknesses) a superior structural efficiency of zee- over hat- stiffeners. “These
experimental results automatically incorporate the effects of certain representative
initial imperfections, not to mention plasticity, discrete rivet attachments, finite
corner radii and they stand as a challenge to theoreticians to confirm or refute
them, and deduce analogous results for other configurations and materials.”
However, it is important to remember that, just as it is unwise to regard exper-
iments as only a check on existing theory, it is as imprudent to be too “practical”
and base one’s design on empirical data only, especially if this data was obtained
under conditions which differ significantly from that of the actual structure. In
structural stability “the proper marriage of theory and experiment is essential”, as
Sechler emphasized again in 1980 in relation to shell research ([1.22]), and only
when they go hand-in-hand is there rapid progress.

1.3 Motivation for Experiments

With the rapid development in computers in the last decades the question of “why
continue to do experiments?” has often been asked in many fields of applied
mechanics. As the computational tools improved and expanded, the idea that
computer simulations can replace the experiments has been voiced occasionally.
For example, in the early sixties computer simulated experiments became popular
and, in the excitement about their advantages and potential, their limitations were
forgotten. For instance Johnston in 1961 [1.23] claimed: “There are many advan-
tages in simulated tests, carried out with the aid of a computer, in comparison
with real tests in an actual testing machine. No machining is involved, no mate-
rials need be acquired, and there is no scatter in the test results! Moreover, the
precision of results, although based on a simulated and idealized material, permits
a study of details of behavior that is not possible in ordinary laboratory tests. It
would be impossible to completely duplicate the observations that may be made
on the basis of the simulated tests reported in this paper.” It was forgotten that the
simulation was so successful because the physical phenomena in this case were
well known and had been extensively explored by very many real experiments.
New phenomena have still to be found and properly understood in physical tests,
before even the powerful computers of today can give a reliable simulation and
then extend the range of parameters.
In a similar vein was the false 1975 prediction for aerodynamics that “Wind
tunnels in 10 years will be used only to store computer print out”.
Hence it is worth the while to reflect in more detail on the purpose of experiments
in the computer era. The question was examined for shell buckling in two reviews
6 Introduction

in the eighties ([1.17] and [1.24]); it will now be re-examined in the broader
context of buckling and postbuckling behavior of structures. One can enumerate
eight primary motives:

(A) Better understanding of buckling and postbuckling behavior and the primary
factors affecting it. In addition to the buckling loads, careful experiments
in which the parameters are varied one at a time yield the behavior of
the structure just before, at and after buckling, and accentuate the main
parameters affecting this behavior. Such a philosophy of “research type
experimental programs” has been strongly advocated for shells by Sechler
[1.22] for many years, and has been implemented in some test programs, for
example in [1.25]. Based on these observed parameters numerical schemes
can be developed, verified, and can also be employed for “experiments on
the computer” to extend the range of the parameters tested. One should
remember that computer methods, like for example finite element analysis,
can converge to non-realistic behavior, unless the physical phenomena are
well understood, or at least well described by appropriate experimentation,
to permit reliable modeling.
(B) To find new phenomena. This reason is a direct extension of the first one
and has been stressed by Drucker ([1.2] and [1.3]), Sechler [1.22] and many
others. In buckling and postbuckling experiments, the new phenomena are
likely to be unexpected behavior patterns or mode interactions.
(C) To obtain better inputs for computations. The mathematical models employed
in modern large multi-purpose computer programs can simulate real
structures fairly closely for buckling, but the simulation depends very much
on the input of correct boundary conditions, in particular joints or bonds,
on material properties, imperfections, residual stresses and load applications.
This has been emphasized by recent experience and definitely applies also
to postbuckling. Better inputs can be provided by subsidiary tests like stub-
column tests for properties of columns or stiffeners, or multiaxial material
tests for more complicated structures or loading conditions. Often improved
inputs can be obtained from appropriate nondestructive tests: for example,
boundary conditions by vibration correlation techniques, imperfection shapes
and amplitudes by imperfection scans, load transfer and eccentricities by
strain measurements and vibration correlation techniques, residual stresses
by X-ray techniques etc. Fully automated recording in experiments has
just begun and much closer interaction between test and computation
is developing.
(D) To obtain correlation factors between analysis and test and for material
effects. Even when large powerful programs are employed, test results may
still differ considerably from predictions. These differences are partly due to
inaccuracies of inputs and partly to variations in buckling and postbuckling
behavior of the mathematical model and the structures tested. They can all be
lumped for design purposes in a “correlation factor”. The advantage of such a
correlation factor is the overall correlation it provides for the designer, but its
Motivation for Experiments 7

weakness is that it is completely reliable only for the structures tested. One
can statistically evaluate a large number of tests to obtain overall lower bound
correlation factors, a method employed extensively for shells, where they
are called “knock-down” factors, but this results in very conservative design.
Other statistical evaluations are extensively employed for columns and plates,
especially for civil and marine engineering design codes, and these too tend
to be conservative. Hence “correlation factors” should be more specialized.
Since many experiments are on laboratory scale structures, extensive studies
comparing the results of laboratory scale and large scale tests are needed
to reassure the experimenter and to guide the designer, in particular for
dynamic loading. Correlation type experiments will therefore continue to be
a major task of research and industrial laboratories for quite some time to
come, as they provide the designer with essential correction factors which
include the effects of new materials and manufacturing techniques and, to
some extent, bridge the gap between the buckling and postbuckling behavior
of the computation model and the realistic structures.
(E) To build confidence in multipurpose computer programs. Extensive experi-
mental verification is an essential element for confidence in a large computer
program. This is therefore a primary motive for buckling and postbuck-
ling experiments, which becomes more important, as the programs become
more sophisticated and ambitious. Though some developers of programs have
promoted and applied extensive experimental confirmation, more correla-
tions of the results obtained from computer programs with test results are
required, as pointed out for example for shells in [1.26]. An example of exten-
sive experimental verification, as well as careful examination of boundary
conditions by combined experimental and numerical studies, is the effort of
Bushnell in building confidence in his BOSOR4 and 5 shell programs ([1.27]
and many others).
(F) To test novel ideas of construction or very complicated elements of a
structure. Exploratory tests of new concepts have been used extensively by
aeronautical, civil, mechanical and ocean engineers, and will continue to be
an important tool. Furthermore, if the structure is elaborate and has many
openings with complicated stiffening and load diffusion elements, model
testing may sometimes be less expensive and faster than computation with a
large multipurpose program, even in the detail design state.
(G) For buckling under dynamic loading and in fluid-structures interaction
problems. These are areas where computation is cumbersome, expensive,
and difficult to interpret reliably. Experiments may therefore be preferable at
this stage, though they too present many difficulties. Theory and numerical
computations should follow these experiments closely, to reinforce and
broaden the partial understanding of the phenomena that the experiments
will provide.
(H) For certification tests of full scale structures. This is the typical industrial
task (see for example Figure 1.2), which will continue till model experiments
8 Introduction

Figure 1.2 A modern full scale aircraft certification test. The Boeing 757 airliner lower forward
fuselage during a test, which illustrates combined compression and shear dominated
buckling (typical of semi-monocoque construction). The photo was taken at 100
percent design limit load (courtesy of the Boeing Commercial Airplane Company)

are sufficiently advanced and integrated with computation to eliminate the


necessity for them. Here computerization of data acquisition and reduc-
tion has made great strides, and has significantly advanced the accuracy
of measurement and interpretation.

Examination of these motives, originally proposed for shells at a Euromech


Colloquium in 1980 [1.24], and recent experience reinforces the conclusion that the
computer does not replace the experiments. It may change their purpose somewhat,
it modifies the techniques, it broadens the capability to acquire results and it can
use the experimental results to improve the computations. The presence of the
computer in the experimental scene enhances and develops new techniques and
capabilties. As pointed out by Birkemoe of the University of Toronto in 1994 (see
[8.162]): “High speed and high quality data acquisition, combined with on-line use
of the data for control of the loading and/or the response of a boundary condition,
present a framework for improved experimental demonstration of stability limits in
structures.” Furthermore, “Improvements in user software for the test environment
continue to make . . . computer control easier”. The experiment remains an essential
link in the analysis also in the computer era, and its scope and usefulness are even
greater today.
It appears therefore that experiments are indeed essential tools in structural
stability research. Why do many investigators still shy away from them? One reason
Bridging Gaps Between Disciplines 9

may be the initial difficulties facing the inexperienced researcher. To quote, for
example, from a 1967 predominantly theoretical doctoral thesis in civil engineering
“Postbuckling Behavior of Tee Shaped Aluminum Columns” by R. Hariri [1.28]:
“The author experienced a great deal of difficulties and some experiments in the
early stages yielded surprising and unexpected results”. He then concludes this
paragraph: “However, the experiences gained and the guidance obtained cannot
be overlooked”. One purpose of this book is to reduce the difficulties in the early
stages and open up the wide horizons of experimental research, and the potential
guidance to the physical phenomena it can provide, to more of the younger, as yet
inexperienced, investigators.

1.4 Bridging Gaps Between Disciplines


As the reader may have noted, the authors of this book hold the view that exper-
iment and theory are complementary and that real progress is contingent upon
experiment and theory proceeding hand-in-hand. Hence, though the book is devoted
to experimental methods, the next two chapters present a discussion and summary
of physical and theoretical concepts as well as analytical and computational tools.
In addition to bridging the gap between the theoretician and the experimentalist,
we also attempt to bridge the gap between the different engineering disciplines
which have to deal with buckling problems. As did the medieval guilds from which
they originated, the various engineering professions tended to keep to themselves,
to their institutions and societies, and develop their own traditions and methods.
Technology transfer between the different disciplines has really only started in
recent decades, and structural stability has been one of the more active fields of
this transfer of knowledge and techniques.
As an example of the difference in approach of civil and aerospace engineers,
one may consider H-section columns. The aeronautical engineers of the late thirties
were interested in the strength of aluminum-alloy extruded H-section columns,
and extensive tests were carried out at the U.S. National Bureau of Standards and
the Aluminum Company of America, under the sponsorship of NACA [1.29], to
provide column curves. These test columns were slender but did not have very thin
webs. Thin-web cross-sections were also widely used as stiffeners in the aircraft
industry, but these were usually made from bent sheet. The local failure modes
(buckling of outstanding flanges as plates), that occurred in these thin-web columns
were also studied extensively in the thirties.
The civil engineers first considered rolled H-section columns, which also did
not have thin webs. Then heavy welded H-sections were studied and used in prac-
tice. The civil engineers studied columns made of structural steel, since this was
the material employed in civil engineering structures. Already the rolling process
introduced some residual stresses, but in welded columns these increased signifi-
cantly, and the effect of residual stresses on the buckling strength became a major
concern. However, little attention was paid to thin-web columns until the early
eighties, when their use increased in manufactured metal building, and a detailed
10 Introduction

experimental study on buckling of thin-web welded H-columns was carried out (see
for example [1.30]). Again the major concern of the civil engineering researchers
was the influence of residual stresses, and they utilized the experience accumulated
by their colleagues. There has, however, been only very little, if any, use by civil
engineers of the aeronautical research and experience, maybe because in this case it
dates back primarily to the thirties and forties, or perhaps because the aeronautical
engineers did not seriously consider this effect of residual stresses. But there could
be more mutual enrichment, or technology transfer, as occurred for example in the
case of stiffened shells in offshore structures, where the aeronautical experience
has been absorbed by the marine and civil engineers (see for example [1.31]).
The civil engineers have been and still are preoccupied with design codes, whose
purpose is to ensure safe structures even when designed by independent engineers
who do not have large staff and extensive computing facilities at their disposal,
and to serve as tools for the proof engineers and licensing officials in their certifi-
cation tasks.
Other engineering disciplines have their design codes too, but rely more on
extensive computational and experimental proof of sufficient strength for safety
and certification. An aerospace engineer, for example, will therefore be surprised
at the continuous correlation and comparison with different design codes in most
experimental and theoretical studies reported in civil engineering publications. His
civil engineering colleague on the other hand will similarly be astonished both
at the numerous different loading cases computed and at the extent of full-scale
“limit load”, ultimate load and fatigue testing required in aerospace practice, as
well as at the very limited employment of design codes. These different approaches
and practices should, however, not deter researchers and engineers of different
disciplines to use each other’s investigations, and in particular experimental studies.
By referring to examples from different disciplines and pointing out the similarities
we will try to guide the reader in this direction.
It is of interest to note that civil engineers themselves have in recent years
been clamouring for confining their codes, which had become far too detailed,
“to principles of design, identifying structural requirements which must be satis-
fied for different classes of structures” [1.32]. These more basic codes would be
supplemented by “cross references to data sheets which give design procedures
satisfying these principles, but leave the designer the choice of method best suited
to his requirements”. The data sheets or data banks would replace “voluminous
codes crammed with complex formulae”. This trend, which also facilitates inter-
national agreement, will bring civil engineering practice closer to the approach of
other engineering disciplines, like aerospace or mechanical.
Professor Dowling of Imperial College, London who made this appeal in 1981,
also pointed out in 1982 a somewhat absurd situation that had arisen in the design
codes for the strength of webs of plate girders [1.33]. He stressed that there would
be large differences in the designs for the same loading produced to the relevant
Swiss, German, British or U.S. codes, and indicated that disparities exist between
the new and draft codes in many countries. Dowling then pleaded for involvement
of the appropriate international bodies to coordinate efforts towards unified research
References 11

and codes for plates and concluded optimistically that “If it was possible to produce
the European Column Strength Curves through international cooperation, surely it
is possible to produce European, or indeed World Plate Strength Curves. Then a
welded steel plate might not know it has changed strength as it crossed the border
between Austria and Germany”! The gap between countries and disciplines appears
indeed to be narrowing.
Another gap that requires some bridging is a kind of “generation gap”. When
searching the literature for previous studies, the young researcher or test engi-
neer will usually focus his attention on recent publications, as he will assume
that only experiments carried out with modern instrumentation and techniques can
be of any relevance to his present-day problem. Furthermore, since his search
will nearly always be carried out with a computerized literature search, which
practically excludes any publications earlier than 25 years prior to the date of
search, many important earlier studies will have escaped his notice. A danger of
“rediscovering America” with more modern means is then imminent. Hence the
authors wish to stress that one should also look to earlier experimental studies,
which, though carried out with less sophisticated instruments, often excelled in the
planning and logic of the experiments and in pin-pointing essential primary and
secondary effects. This is not surprising, since the tests were often carried out by
some of the outstanding scientists and engineers of the time. We will even show
some examples of important experiments, performed more than 90 years ago, but
whose logic and results are still applicable today. The authors hope the reader will
develop respect for the “ancients” also in experimental mechanics.

References

1.1 Hetenyi, M., Handbook of Experimental Stress Analysis, 1st ed., John Wiley & Sons,
New York, 1950, Preface.
1.2 Drucker, D.C., On the Role of Experiment in the Development of Theory, General
Lecture, Proc. 4th US National Congress of Applied Mechanics, ASME, 1962, 15 33.
1.3 Drucker, D.C., Thoughts on the Present and Future Interrelation of Theoretical
and Experimental Mechanics, William M. Murrey Lecture 1967, Experimental
Mechanics, 8, (3), 1968, 97 106.
1.4 Koiter, W.T., Elastic Stability, 28th Ludwig Prandtl Memorial Lecture, Zeitschrift
für Flugwissenschaften und Weltraumforschung, 9, (4), 1985, 205 210.
1.5 Herrmann, G., Stability of Equilibrium of Elastic Systems Subjected to Non-
conservative Forces, Applied Mechanics Reviews, 20, 1967, 103 108.
1.6 Willems, N., Experimental Verification of the Dynamic Stability of a Tangentially
Loaded Cantilever Column, Journal of Applied Mechanics, 33, 1966, 460 461.
1.7 Huang, N.C., Nachbar, W. and Nemat-Nasser, S., On Willems’ Experimental Veri-
fication of the Critical Load in Beck’s Problem, Journal of Applied Mechanics, 34,
1967, 243 245.
1.8 Kolkka, R.W., On the Non-Linear Beck’s Problem with External Damping, Interna-
tional Journal of Non-Linear Mechanics, 14, 1984, 497 505, (in particular paragraph
2 of the introduction).
12 Introduction

1.9 Pindera, J.-T., Patterns and Trends of Advanced Experimental Mechanics, Proceed-
ings of the 11th Canadian Congress of Applied Mechanics, University of Alberta,
Edmonton, Canada, May 31 June 4, 1987, A206 A207.
1.10 Hirschfeld, T., Instrumentation in the Next Decade, Science, 230, 1985, 486 491.
1.11 Koiter, W.T., Current Trends in the Theory of Buckling, in Buckling of Structures,
Proceedings of IUTAM Symposium on Buckling of Structures, Harvard University,
Cambridge, USA, June 17 21, 1974, B. Budiansky, ed., Springer-Verlag, Berlin,
1976, 1 16.
1.12 Chilver, A.H., The Role of Experimentation in the Study of Elastic Stability of
Structures, in: Stability, Solid Mechanics Division, SM Study No. 6, University of
Waterloo, Ontario, Canada, 1972, 63 84.
1.13 Zandonini, R., Recent Developments in the Field of Stability of Steel Compres-
sion Members, in Stability of Metal Structures, Proceedings, 3rd SSRC International
Colloquium, George Winter Memorial Session, Toronto, Canada, Structural Stability
Research Council, 1983, 1 19.
1.14 Arbocz, J., Past, Present and Future of Shell Stability Analysis, Zeitschrift für Flug-
wissenschaften und Weltraumforschung, 5, (6), 1981, 335 348.
1.15 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 Publish-
ing Co., 1980, 65 116.
1.16 Valsgard, S., and Foss, G., Buckling Research in Det norske Veritas, in: Buckling
of Shells in Offshore Structures, J.E. Harding, P.J. Dowling, and N. Agelidis, eds.,
Granada, London, 1982, 491 548.
1.17 Singer, J., The Status of Experimental Buckling Investigation of Shells, in: Buckling
of Shells, E. Ramm, ed., Proceedings of the State-of-the-Art Colloquium, Stuttgart,
Springer-Verlag, Berlin, Heidelberg, New York, 1982, 501 533.
1.18 Babcock, C.D., Shell Stability, Journal of Applied Mechanics, 50, 1983, 935 940.
1.19 Budiansky, B., and Hutchinson, J.W., Buckling: Progress and Challenge, in: Trends
in Solid Mechanics, J.F. Besseling and A.M.A. van der Heyden, eds., Delft Univer-
sity Press, 1979, 93 116.
1.20 Schuette, E.H., Charts for the Minimum-Weight Design of 24S-T Aluminum-Alloy
Flat Compression Panels with Longitudinal Z-Section Stiffeners, NACA Report
No. 827, 1945.
1.21 Hickman, W.A., and Dow, N.F., Compressive Strength of 24S-T Aluminum-Alloy
Flat Panels with Longitudinal Formed Hat-Section Stiffeners Having a Ratio of
Stiffener Thickness to Skin Thickness Equal to 1.00, NACA TN 1439, 1947.
1.22 Sechler, E.E., The Role of Experimentation in Shell Research, in: Mechanics Today,
5, S. Nemat-Nasser, ed., Pergamon Press, Oxford, 1980, 439 449.
1.23 Johnston, B.G., Buckling Behavior Above the Tangent Modulus Load, Journal of
the Engineering Mechanics Division, American Society of Civil Engineers, 87, EM6,
Paper 3019, Dec. 1961, 79 99.
1.24 Singer, J., Buckling Experiments on Shells a Review of Recent Developments, Solid
Mechanics Archives, 7, 1982, 213 313.
1.25 Singer, J., Arbocz, J., and Babcock, C.D., Buckling of Imperfect Stiffened Cylin-
drical Shells under Axial Compression, AIAA Journal, 9, (1), 1971, 68 75.
1.26 Buchert, K.P., Practical Application of Shell Research, in: Buckling of Shells in
Offshore Structures, J.E. Harding, P.J. Dowling, and N. Agelidis, eds., Granada,
London, 1982, 257 283.
References 13

1.27 Bushnell, D., BOSOR 5 Program for Buckling of Elastic-Plastic Complex Shells of
Revolution Including Large Deflections and Creep, Computers and Structures, 16,
1976, 221 239.
1.28 Hariri, R., Post Buckling Behavior of Tee Shaped Aluminum Columns, Doctoral
Thesis, University of Michigan, 1967, University Microfilms International, Ann
Arbor, Michigan.
1.29 Osgood, W.R., and Holt, M., The Column Strength of Two Extruded Aluminum-
Alloy H-Sections, NACA Report No. 656, 1939.
1.30 Avent, R.R., and Wells, S., Experimental Study of Thin-Web Welded H-Columns,
Journal of the Structural Division, Proceedings of the American Society of Civil Engi-
neers, 108, (ST7), 1982, 1464 1480.
1.31 Singer, J., Buckling, Vibrations and Postbuckling of Stiffened Metal Cylindrical
Shells, Proceedings of BOSS 1976 (1st International Conference on Behavior of Off-
Shore Structures), Norwegian Institute of Technology, Trondheim, Norway, August
1976, 765 786.
1.32 Dowling, P.J., Editorial, The Journal of Constructional Steel Research, 1, (3), 1981,
1 2.
1.33 Dowling, P.J., Editorial, The Journal of Constructional Steel Research, 2, (3), 1982, 1.
2
The Concepts of
Elastic Stability

Before one can carry out meaningful experiments on buckling of structures one
has to understand the basic phenomena of structural instability and recognize
the different type of buckling behavior that may occur. Though this book deals
primarily with experimental methods and test results, and it is assumed that the
reader is somewhat familiar with the analysis of buckling, the theoretical concepts
of the basic instability phenomena and the numerical methods used to arrive at
numerical solutions are reviewed in this and the following chapter. This summary
will be brief and the reader may wish to consult some of the well known texts,
like [2.1] [2.8], for a broader introduction and more detailed treatment.

2.1 Physical Concepts Types of Observed Behavior


and Their Meaning
All structural designers know that their structure must satisfy two basic criteria,
namely:

1. the strength criterion, which states that under the specified (foreseeable)
loading conditions the maximum stresses may not exceed the allowable stresses
anywhere in the structure;
2. the stiffness criterion, which specifies the maximum allowable deflections
under the different loading conditions in order not to hinder proper operation
or to avoid undesirable and potentially dangerous behavior such as flutter or
mechanical vibrations.
What often is overlooked is that by carrying out the usual stress and deformation
analysis with the many easily available finite element codes one obtains no infor-
mation as to the stability behavior of the structure. It is by now well known that
thin-walled slender structures, or structures which contain slender members subject
to compressive stresses, may initially fail in one of the many possible instability

Buckling Experiments: Experimental Methods in Buckling of Thin-Walled Structures: Basic Concepts, Columns, Beams and
Plates – Volume 1. J. Singer, J. Arbocz and T. Weller Copyright © 1998 John Wiley & Sons, Inc.
16 The Concepts of Elastic Stability

modes, which in turn may significantly affect the strength or stiffness behavior of
the whole structure. This is especially true for the current trends in design where
with the use of structural optimization techniques one is often producing highly
stressed structures of very slender proportions.
With the sudden and often unexpected occurrence of partial or total structural
failure due to different forms of (at least initially) elastic instabilities, one has
come to rely on so-called buckling tests to provide the data for the development of
safe and reliable design recommendations, as pointed out in Chapter 1. However,
before one can carry out meaningful experiments on buckling of structures, one
has to understand the basic phenomenon of structural instability and recognize the
different types of buckling behavior that may occur during the loading process of
an experiment. In the following, the occurrence of different types of instabilities
will be discussed on hand of relatively simple examples.

2.1.1 Instability of Columns

The problem of a slender, perfectly straight, centrally compressed column, built


in vertically at the base and free at the upper end (see Figure 2.1a) has been first
solved by Leonard Euler in 1744 [2.9]. He found as the smallest critical load
2 EI
Pc D 2.1
4 L2
where E is Young’s modulus, I is the moment of inertia of the cross-section and
the corresponding buckling mode is shown in Figure 2.1b. Euler has assumed in
his work that the cross-section of the column does not distort during buckling and
failure and that the wavelength of the buckling mode is of the order of the column
length.
The buckling load for other boundary conditions can be found easily by direct
solution of the corresponding eigenvalue problem. Thus for simply supported
boundary conditions one must solve (see, for example, [2.1] or [2.2])
wiv C k 2 w00 D 0 for 0  x  L 2.2
00
wDw D0 at x D 0, L

Figure 2.1 Euler’s problem


Physical Concepts Types of Observed Behavior and Their Meaning 17

where  0 D d/dx and k 2 D P/EI. Its solution is


EI 
Pn D n2 2 ; wn D Cn sin kn x; kn D Pn /EI. 2.3
L2
The smallest or critical buckling load occurs for n D 1. Notice that the higher
order buckling loads can be attained only by using very slender columns and
by applying external constraints at the points of inflection to prevent the lateral
deflection associated with the lower order modes.
The perfect column assumption is unrealistic. Using an initial imperfection of
the form
x
w0 x D W01 sin  2.4
L
and a large deflection theory Rivello [2.10] has obtained the results shown in
Figure 2.2. From this figure one can draw the conclusion that the straight position
is the only equilibrium configuration for a column with vanishingly small imper-
fections until P D Pc . Close to and at P D Pc the deflections of a column with
vanishingly small imperfections grow very fast and are approximately given by
Eq. (14.56) of [2.10] until on the concave side the stresses in the extreme fiber
exceed the proportional limit. As can be seen from Figure 2.2 also columns with
measurable imperfections do not bend appreciably until P is very nearly equal
to Pc . 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 Pc .
The dotted curves in Figure 2.2 representing the yield stress limits were also
obtained by Rivello [2.10] for a column with homogeneous cross-section using
an idealized linearly elastic and perfectly plastic material behavior. The dotted

Figure 2.2 Nonlinear behavior of perfect and imperfect columns (from [2.10])
18 The Concepts of Elastic Stability

curves were computed for an idealized H section made of 7075-T6 aluminum


alloy, whereby it was assumed that the web has negligible resistance in bending
and extension but is rigid in shear. It has been shown in [2.4] that if one considers
eccentrically applied axial loading in place of geometric initial imperfections, one
obtains curves similar to the ones shown in Figure 2.2. The instability theory of
Euler accurately describes the buckling behavior of slender columns with solid
or thick-walled cross sections. To obtain a direct measure of slenderness it is
customary to rewrite Euler’s formula (Eq. 2.3) as
Pc E
c D D 2 2.5
A L/2
p
where  = radius of gyration of the cross-section (D I/A). Experimental evidence
indicates that for values of the slenderness ratio L/ > 80 Euler’s formula predicts
the buckling load of columns quite accurately. For values of the slenderness ratio
20 < L/ < 80 one can get reasonably accurate predictions by using Shanley’s
tangent modulus theory [2.11], which essentially consists of replacing in Eq. (2.5)
the modulus of elasticity E by the tangent modulus Et . Finally, for values of the
slenderness ratio L/ < 20 failure occurs mainly by plastic crushing of the cross-
section and c is equal to the compressive strength of the material. For metals one
usually uses c D cy , where cy is the compressive yield strength of the material.
For thin-walled columns Euler’s assumptions that the cross-section does not
distort during buckling and that the wavelength of the buckle is of the order of
the column length must be reexamined. Such columns can be thought of as an
assemblage of thin plates. Thus before addressing the local instability and failure
analysis of thin-walled columns, first the stability analysis of thin plates loaded by
in-plane forces shall be considered.

2.1.2 Instability of Plates

The buckling load of a simply supported rectangular thin flat plate of width b and
length a, subjected to a uniform compressive force per unit length N D h on the
edges x D 0 and x D a while the boundaries y D 0 and y D b are unrestrained
against in-plane motion (see Figure 2.3) has first been derived by G.H. Bryan in
1891 [2.12]. Using the deflection mode shape
mx ny
w D Wmn sin sin where m, n D 1, 2, . . . 2.6
a b

Figure 2.3 Plate subjected to in-plane compressive loading


Physical Concepts Types of Observed Behavior and Their Meaning 19

he found as the smallest critical stress


 2
2 E h
c D kc 2.7
121  2  b
where the plate buckling coefficient kc is the minimum value of
 2
mb 2 a
kmn D Cn 2.8
a mb
obtained for a given plate aspect ratio a/b by proper selection of the integers m
and n. From Eq. (2.8) it is obvious that the minimum value of kmn occurs when
n D 1. To minimize Eq. (2.8) with respect to m, one plots kc as a function of a/b
for different values of m, as shown in Figure 2.4. The minimum value of kc , which
is then used in Eq. (2.7) is given by the lower envelope of the curves, indicated
in Figure 2.4 by the solid line. Considering the case of a long plate (say, a/b ½ 3)
then kc ³ 4.0 and for  D 0.3 Eq. (2.7) becomes
 2
h
c D 3.615E . 2.9
b
For a short plate (say, a/b < 1) one can see from Figure 2.4 that m D 1. Hence
Eq. (2.8) becomes kc D b/a2 [1 C a/b2 ]2 . If a/b − 1 then kc D b/a2 and
Eq. (2.7) becomes
 2
2 E h
c D 2
. 2.10
121    a
Notice that if we consider
p the plate to behave as a simply supported column with
L D a and  D h/ 12 then Eq. (2.5) yields c D 2 E/12 h/a2 . A comparison
of the two expressions indicates that if one replaces in the column equation E
by E/1  2  one obtains Eq. (2.10), the so-called wide column formula. This
difference is due to support that the strips, which make up the wide column, give

Figure 2.4 Compressive buckling coefficients for simply supported plates


20 The Concepts of Elastic Stability

to each other. Or in other words, the restraint against anticlastic bending in the
plate causes biaxial stresses which result in the 1  2  term.
The buckling loads of uniform rectangular plates under constant normal edge
forces have been determined for various boundary conditions either by solving the
appropriate differential equations or by using the Rayleigh Ritz method. In these
simple cases the in-plane stress resultant forces equal the applied edge forces and
the buckling stress can be calculated from Eq. (2.7). However, the value of the
buckling coefficient kc depends upon the type of loading and the edge restraints.
Results from Figure 14 of [2.13] are shown in Figure 2.5 and give the values of kc
as a function of a/b for unaxial compression with various combinations of simply
supported, clamped and free edges.
Notice that kc is essentially independent of the restraint at the loaded edges
when a/b > 3. However, in these cases kc depends strongly upon the restraint of
the unloaded edges. The buckling coefficient kc is nearly constant for long plates
a/b > 3, and as can be seen from Eq. (2.9) c does not depend on “a” and is
inversely proportional to b2 . This is contrary to the behavior of the column or the
wide column (when a/b2 − 1) where as can be seen from Eqs. (2.5) and (2.10)
the length rather than the width is the critical dimension and the important restraint
conditions are at the loaded rather than at the unloaded edges.
It is naturally unrealistic to assume that the plate is perfectly flat. Using a large
deflection theory and an initial imperfection of the form
x y
w0 x, y D W11 sin  sin  . 2.11
a b

Figure 2.5 Compressive buckling coefficients for rectangular plates with various boundary
conditions (from [2.13])
Physical Concepts Types of Observed Behavior and Their Meaning 21

Figure 2.6 Postbuckling stress distributions for plates with uniformly displaced loaded edges
(from [2.14])

Coan [2.14] obtained solutions for the buckling and the postbuckling behavior of
rectangular plates with uniformly displaced loaded edges and either undistorted or
stress-free unloaded edges. As can be seen from Figure 2.6a in the postbuckling
region the axial compressive stress x is no longer uniformly distributed over the
loaded edges as it is before buckling occurs. Instead it has a maximum value at the
simply supported unloaded edges that are held straight. Of considerable importance
are the in-plane stresses y that arise in the postbuckling region. Notice that in the
central region of the plate the y stresses are tensile in character and thus they
stiffen the plate considerably against further lateral deflection. These membrane
stresses together with the fact that the unloaded edges are restrained against out-
of-plane deflection explain why the plate, unlike the column (where there are no
such middle surface forces), can carry axial loads that are much higher than the
buckling load. Notice further that there are no resultant forces in the y-direction
thus the unloaded edges are free to move uniformly, that is v D constant. On the
other hand, as can be seen from Figure 2.6b if the unloaded edges of a plate
are stress free then a contraction occurs at the central region. The absence of
membrane forces in the y-direction accounts for the fact that such a plate carries
smaller postbuckling loads than those of a plate with straight unloaded edges.
The bending (out-of-plane) deformation at the center of perfect and imper-
fect square plates subjected to uniform end-shortening are shown in Figure 2.7.
Comparing these curves with the corresponding column curves of Figure 2.2 one
sees that, unlike for columns, for plates sizeable postbuckling stresses are possible.
Notice that following buckling, the stiffness of the plate decreases; however, failure
occurs only when the axial stress at the unloaded edges reaches the yield stress of
the material used. The buckling and postbuckling behavior of plates subjected to
shear loading is discussed in Chapter 8, Sub-section 8.4.1.

2.1.3 Instability of Columns with Compound Cross-Sections

The different failure modes that can occur with a thin-walled column of varying
length can best be illustrated by considering the lipped channel section shown
22 The Concepts of Elastic Stability

Figure 2.7 Nonlinear behavior of perfect and imperfect plates (from [2.15])

Figure 2.8 A lipped channel section

in Figure 2.8. Assuming that the lateral deflection of the cross-section from the
line of action of the compressive load varies sinusoidally along the length of the
column, then sufficiently long columns will buckle in global or overall modes,
where the half wavelength  of the sinusoidal buckle is equal to the length of a
simply-supported column.
Considering Figure. 2.9 one sees that, depending on the value of the slenderness
parameter /b global buckling can take the form of a flexural mode (Euler mode,
/b ½ 50), in which the cross-section translates but does not rotate, or the form
of a torsional mode (/b D 10), in which the cross-section rotates but does not
translate. There is also a third global mode, called the torsional-flexural mode, in
which the section, as shown in Figure 2.10, both rotates and translates.
In most applications of thin-walled open sections, there exists at least one
axis of symmetry, as illustrated in Figure 2.8. For torsional-flexural instability
of such cases, for example in [2.2], the following characteristic equation has been
derived
P  Pz fI0 /AP  Py P  P   P2 y02 g D 0 2.12
Physical Concepts Types of Observed Behavior and Their Meaning 23

Figure 2.9 Buckling behavior of thin-walled columns (from [2.16])

where
 
2 EIy 2 EIz 2 E
Py D ; Pz D ; P D A/I0  C GJ 2.13
L2 L2 L2

and
I0 D Iy C Iz C Ay02 C z02 
 D Wagner Torsion-Bending Constant
J D Torsional Constant
y0 , z0 D coordinates of the Shear Center S.C.

If Pz is the smallest of the three roots of Eq. (2.12), the column will buckle in
pure bending. Otherwise, the buckling will be combined bending and twisting.
If the cross-section has two axes of symmetry, y0 D z0 D 0 and Eq. (2.12)
simplifies to the form

P  Pz P  Py P  P  D 0. 2.14

In this case the three roots are Py , Pz and P and the column will buckle in pure
bending or pure twisting, depending on which of the three roots is the smallest.
24 The Concepts of Elastic Stability

Figure 2.10 Torsional-flexural buckling mode of a thin-walled column


(from Ref. 16.157, Volume 2)

Also illustrated in Figure 2.9 are cases of local instability in which the cross-
section distorts without translation or rotation. Thus, when the slenderness ratio
/b D 0.75 one gets a local buckling mode, where all the junctions between the
plate elements remain straight while the centers of the plates deflect out-of-plane
as shown. Interestingly enough, when the slenderness ratio /b D 4.0 one gets
another form of local buckling, where only some of the junctions between the
plate elements remain straight. This type of buckling is called flange buckling or
distortional buckling.
Test on short, thin-walled columns show that often, after local buckling has
occurred, the columns still have the ability to carry a greater load before general
failure takes place. Further, it appears that in cases where local buckling occurs
at relatively low stress levels, the stresses at general failure (or crippling) will be
Physical Concepts Types of Observed Behavior and Their Meaning 25

Figure 2.11 Stress distribution after local buckling

noticeably higher. On the other hand, if local buckling takes place at relatively high
mean stress levels (say, at 0.7 cy ) then the buckling and the crippling stresses are
practically the same. Figure 2.11 displays the stress distribution on a thin-walled
cross-section after local buckling has occurred but prior to crippling or failure.
Bending deflections become large after the flanges buckle, and crippling occurs
when the stresses at the supported (essentially straight) edges of the flanges reach
the compressive yield stress cy .
The nonlinear behavior associated with large displacements and plasticity has
prevented the development of a satisfactory analytical solution for the crippling
stress crip of arbitrary thin-walled cross-sections. Hence semi-empirical formulas
are used which are discussed in Chapter 6.

2.1.4 Effect of Modal Coupling

The application of structural optimization techniques has resulted in an increased


use of thin-walled compression members in modern vehicle design like aircraft,
ships, railway, trucks and other applications such as off-shore structures etc. The
initial idea that optimum design requires the equality of the local buckling load
P and the Euler buckling load PE has turned out to be incorrect. Several authors
(for example [2.17] and [2.18]) have shown that when simultaneous or nearly
simultaneous buckling modes do exist nonlinear coupling phenomena can result
in a compound mode of failure whereby explosive like collapse of thin-walled
columns may occur.
To illustrate the importance of modal coupling Van der Neut’s analysis [2.17]
of an idealized thin-walled compression member will be used. As can be seen
from Figure 2.12 the idealized built-up column consists of two axial load carrying
flanges of width b and thickness h. The flanges are held a distance 2c apart by
webs, which are rigid in shear and laterally (normal to mid-plane of flange) but
have no stiffness in the axial direction. The flange plates are assumed to be simply
26 The Concepts of Elastic Stability

Figure 2.12 Van der Neut’s idealized column (from [2.17])

supported along their longitudinal axes by the webs. The length of the column is
L, where h − b − L and both ends are assumed to be pin-jointed.
Obviously, there exist two important critical loads for this built-up column. If
the flange plates remain unbuckled, the column will become unstable at the overall
Euler buckling load
EI
PE D  2 2 . 2.15
L
Conversely, if the column axis remains straight, the simply-supported flange plates
become unstable at the local buckling load
2
P D 2kc D, 2.16
b
where for simply supported edges kc D 4 and D D Eh3 /121  2  is the bending
stiffness of the flange plate.
Van der Neut assumed that the column has failed once its centroidal axis first
bends at P D Pb , the bending buckling load. This appears to be a reasonable
assumption because although an Euler column exhibits a stable symmetric bifur-
cation point its postbuckling strength is very limited. One can actually distinguish
three separate cases.
If PE < P , the column fails by simple Euler buckling at Pb D PE .
Alternatively,
If P < PE , the simply supported flanges will buckle at P D P . This stable
symmetric buckling does not, however, exhaust the load carrying capacity of the
flanges (modelled as simply supported plates), which can carry an appreciable
axial load in the initial postbuckling region with an effective (reduced) Young’s
modulus of
E where
¾ 1/2.45 [2.17]. Ultimately the column fails in overall
Euler buckling at Pb D
PE . Of course, this approach assumes that the effective
stiffness
E is “smeared out” over the whole length of the column.
Finally,
We have the domain of compound failure at Pb D PE D P , where overall buck-
ling and local buckling occur simultaneously.
These three separate domains of buckling behavior are shown in Figure 2.13.
Van der Neut has shown by an Engesser Karman double modulus analysis [2.11]
that the column is in neutral equilibrium when
2
EI
P D  2 . 2.17
1 C
L02
Physical Concepts Types of Observed Behavior and Their Meaning 27

Figure 2.13 Column buckling loads (Pb vs L)

Further he demonstrated that for L2 < L < L0 the equilibrium at P is stable,


whereas for L0 < L < L1 the equilibrium at P is unstable and collapse will occur
explosively. Notice that by equating

Pb D 
PE LDL2 D P LDL0 D PE LDL1 2.18

one can derive the following expressions for L2 and L0 :


 
2
1/2
L0 D L1 D 0.761L1 ; L2 D
1/2 L1 D 0.639L1 . 2.19
1C

To investigate the effect of imperfect flanges it is convenient to replot the


results of Figure 2.13 in the form shown in Figure 2.14. Notice that in the range
1 < PE /P < 1.725, a range that is used frequently in aerospace applications, the
perfect column collapses explosively.
Assuming that the two flanges have the same initial imperfection in the form of
the buckling mode pertaining to the lowest local buckling load,

Figure 2.14 Column buckling loads (Pb /P vs PE /P )


28 The Concepts of Elastic Stability

x y
w D ˛h sin  cos  2.20
b b
where ˛ is the flange waviness parameter, Van der Neut obtained the relation
between the axial load in the imperfect plate strip and the corresponding end-
shortening via a Rayleigh Ritz type approximate solution of the nonlinear plate
equations. Since the two flanges have identical imperfections, the column axis will
remain straight under the load P until the overall bending load Pb is reached.
Using PE as a measure of the column length, Van der Neut [2.17] obtained the
buckling curves of columns with imperfect flanges shown in Figure 2.15. The
broken line for ˛ D 0 is the limit of the smooth curves ˛ > 0. Notice that initial
waviness of the flanges reduces the column failure load Pb considerably when
the ratio PE /P is close to 1.0. This severe imperfection sensitivity is due to the
modal interaction between the two buckling modes corresponding to PE and P ,
respectively. Notice that the magnitude of the imperfection sensitivity decays as the
ratio PE /P changes from unity. It is interesting that all imperfection curves pass
through a single point PE /P D 2.0. It appears that Pb < P when PE /P < 2.0.
For PE /P > 2 the initial waviness appears to be beneficial.
This example illustrates very well the fact that in general unconstrained struc-
tural optimization, whereby here the Euler buckling load of the compound column
and the local buckling load of the flanges occur simultaneously, may lead to
increasingly severe instabilities with pronounced imperfection sensitivity due to the
nonlinear coupling action of the failure modes involved. Mode interaction effects
in columns and plates will be further discussed in Chapters 6 and 4, Volume 2.

2.1.5 Buckling of Frames

Frames, both planar (a two-dimensional frame that is constrained to deform only


in the plane of the frame) and spatial, are structural members that are frequently

Figure 2.15 Buckling loads of columns with imperfect flanges [2.17]


Physical Concepts Types of Observed Behavior and Their Meaning 29

used in many different engineering applications. Thus it is not surprising that


their structural behavior has been studied widely, especially in civil engineering.
Recently, the increasing slenderness resulting from improved design procedures
has emphasized the stability problems of frames that may arise.
Considering a simple framework, if the joints are pinned the frame will become
unstable when the Euler load is first attained in one of the members, which will
then be the only one to exhibit significant distortion. If, however, the members are
rigidly connected at the loaded joint, then the load necessary to cause instability
in one member is enhanced by the other member(s) which restrain the rotation at
the joint, and therefore affect the stiffness of the compressed member.
Thus each member of a frame with rigid joints can be thought of as a bar with
elastically restrained ends. Hence it is often possible to analyze the stability of a
given framework based on the solution of the following model problem. Consider
the column of length L with a constant stiffness EI and spring supported at both
ends shown in Figure 2.16.
The critical value of the axial compressive load P can be obtained by solving the
following 2-point boundary value problem
wiv C k 2 w00 D 0 for 0  ð  L 2.22
EIw000 C Pw0 D k1 w at x D 0 2.23a
EIw00 D ˛1 w0
EIw000 C Pw0 D k2 w at x D L 2.23b
EIw00 D ˛2 w0
where  0 D d/dx, k 2 D P/EI, k1 and k2 are linear spring constants, while ˛1 and
˛2 are torsional spring constants.
It has been shown in [2.2] and [2.4], for example, that the characteristic equation
of this problem can be written as
1  1  2  1 2 2  1 2  2 2 C 1 1 2 4 C 2 1 2 4  sin 
C 2 C 1 2 C 2 2  1 1 4  1 2 4  2 1 4  2 2 4  cos   2 D 0
2.24
where
P EI EI EI EI
 D kL, k D , 1 D , 2 D , 1 D 3
, 2 D
EI ˛1 L ˛2 L k1 L k2 I3

Figure 2.16 Elastically supported column


30 The Concepts of Elastic Stability

are nondimensional parameters, with 1 and 2 defining the rotational elastic


restraints and 1 and 2 the lateral ones.
To illustrate the use of this equation the stability behavior of the portal frame
shown in Figure 2.17 will be investigated. In this case one is interested in finding
the smallest possible load Pc which will cause the frame to buckle. To accomplish
this, one must consider all possible modes of buckling, compute for each mode the
corresponding buckling load and find through comparison the critical buckling load
Pc . The different buckling modes are shown in Figure 2.18. Notice that there is no
possibility of a sway buckling mode when the horizontal bar buckles symmetrically.
At first, the rotational elastic restraint provided to the vertical bars by the hori-
zontal bar must be calculated. This can be done by considering a simply supported
beam subjected to end couples M2 .

Figure 2.17 Clamped portal frame

Figure 2.18 Buckling models for the clamped portal frame


Physical Concepts Types of Observed Behavior and Their Meaning 31

From beam theory the end rotation 2 is found to be for the


M2 Lh M2 2EIh
symmetric case: 2 D hence ˛2 D D
2EIh 2 Lh
and the rotational elastic restraint provided by the horizontal bar is
   
1 L EI
2 D .
2 EI h L v
Similarly for the
M 2 Lh M2 6EIh
antisymmetric case: 2 D hence ˛2 D D
6EIh 2 Lh
and the rotational elastic restraint provided by the horizontal bar is
   
1 L EI
2 D .
6 EI h L v
Notice that if EI/Lh D EI/Lv then for the
1
symmetric case 2 D 2 2.25a
1
antisymmetric case 2 D 6 2.25b
If side motion is suppressed (no sway buckling) then one must solve the buckling
problem shown in Figure 2.19.
The characteristic equation (Eq. (2.24)) becomes then for the
symmetric case F D  sin  C 4 C 2  cos   4 D 0 2.26a
antisymmetric case F D 5 sin  C 12 C 2  cos   12 D 0. 2.26b
The lowest roots of these transcendental equations can easily be found numerically
via Newton’s Method, yielding for the
PL EI 2 EI
symmetric case D D 5.0182; P D 25.1822 2 D 2.5515 2
EI L L
PL EI 2 EI
antisymmetric case  D D 5.5272; P D 30.5498 2 D 3.0953 2 .
El L L

Figure 2.19 Buckling model of the vertical bar


32 The Concepts of Elastic Stability

Figure 2.20 Sway buckling model of the vertical bar

To obtain the characteristic equation for the sway buckling case shown in
Figure 2.18c one must solve the buckling problem displayed in Figure 2.20. In
this case Eq. (2.24) reduces to

F D 6 sin  C  cos  D 0. 2.27

When deriving this equation one must divide Eq. (2.24) by 2 and then take the
limit as 2 ! 1. The lowest root of Eq. (2.27) is

PL EI 2 EI
D D 2,7165; P D 7.3792 D 0.7477 .
EI L2 L2
Comparison of the three buckling loads computed indicates that the characteristic
equation for the critical load of the clamped portal frame is given by Eq. (2.27).
Therefore, as the load P is increased slowly from zero, the frame will sway buckle
when P reaches the value of the lowest root that satisfies Eq. (2.27).
For other worked out examples of frame buckling the interested reader should
consult [2.1], [2.2] and [2.4].

2.1.6 Lateral Buckling of Beams

In general, structural members whose primary function is the transfer of loads


which act perpendicular to their longitudinal axis by bending, are called beams.
Often, by proper design one can ensure that the applied loading acts through the
shear center of the beam cross-section, thereby eliminating any twisting action.
Moreover, in many applications the structure is so arranged that the resulting
bending may be regarded as taking place effectively in the plane of symmetry of the
beam. In such cases major axis bending can be considered as the principal design
variable. Due to this fact the type of cross-section selected is usually relatively
weak in both minor axis bending and twisting. These slender beams loaded in a
plane of symmetry may buckle laterally as shown in Figure. 2.21.
An approximate analysis of lateral instability in terms of a thin-walled elastic
beam theory has been presented in [2.1] and [2.2] (as well as in many other texts).
Physical Concepts Types of Observed Behavior and Their Meaning 33

Figure 2.21 Failure of beam by lateral buckling (courtesy of Prof. Yuhshi Fukumoto)

Consider the beam with two planes of symmetry shown in Figure 2.22. This beam
is loaded only in the xz-plane. That is, it is subjected to transverse forces acting
in the z-direction and to bending couples M0 . Notice that the transverse forces
cause additional bending moment My and when the beam is displaced laterally
the transverse loading also causes a relatively small twisting moment Mx . As
long as the load on the beam remains below the critical value, the beam will
be stable. However, as the load is increased a critical value is reached when a
slightly deflected and twisted form of equilibrium becomes possible. The initial
plane configuration of the beam is now unstable, and the lowest load at which this
deflected condition occurs is called the critical load of the beam. The deflection of
the beam is described by the displacement components v and w of the centroid of
the cross-section and by the angle of rotation  of the cross-section. The axes ,

,  represent coordinates of the deformed configuration and the angles of rotation


34 The Concepts of Elastic Stability

Figure 2.22 Beam in undeformed and deformed configuration symbols and sign convention

v0 and w0 and the angle of twist  are assumed to be small. The sign conventions
for positive moments acting at section AB on the portion of the beam to the left
of the section are shown in Figure 2.22.
The equations of bending and twisting of the buckled beam have been derived,
for example in [2.2], as
EIz v00 D M 2.28a
EIy w00 D M
2.28b
E000  GJ0 D M 2.28c

where  0 D d/dx and it is assumed that  is sufficiently small so that the curvatures
and flexural rigidities in the
 and  planes may be replaced by their values in
the yx and zx planes.
In these equations Iy and Iz are the principal moments of inertia of the cross-
section about the y and z axes, respectively. Similarly, the quantities M
and M
represent the bending moments about
and  the axes. Further, GJ is the torsional
rigidity and E is the warping rigidity. Notice that for small angles of rotation and
a small angle of twist, the bending and twisting moments acting on a deformed
cross-section parallel to the deformed axes ,
,  can be expressed in terms of
Mx and My , the twisting and bending moment acting on the same cross-section
parallel to the undeformed axes x and y. If one neglects products of small angles
and small twisting moments then these expressions are (see also Figure 2.22),
M D Mx C My v0 2.29a
M
D M y 2.29b
M D My . 2.29c

As an example of beams with doubly symmetric cross-sections let us consider


the case of the simply supported I-beam of Figure 2.22 subjected to the end couples
M0 only. The moments Mx and My acting on section AB on the portion of the
Physical Concepts Types of Observed Behavior and Their Meaning 35

beam to the left of the section are


Mx D 0; My D M0 .
From Eq. (2.29) one obtains M D M0 v0 , M
D M0 , M D M0  and then the
governing Eq. (2.28) can be written as
EIz v00  M0  D 0 2.30a
00
Ely w D M0 2.30b
E000  GJ0  M0 v0 D 0. 2.30c
Notice that Eqs. 2.30a and 2.30c must be solved simultaneously. By differentiating
the last equation once, one can then eliminate the v00 term with the help of the first
equation yielding
M2
Eiv  GJ00  0  D 0. 2.31
EIz
The general solution of this constant coefficient differential equation may be written
(see [2.2])
 D C1 e˛x C C2 e˛x C C3 cos ˇx C C4 sin ˇx 2.32
where
 
1/2 1/2
 GJ 2 2 
GJ M
˛D C C 2 0 2.33a
 2E 2E E Iz 
 
1/2 1/2
 GJ  
GJ 2 M20
ˇD  C C 2 2.33b
 2E 2E E Iz 

and the constants C1 , C2 , C3 , C4 must be evaluated from the boundary conditions.


For a simply supported beam where the ends are free to warp, x D 0 at x D 0, L.
It has been shown in [2.2] that this implies at 00 D 0 at x D 0, L. If, in addition, the
ends are prevented from rotating about the longitudinal axis, then  D 0 at the ends.
Thus the boundary conditions for the simply supported I-beam of Figure 2.22 are
 D 00 D 0 at x D 0, L. 2.34
The characteristic equation of the problem then becomes
sinh ˛L sin ˇL˛2 C ˇ2 2 D 0. 2.35
But ˛, ˇ and sinh ˛L are nonzero quantities. Thus the characteristic equation
reduces to
sin ˇL D 0 2.36
with eigenvalues ˇL D m, where m is a positive integer. Substitution and
regrouping yields
 
1/2
 m 2 GJ GJ 2 M20
D C C 2 . 2.37
L 2E 2E E Iz
36 The Concepts of Elastic Stability

It can be seen by inspection that the critical (the smallest) value of M0 occurs
when m D 1. Hence the critical value of M0 is
 1/2
4 EIz E 2 EIz GJ
Mc D C 2.38
L4 L2

with the corresponding buckling mode (see also [2.2])


x
 D C4 sin  . 2.39
L
Notice that the maximum angle of twist occurs at the midspan. Also the magnitude
of the critical moment given by Eq. (2.38) does not depend on the flexural rigidity
EIy of the beam in the vertical plane. This is due to the fact that Eq. (2.28b) is
uncoupled from the other two equations. The assumption that the deflections in the
vertical plane are small is justifiable when the flexural rigidity EIy in the vertical
plane is very much greater than the lateral rigidity EIz .
The interested reader may consult [2.1] [2.3] of [2.19] for an extensive list
of worked out examples and an exhaustive discussion of the different problems
associated with the phenomena of lateral buckling of beam-like structures.

2.1.7 Instability Due to Patch Loading

Another type of initially localized lateral instability may occur if the introduction of
the transverse loading into the thin-walled web of a beam is not properly designed.
Consider the thin-walled structural member shown in Figure 2.23 subjected to
a partial in-plane compressive edge loading. The roller supports could be adjusted
so as to ensure that the load was applied to both webs. This test set-up can be
considered representative of a steel flooring system. In a test program carried out by
K.C. Rockey and his co-workers [2.20] it was found that if the depth-to-thickness
ratio of the web is sufficiently high, then the web will buckle before it fails, where
failure of the test panels was defined by the deformation of a localized yield curve
under the patch load. This program represents an excellent example of the use
of a combination of numerical and experimental methods to arrive at a relatively
simple semi-empirical design formula for a pressing technical problem.

Figure 2.23 Sketch of the test set-up (from [2.20])


Physical Concepts Types of Observed Behavior and Their Meaning 37

Using the finite element method Rockey and Bagchi [2.21] derived the following
formula for the critical value of the compressive patch load Pc which will cause
buckling of a rectangular plate
Pc 2 D
DK 2 2.40
bh d h
where D D Eh3 /121  2  is the flexural rigidity of the plate. The nondimensional
buckling coefficient K is shown in Figure 2.24 for the case where the patch load
on one longitudinal edge is supported by shear forces on the two transverse edges.
In [2.20] and [2.21] interaction curves are also presented for the cases where an in-
plane bending moment or an in-plane shear stress acts in addition to the stress field
set up by the patch loading. Loads were applied to the test specimens in small incre-
ments in the elastic range, and in even smaller increments after yielding has begun.
The lateral deflection of the web was recorded using a specially designed recording
device consisting of seven movable linear-displacement transducers, which could
be adjusted to any position desired. Figure 2.25 displays the variation of the lateral

Figure 2.24 Compressive buckling coefficients for different patch loadings (from [2.20])

Figure 2.25 Lateral displacements at central section under various loads (from [2.20])
38 The Concepts of Elastic Stability

deflection across the central section of the panel for a typical test. Notice that the
largest deformations are located in the upper half of the panel adjacent to the patch
load. In the inelastic range, all plastic flow was allowed to take place at each load
increment before any lateral deflection readings were taken.
In all tests, failure occurred by the formation of a local yield curve as shown
in Figure 2.26. Notice that the yield curve corresponds closely to a segment of a
circle with a width equal to that of the patch load.
The primary purpose of the test program was to determine the ultimate load
carrying capacity of the webs of a sheet steel flooring system. Rockey et al. [2.20]

Figure 2.26 Panel after failure showing yield curve (from [2.20])

Figure 2.27 Variation of Pu /Pc vs d/h for a square web b/d D 1


Physical Concepts Types of Observed Behavior and Their Meaning 39

found that for d/h  250 there exists a linear relationship between the ultimate
failure load Pu and the theoretical buckling load Pc given by Eq. (2.40). By curve
fitting the experimental results shown in Figure 2.27 they obtained the following
semi-empirical formula
Pu  c d
D 103 4.5 C 6.4 . 2.41
Pc b h
Notice that for d/h > 250 the relationship between the ultimate failure load Pu
and the theoretical buckling load Pc becomes nonlinear.

2.1.8 Buckling of Beam-Columns

Slender beams subjected to both axial compression and bending are called beam-
columns. As an example of beam column analysis let us consider the case of
a simply supported beam with a doubly symmetrical cross-section subjected to
compressive loads P and end moments M0 as shown in Figure 2.28. The loads P
are applied at the centroid of the cross-section and Iy × Iz . Thus the member can
bend in the xz-plane or bend and twist out of that same plane.
The analysis follows closely the one described earlier in section 2.1.6 for the
lateral buckling of beams. Using the symbols and the sign convention defined in
Figure 2.22, the equations of bending and twisting of the buckled beam-column
have been given earlier by Eqs. 2.28. Differentiating the last equation with respect
to x yields
EIz v00 D M 2.42a
00
EIy w D M
2.42b
Eiv  GJ00 D M0 2.42c
where  0 D d/dx and M , M
, M are the twisting and bending moments acting
at section AB on the portion of the beam to the left of the section and parallel
to the deformed axes ,
, . The additional bending moments due to the axial
load P acting at the deformed cross section are M
D Pw and M D Pv. The
rate of change of the twisting moment for a doubly symmetrical cross-section
(where y0 D z0 D 0) has been shown in [2.2] to be M0 D I0 /AP00 , where I0 is
the polar moment of inertia. From Eqs. 2.29 the corresponding quantities due to
the end moments M0 are M0 D M0 v00 , M
D M0 and M D M0 . Introducing

Figure 2.28 Beam-column in undeformed configuration


40 The Concepts of Elastic Stability

these quantities into Eq. (4.42) one obtains after some regrouping the following
governing equations for the beam-column,
EIz v00 C Pv  M0  D 0 (2.43a)
EIy w00 C Pw D M0 (2.43b)
I0 00
Eiv  GJ00 C P  M0 v00 D 0. (2.43c)
A
Notice that the second equation is uncoupled from the other two equations. It
governs the bending in the xz-planes and because of the presence of the term Pw
the bending is nonlinear. The general solution of Eq. (2.43b) is
M0
w D C1 sin kx C C2 cos kx C 2.44
P
where k 2 D P/EIy and the constants C1 and C2 can be evaluated from the specified
boundary conditions. For simply supported ends w D 0 at x D 0, L and the solution
becomes
M0
wD fsin kL  sin kx  sin kL  xg. 2.45
P sin kL
Notice that as the load P approaches the value Py D 2 EIy /L 2 , the eigenvalue
for loss of stability by bending in the xz-plane, the factor sin kL in Eq. (2.45)
approaches zero and the magnitude of the displacement w approaches infinity.
For values of P smaller than Py , the allowable values of P and M0 are limited
by the strength of the beam-column material.
To investigate the lateral-torsional instability of the beam-column Eqs. (2.43a)
and (2.43c) must be solved simultaneously. For simply supported boundary condi-
tions at x D 0, L
 D 00 D 0 and v D v00 D 0 2.46

solutions of the form


x x
 D A1 sin  ; v D A2 sin  2.47
L L
satisfy the specified boundary conditions identically, and upon substitution into
Eqs. (2.43a) and (2.43c) for nontrivial solution yield the following characteristic
equation
I0 /AP  PPz  P  M20 D 0 2.48

where P and Pz are defined by Eq. (2.13).


For given values of M0 (or for a beam-column subjected to eccentric compressive
loads P which cause end moments M0 D Pe), the critical load Pc is the smaller of
the two roots of Eq. (2.48).
Conversely, for given values of P, the critical end moment Mc is

Mc D fP  PPz  PI0 /Ag1/2 . 2.49


Physical Concepts Types of Observed Behavior and Their Meaning 41

For additional examples dealing with the behavior of beam-columns under


different lateral loads the interested reader may consult [2.1], [2.3] or [2.22].

2.1.9 Buckling of Rings and Arches

The stability behavior of thin rings and arches exhibit several features that are not
encountered when one is dealing with straight columns and flat plates. Thus, for
instance, the perfect thin ring under lateral loading undergoes a nonzero lateral
displacement w prior to loss of stability, whereas for perfect columns and plates
w D 0 for the unbuckled state.
For a complete historical sketch of investigations dealing with the buckling of
thin circular rings the interested reader should consult [2.1]. The following analysis
based on a small strain, moderate rotation theory is due to Brush and Almroth [2].
They based their ring-bending theory on the simplifying assumption that normals to
the undeformed centroidal surface remain straight, normal and inextensional during
deformation. Thus the extensional strain of a circumferential line element located
at a distance z from the centroidal surface (see Figure 2.29) can be expressed in
terms of centroidal-surface relations as follows;

εN D ε C z 2.50

where
1 0 1 1 1 0
εD v C w C ˇ2 ; ˇD v  w0 ; D ˇ 2.51
R 2 R R
and  0 D d/d .
It has been shown by Bodner [2.23] that a ring subjected to a uniform external
pressure is a conservative system. Thus the governing equilibrium equations can
be conveniently derived by the stationary potential energy criterion, which states
that a conservative system is in equilibrium if its potential energy is stationary
(see for example [2.24] or [2.25]). Further, for a conservative system the change
in potential energy of the applied loads as the structure deforms is the negative
of the work done by the loads during the deformation. This yields the following

Figure 2.29 Circular ring subjected to uniform external pressure


42 The Concepts of Elastic Stability

variational problem
 2
υ D υUm C Ub C p  D υ F , v, v0 , w, w0 , w00 R d D 0 2.52
0
where

EA 2 2
Um D membrane energy D ε R d 2.53
2 0

EI 2 2
Ub D bending energy D  R d 2.54
2 0
p D potential of the applied load. 2.55
When working with external pressure it is customary to distinguish between
fluid pressure loading and centrally directed pressure loading. In the first case the
pressure at each point on the ring surface remains normal to the surface as the
ring deforms (the so-called “live load”). In the second case the pressure remains
centrally directed at each point on the ring surface (the so-called “dead load”). As
a consequence of these assumptions the potential energy of the applied load differs
considerably for these two cases, yielding for
 2  
1 2
Fluid pressure loading: p D q wC v  vw C v w C w  R d
0 0 2
0 2R
2.55a
 2  
1 2
Centrally directed pressure loading: p D q wC v R d . 2.55b
0 2R
The condition that υ D 0 implies that the integrand in Eq. (2.52) must satisfy the
corresponding Euler equations of the calculus of variation. The Euler equations for
an integrand of the form as indicated in Eq. (2.52) are
∂F d ∂F
 D0 (2.56a)
∂v d ∂v0
∂F d ∂F d2 ∂F
 C D 0. (2.56b)
∂w d ∂w0 d 2 ∂w00
Calculating the required partial derivatives for fluid pressure loading, substitution
and regrouping yields
   0      
v0 C w 1 v  w0 2 1 v  w0 00 v0 C w 1 v  w 0 2
C C  C
R 2 R AR2 R R 2 R
   
v  w0 qR v  w0
ð  D0 2.57a
R EA R
       0  

1 v  w0 000 v0 C w 1 v  w 0 2 v C w 1 v  w0 2
 C  C
AR2 R R 2 R R 2 R
 0  
v  w0 qR v0 C w qR
ð  D . 2.57b
R EA R EA
Physical Concepts Types of Observed Behavior and Their Meaning 43

Under axisymmetric loading circular equilibrium configurations of the ring exist


for all values of the applied load q < qc . The critical load qc is the smallest load
for which the ring may be maintained in equilibrium in an adjacent noncircular
configuration.
To investigate the possible existence of adjacent equilibrium position one gives
small increments to the displacement variables
v D v0 C vO
2.58
w D w0 C wO
where v0 , w0 are the circular prebuckling solutions and vO , wO are small noncircular
perturbations at buckling. Direct substitution into Eqs. (2.57a) and (2.57b) and
deletion of squares and higher order products of the perturbation quantities yields
a set of nonlinear governing equations for the prebuckling quantities v0 , w0 and
a set of linearized stability equations governing the perturbation quantities vO , w.O
Notice that the nonlinear equations governing the prebuckling state variables v0 ,
w0 are identical to Eq. (2.57).
Recalling that for a circular prebuckling state v0 and all derivatives of v0 and w0
are identically zero, Eq. (2.57a) is identically satisfied and Eq. (2.57b) reduces to
w0 qR  w0 
C 1C D 0. 2.59
R EA R
For thin rings w0 /R − 1, thus w0 D qR2 /EA. Notice that for the prebuckling
solution
qR2
v0 D 0, w0 D  2.60
EA
the linearized stability equations become
EAR2 Ov0 C w
O 0 C EIOv  wO 0 00 D 0 (2.61a)
EAR2 Ov0 C w
O  EIOv  wO 0 000 C qR3 wO 00 C w
O D0 (2.61b)
a set of homogeneous linear equations in vO and w. O For a complete ring the boundary
conditions simply require that vO , wO and their derivatives be periodic in . Thus a
solution of the form
vO D B cos n
2.62
wO D C sin n
where B, C are constants and n is a positive integer, satisfies the periodicity
condition and reduces the solution of the set of differential equations to a simple
matrix eigenvalue problem. The roots of the characteristic equation can be put into
the form
n2  1
qn D n D 1, 2, 3, . . . . 2.63
1 C I/AR2
For n D 1 the eigenvalue is q1 D 0. However, the corresponding eigenvector
vO D B cos
wO D B sin
44 The Concepts of Elastic Stability

represents a rigid body translation and not a noncircular buckling mode. The ring
is thought to be constrained against such translation and one considers only buck-
ling modes for which n is greater than unity. The smallest eigenvalue is seen to
correspond to n D 2 (ovalization of the ring) and since for thin rings I/AR2 − 1
its value is
EI
qc D 3 3 . 2.64
R
This result, obtained by Bresse in 1866 [2.26] and independently in 1884 by
Lévy [2.27] is considered to be the classical solution for a ring subjected to external
fluid pressure.
It may be of some interest to mention here that besides the fluid pressure loading,
which best represents the real load case of external pressure, and the centrally
directed pressure loading with a lowest eigenvalue of qc D 4.5EI/R3 [2.28], the
case of a thin ring loaded by external pressure, where the load remains parallel to
its original direction has also been solved. Though it is difficult to conceive of a
practical application for this last case, Singer and Babcock [2.29] have shown that
such a thin ring is unstable as a rigid body and will rotate under arbitrarily small
pressure that remains parallel to its original direction.
Equations (2.61) can also be used to investigate the buckling behavior of the
high circular arch under normal pressure loading shown in Figure 2.30. Notice that
in this case it is assumed that initially the arch is uniformly contracted so that a
fundamental state exists, which is identical to that of the complete ring. Then at
the instant of buckling the supports become immovable and the arch buckles in an
antisymmetric mode as shown in Figure 2.30. If the arch is simply supported at
both ends, then the boundary conditions at D š˛ are

EI 0 EA 0
wO D 0; O D
M Ov  wO 00  D 0; O D
N Ov C wO 00  D 0. 2.65
R2 R

Figure 2.30 Pinned circular arch submitted to uniform external pressure


Physical Concepts Types of Observed Behavior and Their Meaning 45

Notice that a solution of the form


n
vO D B cos
˛
2.66
n
wO D C sin
˛
where B, C are constants and n is a positive integer satisfies the above simply
supported boundary conditions identically and reduces the solution of the set of
differential equations (2.61) to a simple matrix eigenvalue problem. For this case
the roots of the characteristic equation can be put in the form

[n/˛2  1] EI
qn D n D 1, 2, 3, . . . . 2.67
1 C I/AR2 R3

The critical condition corresponds to n D 1. Thus for thin arches, where I/AR2 − 1
  
 2 EI
qc D 1 3. 2.68
˛ R
This solution was obtained by Hurlbrink in 1908 [2.30] and independently by
Timoshenko in 1910 [2.31].
Notice that when ˛ D  one obtains a complete ring and from Eq. (2.68) qc D 0.
As has been pointed out in [2.4], this unrealistic result is due to the fact that for
˛ D  one has a complete ring with a hinge, which is free to rotate as a rigid body
about this hinge for arbitrarily small pressure. That the continuous complete ring
corresponds to ˛ D /2 can be deduced also from the fact that then Eq. (2.68)
reduces to Eq. (2.64). Solutions for other boundary conditions and for arches of
other forms can be found in [2.1], [2.4] and [2.8].
Finally, it must be mentioned that the above results derived by assuming the
buckling modes given by Eq. (2.62) or Eq. (2.66) are not applicable to shallow
arches. The low arch problem will be treated in the next section.

2.1.10 Buckling of Shallow Arches

Transversely loaded shallow arches represent a class of widely used structural


elements. Their stability behavior differs from those of the preceding examples in
that the fundamental path is not identified with w D 0 (zero lateral displacement)
prior to loss of stability. On the contrary, the fundamental path is highly nonlinear
and depending on the value of the dimensionless rise parameter K it exhibits limit
point and bifurcation points in the load versus lateral displacement plane.
To illustrate these points the stability behavior of the clamped shallow arch under
“dead” pressure loading shown in Figure 2.31 will be investigated. The analysis is
based on the work by Schreyer and Masur [2.32] and of Kerr and Soifer [2.33]. To
derive the equilibrium equations the stationary potential energy criterion, described
in the previous section, will be used. For shallow arches (for arches where the
46 The Concepts of Elastic Stability

Figure 2.31 Clamped circular arch under uniform pressure

arch rise H is small compared to the arch span 2b) one assumes that the rotation
ˇ behaves very much like that of a straight beam, thus
1 0
ˇD w 2.69
R
Further, following Koiter’s work (see [2.34], it will be assumed that the potential
energy of the applied pressure may be represented by
 2
p D q wR d . 2.70
0

With the help of Eqs. 2.51 2.54 the Euler equations of the variational problem
can then be written as
N0 D 0 (2.71a)
M00  RN C Nw00  qR2 D 0 (2.71b)
where  0 D d/d and use has been made of the constitutive equations
N D EAε; M D EI. 2.72
Notice that Eq. (2.71a) can be integrated directly yielding
N D constant D N0 . 2.73
This result, that the in-plane force N is constant is very useful, for it can be used
to obtain an exact, closed form solution for the complete nonlinear buckling and
post-buckling problem of shallow arches. Notice that Eq. (2.71b) can be put into
the form
wiv C 2 w00 D 2 R[1  qO K/ ˛2 ] 2.74
where 2 D N0 R2 /EI, qO D qR2 h/EI and the dimensionless rise parameter K D
˛2 R/h. The general solution of this linear ordinary differential equation with
Physical Concepts Types of Observed Behavior and Their Meaning 47

constant coefficients is

w D A1 C A2 C A3 sin C A4 cos C R/2[1  qO K/ ˛2 ] 2 . 2.75

Recalling that for clamped arches the boundary conditions at D š˛ are

w D w0 D 0; vD0 2.76

then one can attempt to evaluate the four constants A1 , A2 , A3 , A4 from the first
two conditions. It is found that whereas A1 and A4 can be determined uniquely,
nonzero values of A2 and A3 are possible only if the characteristic equation

tan ˛ D ˛ 2.77

is satisfied.
Considering, at first, the case when the characteristic equation is not satis-
fied (that is, tan ˛ 6D ˛) then A2 D A3 D 0, and Eq. (2.75), the general solution,
becomes


2 cos  cos ˛ 1  2
w D hK[1  qO K/ ˛ ]  1 . 2.78
 ˛ sin ˛ 2  ˛2

Notice that this expression is even in , thus it represents the solution for the
symmetrical deformations. Further it should be noted that in this equation is as
yet an unknown parameter, which by definition is related to the constant in-plane
force N0 . Its value can be determined from Eq. (2.73) and the remaining boundary
conditions v D 0 at D š˛.
To achieve this one rewrites Eq. (2.73) with the help of Eqs. (2.51) and (2.69)
and then integrates the resulting equation from ˛ to C˛ yielding
 C˛  C˛  
N0 R
v d D
0
 0 2
 w  w  d . 2.79
˛ ˛ EA
Notice that the left-hand side vanishes identically because of the boundary
conditions v D 0 at D š˛. Substitution for w and w0 from Eq. (2.78) and then
performing the integrals one obtains
 
K2 5 2 2 2
qO 4   ˛  3 ˛ cot ˛   ˛ cot ˛
 ˛4 3
 
2K 2 2 2
 qO 2   ˛   ˛ cot ˛   ˛ cot ˛
 ˛2
 
1 2 1  ˛4 2 2
C   ˛  C  ˛ cot ˛   ˛ cot ˛ D 0. 2.80
3 3 K2
Equations (2.78) and (2.80) represent the exact solution for the symmetrical defor-
mations. For a given value of the nondimensional rise parameter K one can obtain
a numerical solution as follows. Initially, for a pre-selected set of ˛-values one
48 The Concepts of Elastic Stability

solves for the corresponding values of qO using Eq. (2.80). Next, for each pair of
˛ and qO values one calculates the corresponding radial displacement w0/h
from Eq. (2.78). The results for the first buckling mode using K D 10 are shown
in Figure 2.32. Notice that two limit points are obtained, an upper limit point at
qO u D 2.2681 2.81
and a lower limit point at
qO L D 0.4808. 2.82
Let us now consider the case when A2 and A3 are not identically equal to
zero. Earlier it was found that for such nonsymmetric deformation to occur the
characteristic equation tan ˛ D ˛ must be satisfied. Its roots are
Ł ˛ D 4.4934, 7.7253, . . . . 2.83
Notice that
NŁ0 R2
Ł ˛ D constant D 4.4934 D ˛ 2.84
EI
implies that NŁ0 is constant not only throughout the arch but it remains constant
throughout the nonsymmetrical deformations shown in Figure 2.32. N0 is equal
to its value at the bifurcation point and is denoted by NŁ0 . Thus the bifurcation
pressure is obtained by substituting the characteristic equation tan Ł ˛ D Ł ˛ into
Eq. (2.80) of the symmetrical equilibrium branch. This results in the following

Figure 2.32 Load-displacement curves for K D 10 (from [2.33])


Physical Concepts Types of Observed Behavior and Their Meaning 49

equation for qO b

Ł 2 Ł 4 Ł 2
6  ˛  ˛  ˛
qO b2  qO b C 1C D0 2.85
5 K 5K2 K2

which yields   
Ł
 ˛ 
2 Ł 2
5  ˛
qO b D 3š2 1 . 2.86
5K  4 K2 

From the discriminant of this equation it follows that for bifurcation to take place
p
5 Ł
K>  ˛. 2.87
2
Since the lowest value of Ł ˛ is 4.4934, it follows from Eq. (2.87) that when
K  5.024 the shallow arch deforms only symmetrically. Instability will then occur
at the upper limit point in what is known as snap-through buckling or oil-canning.
In the present case for K D 10 and Ł ˛ D 4.4934 one obtains from Eq. (2.86)
the following bifurcation pressures;

qO b D 1.9098 and qO b D 0.51131. 2.88

Notice that qO b D 1.9098 corresponds to the bifurcation point A, whereas qO b D


0.5131 corresponds to the bifurcation point B in Figure 2.32.
It can easily be shown that when the characteristic equation tan Ł ˛ D Ł ˛ is
satisfied then
A2 D A3 Ł cos Ł ˛. 2.89

The corresponding radial displacement becomes

w D ws  A3 [ Ł  cos Ł ˛  sin Ł ] 2.90

where ws is given by Eq. (2.78) with replaced by Ł . The only remaining


unknown constant A3 is determined from the boundary conditions v D 0 at D š˛
rewritten in the form given by Eq. (2.79). Substituting for w and w0 from Eq. (2.90)
and carrying out the integrals yields

1/2
hK 1 25 K
2
2K 1  Ł ˛2
A3 D š Ł 2 Oq  qO Ł 2  1C .
 ˛ cos Ł ˛ 3  Ł ˛4  ˛ 3 K2
2.91
Equations (2.90) and (2.91) constitute the exact solution for the nonsymmetrical
deformations.
Notice that for D 0, K D 10 and Ł ˛ D 4.4934 Eq. (2.90) reduces to the
equation of a straight line, it becomes
w0
 D 7.7775  3.8509Oq. 2.92
h
50 The Concepts of Elastic Stability

Schreyer and Masur [2.32] have carried out an extensive investigation as to the
character of the buckling behavior of shallow arches under dead pressure loading as
a function of the dimensionless rise parameter K. Their results can be summarized
as follows:
0 < K < 2.85 very shallow arch, no buckling occurs
2.85 < K < 5.02 symmetric limit point (or snap), buckling occurs at qO D qO u
(see Figure 2.32)
5.02 < K < 5.74 asymmetric bifurcation point occurs after the limit point
5.74 < K asymmetric bifurcation occurs at point A (see Figure 2.32)
before the limit point

For further results the interested reader may consult [2.1], [2.4] and [2.5].

2.1.11 Buckling of Circular Cylindrical Shells

Thin-walled shells are frequently used structural elements in such diverse applica-
tions as cooling towers, legs of offshore bore islands, aircraft fuselages or as the
main load carrying elements of aerospace launch vehicles. The popularity of shells
is due to the fact that they are very efficient load carrying structures. However,
unfortunately, often they are prone to “catastrophic” elastic instabilities. Thus a
thorough understanding of the stability behavior of thin-walled shells is a must
for all those who employ them. This was realized already in the last century, as
pointed out in the historical introduction to Chapter 9, Volume 2.
Circular cylindrical shells will be treated separately because their stability
equations are much simpler than those of shells of general shape, and thus can
be used very conveniently to illustrate the different types of instabilities that may
occur. In the present analysis the relatively simple Donnell type shell theory
will be employed. These equations give accurate results for cylindrical shells
whose displacement components in the deformed configuration are rapidly varying
functions of the circumferential coordinate. For the sign convention used see
Figure 2.33.
In the age of computerized shell stability analysis the interest in using the
Donnell type shell equations has practically disappeared. However, their relative
simplicity makes them ideally suited for rapid approximate analytical develop-
ments and hence also for the following introductory analytical examination of
shell stability.
The Donnell equations are based on the following middle-surface kinematic
relations
εx D u,x C 12 ˇx2 ˇx D w,x x D ˇx,x
w
εy D u,y C C 12 ˇy2 ˇy D w,y y D ˇy,y 2.93
R
xy D u,y C v,x C ˇx ˇy xy D 21 ˇx,y C ˇy,x 
Physical Concepts Types of Observed Behavior and Their Meaning 51

Figure 2.33 Circular cylindrical shell symbols and sign convention

Comparing the circumferential rotation ˇy with the one used for the ring problem
in Eq. (2.51) one sees that according to Donnell’s approximation the circumferen-
tial displacement component v is neglected relative to the gradient of the normal
displacement component in the circumferential direction w,y .
Employing the stationary potential energy criterion, the following set of
nonlinear governing equations are derived in [2.2] for isotropic circular cylindrical
shells.
Nx,x C Nxy,y D 0 (2.94a)
Nxy,x C Ny,y D 0 (2.94b)
1
Mx,xx C Mxy C Myx ,xy C My,yy  Ny C Nx w,xx
R
C 2Nxy w,xy C Ny w,yy D p. (2.94c)
Three equations in three variables, the displacements u, v, w may be obtained by
introduction of the isotropic constitutive equations
Nx D Cεx C εy  Mx D Dx C y 
Ny D Cεy C εx  My D Dy C x 
2.95
1   Mxy C Myx
Nxy D C xy D D1  xy
2 2
and the kinematic relations from Eq. (2.93) into Eq. (2.94). The extensional
and the bending stiffness parameters are, respectively, C D Eh/1  2  and
D D Eh3 /121  2 .
A simpler set of two equations in two variables w and f can be derived as
follows. Notice that if one defines an Airy stress function f such that
Nx D f,yy , Ny D f,xx , Nxy D f,xy 2.96
52 The Concepts of Elastic Stability

then the in-plane equilibrium equations (2.94a) and (2.94b) are identically satis-
fied. The remaining out-of-plane equilibrium equation (2.94c) and the compatibility
equation
1
εx,yy C εy,xx  xy,xy D w,2xy w,xx w,yy C w,xx 2.97
R
yield upon substitution and regrouping
1
Dr4 w C f,xx  f,yy w,xx  2f,xy w,xy C f,xx w,yy  D p (2.98a)
R
1
r4 f  Ehw,2xy w,xx w,yy C w,xx  D 0 (2.98b)
R
where
r4   D  ,xxxx C 2 ,xxyy C  ,yyyy . 2.98c

These equations were first presented by Donnell as three equations in 1933 [2.35].
When talking about buckling of thin-walled shells one must distinguish between
collapse at the maximum point of a load-deflection curve and bifurcation buckling,
the same types of behavior as encountered earlier by shallow arches. Thus if
one employs the general nonlinear analysis governed by Eqs. (2.98), the axially
compressed perfect isotropic shell initially deforms axisymmetrically along the
path OA (see Figure 2.34) until a maximum (or limit) load A is reached at point
A. However, in this case there exist many bifurcation points along the fundamental
path between O and A. Hence, once the lowest bifurcation load c is reached, the
initial failure of the perfect structure will be characterized by a rapidly growing
asymmetric deformation along the path BD with a decreasing axial load . Notice
that in this case, the (axisymmetric) collapse load of the perfect structure A is of
no engineering significance.
The linearized stability equations for the determination of the critical load c at
the bifurcation point can be derived by the application of the adjacent equilibrium
criterion. To investigate the existence of adjacent equilibrium configurations one

Figure 2.34 Bifurcation point and limit point via nonlinear analysis
Physical Concepts Types of Observed Behavior and Their Meaning 53

assumes that the two variables w, f are given by


w D w0 C w,
O O
f D f0 C f 2.99
where w0 , f0 represent the prebuckling solutions along the fundamental path and w,O
O
f represent small perturbations at buckling. Direct substitution of these expressions
into Eqs. (2.98a) and (2.98b) and deletion of squares and products of the pertur-
bation qualities, yields a set of nonlinear governing equations for the prebuckling
quantities w0 , f0 which are identical in form to Eq. (2.98), and a set of linearized
O f
stability equations governing the perturbation quantities w, O
1O O D0
Dr4 wO C f,xx  LNL f0 , w
O  LNL w0 , f (2.100a)
R
O  Eh wO ,xx C Eh LNL w0 , w
r4 f O D0 (2.100b)
R
where
LNL S, T D S,yy T,xx  2S,xy T,xy C S,xx T,yy . 2.101

a. Axial Compression

First consider the stability of a cylindrical shell that is simply supported at its
ends and subjected to a uniformly distributed axial compressive load P. Under
this loading the prebuckling deformation of the shell is axisymmetric as shown in
Figure 2.35a. The critical load Pc is the lowest axial load at which the axisymmetric
equilibrium state ceases to be stable.
Assuming that the shell is sufficiently long so that the effect of bending of the
shell wall close to the ends can be neglected, then the prebuckling state can be
approximated by the following membrane state
P Eh2
Nx0 D  D  , Ny0 D Nxy0 D 0, w0 D constant 2.102
2R cR
where D Nx /Nc and Nc D Eh2 /cR. See also below an alternate definition given
for by Eqs. (2.110) and (2.111).

Figure 2.35 Axially compressed cylinder


54 The Concepts of Elastic Stability

Thus the axisymmetric form in Figure 2.35a in effect is replaced by that in


Figure 2.35b. Notice that this membrane state satisfies the nonlinear governing
equations of the prebuckling path, Eq. (2.98), identically and reduces the linearized
stability equations, Eq. (2.100) to the following set of constant coefficient equations
1O Eh2
Dr4 wO C f,xx C wO ,xx D 0 2.103a
R cR

r4 fO  Eh wO ,xx D 0. 2.103b
R
Recalling that for constant D and simply supported boundary conditions wO D
wO ,xx D 0 at x D 0, L, then these equations admit separable solutions of the form
x y O D B sin m x cos n y
wO D A sin m cos n , f 2.104
L R L R
leading to a standard eigenvalue problem with the eigenvalues
 
1 ˛2m C ˇn2 2 ˛2m
c,mn D C 2 2.105
2 ˛2m ˛m C ˇn2 2

and the eigenfunctions


3
x y O D  Eh ˛2m x y
O D h sin m
W cos n , f sin m cos n 2.106
L R 2c ˛2m C ˇn2 2 L R
where
  2  2 
2 Rh 1
2 Rh
˛2m Dm ; ˇn2 Dn ; cD 31  2 . 2.107
2c L 2c R
Notice that the eigenvalues c,mn depend not only on the geometric parameters
but also on the axial and circumferential wave numbers m and n.
For cylinders of intermediate length, a close estimate of the smallest eigenvalue
may be obtained directly by analytical minimization of c,mn with respect to the
quantity mn D ˛2m C ˇn2 2 /˛2m in Eq. (2.105). Differentiation leads to the result
that c,mm is a minimum for
˛2m C ˇn2 2
mn D D 1. 2.108
˛2m
Thus all mode shapes which satisfy Eq. (2.108) have the same (lowest) eigenvalue
of c D 1. Regrouping Eq. (2.108) one gets the well known Koiter circle [2.36]

˛2m C ˇn2  ˛m D 0 2.109


which is the locus of a family of modes belonging to the lowest eigenvalue
c
c D D 1. 2.110
c
Physical Concepts Types of Observed Behavior and Their Meaning 55

Notice that Eq. (2.110) is normalized by


Eh E h
c D D 2.111
cR 31  2  R
the critical buckling stress for axially compressed circular cylindrical shells, derived
shortly after the turn of the century independently from each other by Lorenz [2.37]
and [9.44], Timoshenko [9.45] and Southwell [9.29], Volume 2. See also the histor-
ical review of shell buckling in Chapter 9, Volume 2.
For short cylinders, because m and n are integers the analytical minimization to
arrive at Eq. (2.108) is inadmissible. In such cases Eq. (2.105) must be evaluated
repeatedly for different values of m and n in a trial-and-error procedure to determine
the critical load. If the cylinder is so short that
 
R 2 R
 > 2c 2.112
L h
then during buckling only a half-wave in the axial direction will be formed and
the smallest value of Eq. (2.105) is obtained for n D 0. Thus
 

   

c 1 Rh  2 1
c D D c,m0 D C   . 2.113
c 2  2c L Rh  2 
2c L
By taking the length of the cylinder shorter and shorter, the second term in
Eq. (2.113) becomes smaller and smaller in comparison with the first term. Thus,
by neglecting it one obtains
c Rh   2
c D D 2.114
cl 4c L
or  2
2 E h
c D 2
2.115
121    L
which is Euler’s formula for a “wide column”, i.e. a flat plate that is simply
supported at the loaded edges and free along the unloaded edges.
A very long cylinder can buckle as an Euler column with undeformed cross-
section (m D n D 1). The Donnell formulation used does not yield the correct
result for this case as can be seen from Figure 2.36a. Comparing these results with
the values displayed in Figure 2.36b, which are based on Love’s theory (Eq. (i) on
p. 464 of [2.1]), one sees that Donnell’s approach also yields somewhat inaccurate
results for moderately long cylinders. The differences between the predictions of
the two theories can be seen more precisely in Figure 2.37. Notice that the results
of Love’s theory show the proper limiting behavior for very long shells.
The Euler buckling load of very long thin-walled cylinders can be obtained by
setting I D R3 h and A D 2Rh in the appropriate column Eq. (2.3) yielding
 
Pc EI E R 2
c D D 2 2 D 2 . 2.116
A AL 2 L
56 The Concepts of Elastic Stability

Figure 2.36a Buckling diagram for axial compression based on Donnell’s theory, R/h D 1000

Figure 2.36b Buckling diagram for axial compression based on Love’s theory, R/h D 1000
Physical Concepts Types of Observed Behavior and Their Meaning 57

Figure 2.37 Comparison of buckling load predictions based on Donnell and Love type theories

Additional results dealing with orthotropic and anisotropic shells can be found, for
example, in [2.38], [2.39], [2.40] and [2.41].

b. Combined External Pressure and Axial Compression

If the shell is simply supported at its ends then under the simultaneous action of
uniform lateral pressure and axial compression the prebuckling deformation of the
shell is axisymmetric as shown in Figure 2.38a. The critical pressure pc is defined
as the lowest pressure at which the axisymmetric form loses its stability.
Again it is assumed, for simplicity, that the shell is sufficiently long so that the
prebuckling state can be approximated by the following membrane state;
Eh2 Eh2
Nx0 D  ; Ny0 D pe R D pN e ; Nxy0 D 0;
w0 D constant .
cR cR
2.117
Notice that thus, in effect, the axisymmetric form in Figure 2.38a is replaced by
that in Figure 2.38b. It can easily be verified that this membrane state satisfies
the nonlinear governing equations of the prebuckling path, Eq. (2.98), identically
(whereby p D pe ) and reduces the linearized stability equations, Eqs. 2.100, to
the following set of constant coefficient equations;
1O Eh2 Eh2
Dr4 wO C f,xx C pN e wO ,yy C wO ,xx D 0 (2.118a)
R cR cR

Figure 2.38 Cylinder subjected to uniform external lateral pressure


58 The Concepts of Elastic Stability

O Eh
r4 f wO ,xx D 0. (2.118b)
R
The boundary conditions and the separable solutions are the same as for the
preceding example. The use of Eqs. (2.104) leads to a standard eigenvalue problem
with eigenvalues
 
1 ˛2m C ˇn2 2 ˛2m ˇn2
D C  N
p e . 2.119
2 ˛2m ˛2m C ˇn2 2 ˛2m
A single parameter eigenvalue can be obtained introducing the relation
D RO pN e 2.120
where RO is a nondimensional constant. Notice that if RO D 0 the pressure acts only
on the lateral surface, whereas if RO D 1/2 then the pressure contributes also to axial
compression through the end plates, forming the so-called hydrostatic pressure case.
With the help of this expression the eigenvalues can be written as
 
˛2m ˛2m C ˇn2 2 2
˛m
pN c,mn D C 2 . 2.121
O 2m C ˇn2 
2R˛ ˛2m ˛m C ˇn2 2
The eigenfunctions are the same as for the preceding example (see Eqs. (2.106)).
Considering Eq. (2.121), a distinct eigenvalue corresponds to each pair of values
m and n and it is seen that the smallest eigenvalue corresponds in every case to
m D 1. For particular values of L/R and R/h, the n corresponding to the smallest
eigenvalue may be determined by trial-and-error.
Numerical results based on Eq. (2.121) are shown in Figure 2.39. From these
curves, calculated for different R/h ratios, it is seen that for shorter tubes the
critical external pressure pc increases rapidly as the ratio L/R decreases. On the
other hand for long tubes, for L/R > 50 say, the critical external pressure does not
depend on the length. Its value can also be deduced from Eq. (2.64), the critical
pressure for a ring subjected to external fluid pressure, as follows. Recalling that
the compressive force per unit length Ny acting on the elemental ring of unit width
is equal to pc R, then from Eq. (2.64)
EI
Nyc D pc R D 3 . 2.122
R2
If one now replaces E by E/1  2  and sets I D h3 /12, then Eq. (2.122) yields
 3
E h
pc D 2
2.123
41    R
the critical buckling pressure for long tubes subjected to uniform external pressure.
It also becomes apparent from the results displayed in Figure 2.39 that for n D 4
or less there is a noticeable difference between the predictions of Eq. (2.121), which
is based on Donnell’s theory, and the results of the Love theory of [2.1] (Eq. (d) on
p. 496). For n D 2, as well known, the Donnell values are about 33 percent too high.
Physical Concepts Types of Observed Behavior and Their Meaning 59

Figure 2.39 Buckling diagram for uniform external lateral pressure

As described in the historical review of shell buckling in Chapter 9, Volume 2


the first investigations of the stability of externally pressurized tubes were made
by Southwell [9.26] and [9.29], Volume 2) and von Mises [9.27] and [9.31],
Volume 2). For results dealing with orthotropic and anisotropic shells the reader
may consult [2.42], [2.43] and [2.44].

c. Combined Torsion and Axial Compression

Assuming, for simplicity, that the shell is sufficiently long, then the prebuckling
solution under the simultaneous action of axial compression and torsion can be
approximated by the following membrane state;

Eh2 Mt Eh2
Nx0 D  ; Ny0 D 0; Nxy0 D D N
 ; w0 D constant 2.124
cR 2R2 cR
where Mt is the applied torsional moment.
60 The Concepts of Elastic Stability

Direct substitution shows that this membrane state satisfies the nonlinear
governing equations of the prebuckling path, Eqs. (2.98) identically and that the
linearized stability equations, Eqs. (2.100), are reduced to the following set of
constant coefficient equations;
1O Eh2 Eh2
Dr4 wO C f,xx  2N wO ,xy C wO ,xx D 0 (2.125a)
R cR cR

r4 fO  Eh wO ,xx D 0. (2.125b)
R
Notice that these equations differ markedly from the previously derived stability
equations (see Eqs. (2.103) and (2.118)) in that in the out-of-plane equilibrium
equation one encounters both odd and even derivatives of wO with respect to the
same independent variable. This indicates that one can no longer satisfy the stability
equations by using separable solutions in the form of simple products of sines and
cosines. Physically this means that there are no generators which remain straight
during buckling and which form a system of straight nodal lines for a buckled
surface.
Under torsional loading the buckling deformation of a cylindrical shell consists
of a number of circumferential waves that spiral around the cylinder from one end
to the other. If one now assumes that the buckling mode is represented by
 N   N 
 x y  x y
wO D h Cmn sin m sin n C h Dmn sin m cos n 2.126
mD1
L R mD1
L R

an expression that satisfies simply supported boundary conditions wO D wO ,xx D 0 at


x D 0, L, then an approximate solution of the linearized stability equations can be
obtained as follows.
First, the compatibility equation (2.125b) is solved exactly for the stress function
fO in terms of the assumed radial displacement w. O Since it is assumed that the shell
is sufficiently long so that the effect of bending of the shell wall close to the
ends can be neglected, only a particular solution of equation (2.125b) needs to be
considered. Secondly, the equation of equilibrium (2.125a) is solved approximately
by substituting therein f O and wO and then applying Galerkin’s procedure. Carrying
out the steps yields for a given number of circumferential full waves n the following
homogeneous system of two simultaneous algebraic equations;
1
Mm Cmn C Nmj Djn D 0
N
m D 1, 2, 3, . . . , N 2.127
1
Nmj Cjn C Mm Dmn D 0
N
where

Mm D m Nc,mn
4
Physical Concepts Types of Observed Behavior and Their Meaning 61

N   N  
jÐm jÐm
Nmj Djn D jšm Djn ; Nmj Cjn D jšm Cjn
jD1
m 2  j2 jD1
m 2  j2

Nxy xy
N D D (2.128)
hc cl
 2 
1 ˛2m C ˇn2 ˛3m ˛m
Nc,mn D C  2 
4 ˛m ˇn ˇn ˛2m C ˇn2 2 ˇn

jšm D 1 if j š m D odd integer
D 0 otherwise.
Using matrix notation Eqs. (2.127) can be put into the form of a standard eigen-
value problem
[[A]  N [B]]X D 0 2.129
¾
which can be solved routinely on a digital computer. Since the structure buckles
at the lowest stress at which instability can occur, for a given shell N is minimized
with respect to the circumferential wave number n. This is done by truncating the
determinant of the coefficients of Eq. (2.129) and finding the lowest eigenvalue
by matrix iteration. The size of the determinant is increased until the eigenvalue
converges to the desired accuracy (say, five significant figures).
Results for R/h D 1000 and different L/R ratios are displayed in Figure 2.40.
As can be seen, for shorter shells the critical normalized torque parameter Nc
increases rapidly as the ratio L/R decreases. Notice also, that by taking the radius
of the cylinder larger and larger, while keeping its length constant, the lower bound
festoon curve for Nc approaches the critical shear load of an infinitely long strip
with simply supported edges obtained by Southwell and Skan [2.45]
2 D
c D 5.35 2.130
L2 h
where again D D Eh3 /121  2 .
Limiting results for large values of L/R, when the shell will buckle with two full
waves in the circumferential direction, have been derived, for example, in [2.2]
using Donnell’s theory yielding
 3/2
E h
c D 0.272 2 3/4
2.131
1    R
and in [2.1] using a Love type theory yielding
 3/2
E h
c D 0.236 . 2.132
1  2 3/4 R
Once again, as noted earlier, for n D 2 Donnell’s equations are inaccurate.
Also shown in Figure 2.40 are solutions based on the following buckling mode
 x y
wO D hCmn sin m  n 2.133
L R
62 The Concepts of Elastic Stability

Figure 2.40 Buckling diagrams for cylinders subjected to torsion

where Cmn is a constant and m, n are integers. Equation (2.133) satisfies the
requirement of periodicity in the circumferential coordinate, but does not satisfy
any of the commonly used boundary conditions at the cylinder ends. Consequently,
this simple expression may only be used for sufficiently long cylinders, whose end
conditions have little influence on the magnitude of the critical load.
Using the Donnell type theory and proceeding as outlined earlier one obtains a
particularly simple solution of the linearized stability equations (2.125) with the
following expression for the eigenvalues
 
1 ˛2m C ˇn2 2 ˛3m ˛m
Nc,mn D C 2 2 2
 2.134
4 ˛m ˇn ˇn ˛m C ˇn  2 ˇn

where Nc,mn D xy /c . Notice that the eigenvalues Nc,mn depend not only on the
geometric parameters and the specified axial load D /c , but also on the axial
and circumferential wave numbers m and n.
Timoshenko used expressions similar to the one given by Eq. (2.133) to solve
Love type stability equations in [2.1]. His solution curve agrees well for n ½ 4 full
waves in the circumferential direction with the one based on Donnell’s stability
equations. However, for large values of L/R when the shells buckle with two full
Physical Concepts Types of Observed Behavior and Their Meaning 63

waves in the circumferential direction, one can observe the well known fact that
Donnell’s equations yield about 10 per cent higher values than solutions based on
the more accurate Love type theory.
The first investigation of buckling of cylindrical shells under torsion is due
to E. Schwerin [2.46]. Buckling under torsion is further discussed in Chapter 9,
Volume 2. For a complete review of the torsion problem the interested reader may
consult Yamaki’s book [2.47].

d. Combined Bending and Axial Compression

If a cylindrical shell is relatively short and the shell edges are held circular, then
the circumferential flattening of the cylinder cross section caused by the bending
moment can be neglected. In this case the prebuckling state under an external load
consisting of combined bending and axial compression can be approximated quite
accurately by the following membrane state;
Eh2  y
Nx0 D  RO a C RO b cos , Ny0 D Nxy0 D 0, w0 D constant 2.135
cR R
where
N0 Nb P M0
RO a D ; RO b D ; N0 D ; Nb D ;
Nc Nc 2R R2
Eh2 
Nc D ; cD 31  2 .
cR
Notice that this membrane state does not satisfy rigorously the nonlinear
equations governing the prebuckling state, Eqs. 2.98. However, because of its
simplicity it has been widely used in the literature (see [2.40], [2.48] and [2.49]) to
obtain approximate solutions. The linearized stability equations (2.100) then reduce
to the following set of variable coefficient equations
1O Eh2  y
Dr4 wO C f ,xx C O
R a C RO b cos wO ,xx D 0 (2.136a)
R cR R
O  Eh wO ,xx D 0.
r4 f (2.136b)
R
If one now assumes that the buckling mode is represented by
 N 
x  y
wO D h sin m Cmn cos n 2.137
L nD1 r

an expression that satisfies simply supported boundary conditions wO D wO ,xx D 0 at


x D 0, L, then an approximate solution of the linearized stability equations can be
obtained as follows.
First the compatibility equation (2.136b) is solved exactly for the stress function
O
f in terms of the assumed radial displacement w. O Here it is assumed that the
64 The Concepts of Elastic Stability

effect of bending of the shell wall close to the ends can be neglected. Thus only
a particular solution of equation (2.136b) needs to be considered. Secondly, the
equation of equilibrium (2.136a) is solved approximately by substituting therein
fO and wO and then applying Galerkin’s procedure. Carrying out the steps yields
for a given number of axial half waves m the following homogeneous system of
algebraic equations;
 c,mn  RO a Cmn  21 RO b [1 C υ1n  υ0N Cm,n1 C 1  υ0n Cm,nC1 ]
n D 1, 2, 3, . . . , N 2.138
where
 2 
1 ˛2m C ˇn2 ˛2
c,mn D C  2 m 2
2 ˛2m ˛m C ˇn
2.139
 2
Rh   2 2 Rh 1 
˛2m D m2 , 2
ˇn D n , cD 31  2 
2c L 2c R
and υ1n , υon , υoN are Kronecker deltas.
Using matrix notation Eq. (2.138) can be put into the following form;
  
c,m0  RO a  12 RO b Cm0
 RO c,m1  RO a  21 RO b  
 b   Cm1 
  
  1 RO b c,m2  RO a  21 RO b  Cm2 
 2  D0
 ... ... ...  ... 
 
  21 RO b c,mN1  RO a  21 RO b   . . . 
 21 RO b c,mN  RO a CmN

(2.140)
This tridiagonal matrix eigenvalue problem can be solved very conveniently by a
recursive Gaussian elimination scheme originally derived by Potter [2.50] and used
later extensively by the Harvard group under Budiansky [2.51]. Either the normal-
ized axial load parameter RO a D  or the normalized bending moment parameter
RO b can be chosen as the eigenvalue, whereby in each case the other load param-
eter has a specified fixed value. Since the shell buckles at the lowest stress at
which instability can occur, the eigenvalue chosen is minimized with respect to
the axial half-wave number m. This is done by truncating the size of the matrix
in Eq. (2.140) for a given value of m and finding the lowest eigenvalue by matrix
iteration. The size of the determinant is increased until the eigenvalue converges
to the desired accuracy (say, five significant figures).
Numerical results for R/h D 100 and different L/R ratios are displayed in
Figure 2.41 for pure bending (RO a D 0). Notice that for shorter shells the normalized
bending stress ratio RO b increases rapidly as the ratio L/R decreases. Further,
whereas for shorter shells (L/R < 0.5, say) for certain L/R ratios RO b may vary
noticeably, for longer shells (L/R > 1.0, say) the critical normalized bending stress
ratio RO b can be set equal to the lower bound of the festoon curves of about 1.014.
Physical Concepts Types of Observed Behavior and Their Meaning 65

Figure 2.41 Variation of buckling stress ratio with cylinder length R/h D 100

Thus for pure bending the maximum critical bending stress is only slightly higher
than the critical stress for axial compression only.
Here it must be mentioned that the statement made in [2.1] on p. 483 about the
maximum critical stress for bending alone is 1.3 times the critical stress for pure
compression does not hold in general. It is only true for the particular set of geometric
and material properties used by Flügge [2.48] for his “habilitation” paper.
Finally to check whether it is safe to neglect the effect of ovalization of the
circular cross-section caused by the applied bending moment one can use the
results of [2.52] here reproduced in part in Figure 2.42. The authors of this paper

Figure 2.42 Comparison of collapse moments of cylinders under pure bending with classical
results (from [2.52])
66 The Concepts of Elastic Stability

used the two-dimensional finite difference code STAGS [2.53] to calculate the
collapse bending moment while taking the effects of boundary conditions and
geometric nonlinearities in the prebuckling state into account. As can be seen from
Figure (2.42) for shells of moderate length (L/R < 3, say) the results obtained with
Eq. (2.140) are quite accurate.

2.1.12 Buckling of Shells of Revolution

Besides circular cylindrical shells many structural applications of thin-walled shells


consist of general shells of revolution, the middle surface of which is obtained by
rotating a plane curve about an axis in the plane of the curve (see Figures 2.43
and 2.44). The lines of principal curvature on a shell of revolution are called the
parallels and the meridians. The parallels are formed by the intersection of planes
normal to the axis of revolution with the shell surface, whereas the meridians are
the intersections with the shell surface of planes that contain the axis of revolution.
Points on the middle surface are referred to coordinates  and , where  denotes
the angle between the axis of revolution and a normal to the surface, whereas
is the circumferential coordinate. The principal radii of curvature of the surface in
the - and direction are R and R , respectively. It is convenient to introduce an
additional variable r defined as

r D R sin . 2.141

Figure 2.43 Shell of revolution notation and sign convention

Figure 2.44 Meridian of a shell of revolution


Physical Concepts Types of Observed Behavior and Their Meaning 67

Then, as can be seen from Figure 2.43


ds D R d; ds D r d . 2.142
Furthermore, from Figure 2.44
dr D ds cos  2.143
Combining the last two equations one obtains the additional relationship
dr
D R cos . 2.144
d
Notice that for a shell of revolution the geometric quantities are independent of
the circumferential coordinate .
In this section the relatively simple Donnell Mushtari Vlasov type quasi-
shallow equations will be used, which are based on the following middle-surface
kinematic relations
1 1
ε D ε C ˇ2  D ˇ, 2.145
2 R
1 1 cos 
ε D ε C ˇ 2  D ˇ , C ˇ
2 r r
   
1 r 1 1
 D ε C ˇ ˇ  D ˇ C ˇ, .
2 R r , r

For quasi-shallow shells the terms containing u and v are omitted from the rotation
expressions, thus
1
ˇ D  w, 2.146
R
1
ˇ D  w,
r
and
1
ε D u, C w
R
1
ε D v, C u cos  C w sin  2.147
r
r v 1
ε D C u, .
R r , r
Employing the stationary potential energy criterion, the following set of nonlinear
equations are derived in [2.2] for isotropic shells of revolution;
rN , C R N ,  R cos N D rR p 2.148a
rN , C R N , C R cos N D rR p 2.148b
68 The Concepts of Elastic Stability
   
1 1
rM , C R M , C cos M , C M C M  ,
R , r ,
   
cos  N N
C 2 R M  rR C
r , R R
 [rN ˇ C rN ˇ , C R N ˇ C R N ˇ , ] D rR p. 2.148c
The variables R , R and r characterize the shape of the middle surface of the
undeformed shell, and are functions of  alone. Three equations in the three vari-
ables u, v, w, which denote middle surface displacement components in the ,
and normal directions, respectively, can be derived by introducing the isotropic
constitutive equations
N D Cε C ε  M D D C   2.149
N D Cε C ε  M D  C  
1   M C M 
N D C  D D1  
2 2
and the kinematic relations from Eq. (2.145) into Eq. (2.148).
The linearized stability equations for the determination of the critical load at the
bifurcation point may be obtained from the nonlinear equilibrium Eqs. (2.148) by
application of the adjacent equilibrium criterion. To investigate the existence of
adjacent equilibrium configurations one assumes that
u D u0 C uO , v D v0 C vO , w D w0 C wO 2.150
where (u0 , v0 , w0 ) represents the equilibrium whose stability is under investigation,
(u, v, w) is an adjacent equilibrium configuration at the same value of applied load
as the configuration (u0 , v0 , w0 ) and (Ou, vO w)
O is an arbitrarily small incremental
displacement. Direct substitution of these expressions into Eqs. (2.148) and deletion
of squares and products of the perturbation quantities, yields a set of nonlinear
governing equations for the prebuckling quantities (u0 , v0 , w0 ) which are identical
in form to Eqs. (2.148), and a set of linearized stability equations governing the
perturbation quantities (Ou, vO w).
O
If the applied load is axisymmetric, then the deformation prior to the loss of
stability is also axisymmetric. If, further, one assumes that the prebuckling rotations
are zero then the axisymmetric equations governing the prebuckling state become
rN0 ,  R cos N 0 D rR p (2.151a)
rN 0 , C R cos N 0 D rR p (2.151b)
 
1
rM0 ,  cos M 0 ,  rN0 C R sin N 0  D rR p (2.151c)
R ,

and the linearized stability equations are


O  , C R N
r N O  ,  R cos N
O D 0 (2.152a)
r N O , C R cos N
O  , C R N O  D 0 (2.152b)
Physical Concepts Types of Observed Behavior and Their Meaning 69
 
1 R
O  ,
r M C M O , C M
O ,  cos M O  C M
O  ,
R , r
R
C2 cos M O  C R sin N
O  ,  r N O   fN0 r Ǒ  , C N 0 r Ǒ ,
r
C N 0 R Ǒ , C N 0 R Ǒ , g D 0 (2.152c)
where
O  D COε C Oε 
N O  D DO  C O 
M 2.153
O D COε C Oε 
N O D DO C O  
M
1   O  C M
M O 
O  D C
N εO  D D1  O
2 2
and
1 Ǒ D  1 wO ,
εO  D uO , C w
O 2.154
R R
1 Ǒ D  1 wO ,
εO D Ov, C uO cos  C wO sin 
r r
 
r vO 1
εO  D C uO ,
R r , r
with
1 Ǒ
O  D , 2.155
R
1 Ǒ cos  Ǒ
O D , C 
r r
   
1 r 1 Ǒ 1 Ǒ
O  D C , .
2 R r , r
In the following these stability equations will be used to solve buckling problems
of common structural configurations other than cylindrical shells which can be
represented by symmetrically loaded shells of revolution.

a. Externally Pressurized Shallow Spherical Caps

A shallow section S0 of a complete spherical shell is imagined to be isolated as


shown in Figure 2.45. To satisfy the shallowness criterion the rise H of the shell
must be much smaller that the base radius a. The position of a point on the middle-
surface is described by polar coordinates r, . Notice that R D R, a constant and
sin  D r/R. Furthermore, approximately, cos  D 1 and dr D R d. Thus
1
 , D  ,r . 2.156
R
70 The Concepts of Elastic Stability

Figure 2.45 Shallow spherical cap notation and sign convention

Using these approximations the axisymmetric equations governing the prebuckling


state (Eqs. 2.152) become with p D p D 0 and p D pe
rNr0 ,r  N 0 D 0 (2.157)
rNr 0 ,r C Nr 0 D 0
1 1 1
rMr0 ,rr  M 0,r  Nr0 C N 0  D pe
r r R
and the linearized stability equations reduce to

r N O r ,  N
O r ,r C N O D 0 (2.158a)
O r ,r C N
r N O , C N
O r D 0 (2.158b)
1 2 r
O r ,rr C M
r M O ,  M O ,r C M
O r C M O r ,r C M O r ,  N Or CN
O 
r r R
 frNr0 Ǒ r ,r C rNr 0 Ǒ ,r C Nr 0 Ǒ r, C N 0 Ǒ , g D 0. (2.158c)

Assuming that prior to buckling the perfect spherical shell is in a uniform membrane
state of stress then
Nr0 D N 0 D  12 pe R, Nr 0 D 0 2.159

with an associated uniform inward radial displacement of

1   R2
w0 D pe . 2.160
2 Eh
Notice that this uniform membrane state of stress satisfies the axisymmetric
equations governing the prebuckling state identically and that this prebuckling
state is rotation free. Substitution into the linearized stability equations yields
O r ,r C N
r N O r ,  N
O D 0 (2.161a)
r N O , C N
O r ,r C N O r D 0 (2.161b)
Physical Concepts Types of Observed Behavior and Their Meaning 71

1 2
r MO r ,rr C M O ,  MO ,r C M
O r C M O r ,r C M O r ,
r r
r 1
 N Or CN O  C pN e Rfr Ǒ r ,r C Ǒ , g D 0. (2.161c)
R 2
Introduction of the appropriate incremental constitutive and kinematic relations for
the stress and moment resultants leads after some regrouping to a coupled set of
three homogeneous equations in uO , vO , w. O As has been shown, for instance in [2.2],
a simpler set of two equations in two unknowns can be derived as follows. If one
defines an Airy stress function f O such that
1O 1 O O ,rr ; N 1O 1 O
Nr D f ,r C 2 f, : NO D f O r D  f ,r C 2 f, 2.162
r r r r
then the in-plane equilibrium equations (Eqs. (2.161a) and (2.161b)) are identically
satisfied and, with the help of the appropriate constitutive and kinematic relations
for M O and M
O r, M O r the out-of-plane equilibrium equation (Eq. 2.161c) can be
reduced to
1 O 1
DrwO C r2 f C pe Rr2 wO D 0 2.163
R 2
where
1 1
r2   D  ,rr C  ,r C 2  , 2.164
r r
r4   D r2 r2  . 2.165
A second equation in terms of fO and wO is the compatibility equation that as has
been shown in [2.2] can be written as
1 4O 1
r f  r2 wO D 0. 2.166
Eh R
The homogeneous Eqs. (2.163) and (2.166) have nontrivial solutions only for
discrete values of the external pressure pe . The smallest such value is the critical
pressure pe .
A particularly simple solution was presented by Hutchinson in [2.54]. Using the
coordinate transformation
x D r cos , y D r sin 2.167
the Laplacian operator reduces to the cartesian form. That is
1 1
r2   D  ,rr C  ,r C 2  , D  ,xx C  ,yy 2.168
r r
Applying this transformation it is seen that Eqs. (2.163) and (2.166) admit separable
solutions of the form
x y
wO D A cos x cos y 2.169
R R
fO D B cos x x cos y y
R R
72 The Concepts of Elastic Stability

leading to a standard eigenvalue problem with the eigenvalues


p 1
pN D D ˛2x C ˛2y  C 2 2.170
h/Rc ˛x C ˛2y 

where
 2  2 
Rh1 Rh1 Eh
˛2x D 2x ; ˛2y D 2y ; c D ; cD 31  2 
2c R 2c R cR
and x , y are integers.
Notice that a close estimate of the smallest eigenvalue may be obtained
directly by analytical minimization of pN with respect to the quantity  D ˛2x C ˛2y
Eq. (2.170). Differentiation leads to the result that pN is a minimum for

 D ˛2x C ˛y2 D 1. 2.171

Thus all mode shapes which satisfy Eq. (2.171) have the same (lowest) eigenvalue
of pN D 2. Thus the critical pressure for externally pressurized (shallow) caps is
 2
h 2E h
pc D 2 c D  . 2.172
R 31  2  R
This is the same result that was obtained for a complete spherical shell using
Legendre functions pn cos  by Flügge ([2.55], p. 477) and by Timoshenko and
Gere ([2.1], p. 517).
It is interesting, that if one calculates the corresponding critical stress from
Eq. (2.159) then
  2 
1 2E h
Nr0 D N 0 D c h D   R
2 31  2  R
E h
c D  
31  2  R
the same magnitude as the critical stress for an axially compressed cylindrical shell
of radius R and wall-thickness h (see Eq. (2.111)). Finally, it must be remembered
that the assumed functions in Eq. (2.169) of the separable solution do not satisfy
the boundary conditions at the edges of the spherical cap. Thus the validity of
the present simplified analysis is limited to cases where the wavelengths of the
buckling pattern are small compared to the radius of the shell or what is the
same if the wave numbers x and y are both large compared to unity.

b. Toroidal Shell Segments under External Pressure p D pe 

As mentioned earlier the middle surface of a shell of revolution may be formed


by rotation of a plane curve about an axis in the plane of the curve. If the plane
Physical Concepts Types of Observed Behavior and Their Meaning 73

Figure 2.46 Toroidal shell segments notation and sign convention

curve is a circular arc of radius R D b then the surface formed is a segment of a


torus. As can be seen from Figure 2.46 for the middle surface of the segment of a
toroidal shell
R D b, R D a and r D a  b1  sin . 2.173
Further, for a sufficiently shallow L/R − 1 segment in the region of the equator
of the torus, the angle  is approximately equal to /2. Then cos  D 0, sin  D 1
and r D a. The governing equations can further be simplified if one introduces
x and y as axial and circumferential coordinates, as indicated in Figure 2.46.
Notice that
dx D Rx d D b d and dy D Ry d D a d . 2.174
The equations governing the membrane prebuckling state (Eqs. (2.151)) become
Nx0,x D 0 2.175a
Nxy0,x D 0 2.175b
Nx0 Ny0
C D0 2.175c
b a
and the linearized stability equations reduce to
N O xy,y D 0
O x,x C N (2.176a)
NO xy,x C N
O y,y D 0 (2.176b)
 
Ox
N Oy
N
O x,xx C M
M O xy C M
O yx ,xy C M
O y,y  C
b a

 [Nx0 Ǒ x C Nxy0 Ǒ y ,x C Ny0 Ǒ y C Nxy0 Ǒ x ,y ] D 0 (2.176c)


where
O x D COεx C Oεy 
N O x D DO x C O y 
M 2.177
O y D COεy C Oεx 
N O y D DO x C O y 
M
1   O xy C M
M O yx
O xy D C
N O xy D D1  Oxy
2 2
74 The Concepts of Elastic Stability

and
εO x D uO ,x Ǒ x D wO ,x O x D Ǒ x,x 2.178
εO y D vO ,y Ǒ y D wO ,y O y D Ǒ y,y
O xy D uO ,y C vO ,x O xy D 21  Ǒ x,y C Ǒ y,x .
Stability analysis of both bowed-out and bowed-in shallow toroidal shell segments
under three different loading conditions have been presented by Stein and McElman
in [2.56] and Hutchinson in [2.57].
Introducing an Airy stress function fO such that

N O ,yy ,
Ox Df N O ,xx ,
Oy D f O ,xy
O xy D f
N 2.179
then the in-plane linearized stability equations (Eqs. (2.176a) and (2.176b)) are
identically satisfied and the out-of-plane stability equation (Eq. (2.176c)) can be
written as
1O 1O
Dr4 wO C f ,yy D f,xx  Nx0 w
O ,xx C 2Nxy0 wO ,xy C Ny0 wO ,yy  D 0. 2.180
b a
A second equation involving wO and f O is provided by the appropriate compatibility
condition, which can be derived by eliminating the in-plane displacements uO , vO in
the strain displacement relations of Eqs. (2.178). This yields
1 4O 1 1
r f  wO ,yy  wO ,xx D 0. 2.181
Eh b a
Assuming that the prebuckling state is torsion free, that is Nxy0 D 0, and using the
simply supported conditions
O Df
wO D wO ,xx D f O ,xx D 0 at x D 0, L 2.182
then together with Eqs. (2.180) and (2.181) one has a linear eigenvalue problem
for determining the critical buckling load. Using the eigenfunction
x y
wO D h sin m sin n 2.183
L a
3 2 2
O D  Eh ˛m C ˇn a/b sin m x sin n y
f
2c ˛2m C ˇn2 2 L a
separation of variables yields the following characteristic equation
 
1 2 2 2 ˛2m C ˇn2 a/b2 Eh2
˛m C ˇn  C C Nx0 ˛2m C Ny0 ˇn2  D 0 2.184
2 ˛2m C ˇn2 2 ca

where
 2
ah   2 ah 1 
˛2m D m2 , ˇn2 D n2 , cD 31  2 . 2.185
2c L 2c a
Physical Concepts Types of Observed Behavior and Their Meaning 75

Assuming now that the perfect shallow toroidal segment is loaded by uniform
external lateral pressure only, then the corresponding membrane stress state

Nx0 D 0, Ny0 D pe a, Nxy0 D 0 2.186

satisfies the equations governing the prebuckling state (Eqs. (2.175)) identically
and from the characteristic Eq. (2.184) one obtains the following expression for
the eigenvalue;
 
2 2 2
1 ˛ C ˇ a/b
pN c,mn D 2 ˛2m C ˇn2 2 C m 2 n 2 2 2.187
2ˇn ˛m C ˇn 

where pN c,mn D pe /Eh2 /ca2 . The critical buckling pressure pN c corresponds to the
minimum value of pN c,mn among all possible integer values of m and n. It is easily
shown that for the two loading conditions considered in this section the minimum
value of pN c,mn always occurs for m D 1 if ˇn is treated as a continuous variable.
This is valid if n is sufficiently large, which must be checked á posteriori. Notice
that the restriction to n > 4, say, is necessary in any case since Donnell type
equations are used.
Considering the predicted critical buckling pressures plotted in Figure 2.47, one
must notice the significant difference between the predicted buckling strengths
of the bowed-out Ry /Rx > 0 and Ry /Rx > 0 the bowed-in shells which have
otherwise essentially the same dimensions.
Considering now the hydrostatic pressure case there is a prebuckling axial
compressive stress in addition to the circumferential stress. Notice that the

Figure 2.47 Buckling diagrams for toroidal shell segments under external lateral pressure
76 The Concepts of Elastic Stability

membrane stress state

Nx0 D  21 pe a, Ny0 D  12 pe a2  a/b, Nxy0 D 0 2.188

satisfies the equations governing the prebuckling state (Eqs. (2.175)) identically
and that the characteristic Eq. (2.184) yields now the following expression for the
eigenvalue
 
2 2 2
1 ˛ C ˇ a/b
pN c,mn D 2 2
˛m C ˇn2 2 C m 2 n 2 2 2.189
˛m C 2  a/bˇn2 ˛m C ˇn 

where pN c,mn D pe /Eh2 /ca2 . Once again the critical buckling pressure pN c corre-
sponds to the minimum value of pN c,mn for m D 1 and treating ˇn as a continuous
variable. From the plots of the predicted critical buckling pressures shown in
Figure 2.48 one must conclude that the trends are similar to those of the lateral pres-
sure case, that is, there is significant discrepancy between the buckling pressures
of the bowed-out Ry /Rx > 0 and the bowed-in Ry /Rx < 0 shells.
Notice that when Ry /Rx D 1 the shell is locally spherical at each point on its
surface and the prebuckling stresses are exactly those corresponding to a complete
spherical shell of similar radius and thickness, namely Nx0 D Ny0 D  21 pe a. As
can be seen from Figure 2.48 the critical buckling pressure of the Ry /Rx D 1.0
case for Z ½ 3 is also that for a complete spherical shell
 2
Eh2 2E h
pc D 2.0 2
D . 2.190
ca 31  2  a

Figure 2.48 Buckling diagrams for toroidal shell segments under hydrostatic pressure
Physical Concepts Types of Observed Behavior and Their Meaning 77

c. Toroidal Segments under Axial Tension

The prebuckling state of stress of a perfect toroidal shell segment carrying a


uniformly distributed tensile axial force N0 at its edges is

Nx0 D N0 , Ny0 D N0 a/b, Nxy0 D 0 2.191

as can easily be derived from the equations governing the prebuckling state
(Eqs. (2.175)). Notice that buckling is due to the compressive circumferential stress
Ny0 , which will be induced only if Ry /Rx > 0. In other words, buckling under a
tensile force occurs only for the bowed-out shells. The characteristic Eq. (2.184)
yields in this case the following expression for the eigenvalue
 
T 1 2 2 2 ˛2m C ˇn2 a/b2
c,mn D ˛m C ˇn  C 2.192
2ˇn2 a/b  ˛2m  ˛2m C ˇn2 2

T
where c,mn D N0 /Eh2 /ca. The critical normalized axial tensile force T
c corre-
T
sponds to the minimum value of c,mn for m D 1 and treating ˇn as a continuous
variable. Results of such buckling load calculations are shown in Figure 2.49,
whereas a typical buckling mode is displayed in Figure 2.50. In all cases the
buckled shape is similar to that of a cylinder which buckled under radial pressure,
with one half-wave in the axial direction and many small waves in the circumfer-
ential direction.

Figure 2.49 Buckling diagrams for bowed-out toroidal shell segments under axial tension
78 The Concepts of Elastic Stability

Figure 2.50 Typical buckling mode of a bowed-out toroidal shell


segment under axial tension

d. Domed (torispherical) End-Closures under Internal Pressure

This problem is of special interest to designers of pressure vessels, many of which


have torispherical domes as end-closures. As in the case of shallow toroidal shell
segments under axial tension buckling is caused by the occurrence of compressive
circumferential stresses, which are induced by the internal pressure over parts of
the end-closure (see also Figure 2.51).
The possibility of nonaxisymmetric buckling of internally pressurized torispher-
ical end-closures was first predicted by Galletly in [2.58]. Applying the membrane
equations of an axisymmetric shell of revolution with no torque acting to the tori-
spherical end-closure shown in Figure 2.52 (a torispherical shell is a toroidal shell
jointed to a spherical cap) one gets
d
rN   R N cos  D 0 (2.193a)
d
N N
C D p. (2.193b)
R R
Solving the second equation for N and substituting it into the first equation, one
obtains after multiplying the resulting equation by sin  and some regrouping
d
r sin N  D R R p sin  cos  2.194
d
an expression that can be integrated directly yielding
 
1
N D R R p sin  cos  d 2.195
r sin  0
Ways to evaluate the constant of integration involved are discussed in great detail
by Flügge in [2.55] (see pp. 23 48). Recalling that and that r D R sin  and that
dr D R cos  d (see Eqs. (2.142) and (2.143)) then
 r
p
N D r dr 2.196
R sin2  0
an integral that can be evaluated independently of the shape of the meridian yielding

N D 12 pR . 2.197
Physical Concepts Types of Observed Behavior and Their Meaning 79

Figure 2.51 Simple rig to demonstrate buckling due to internal pressure (courtesy of Prof.
G.D. Galletly)

Figure 2.52 Torispherical end-closure notation and sign convention


80 The Concepts of Elastic Stability

The hoop stress resultant N can then be found from Eq.(2.193b) yielding
 
1 R
N D pR 2  . 2.198
2 R
Considering the torispherical end-closure depicted in Figure 2.52 both radii of
curvature are positive, therefore for internal pressure p > 0 the meridional stress
resultant N will always be positive (tensile). Notice that the hoop resultant N
can be positive or negative, depending on the ratio R /R . From Eq. (2.198) it is
evident that if
R
>2
R
then N will be negative (compressive). The existence of compressive hoop stresses
due to internal pressure indicates that buckling with an asymmetric buckling mode
may occur.
To calculate the critical buckling pressure one must solve the linearized stability
Eqs. (2.153), whereby with no external torque applied N 0 D 0. The other two
prebuckling stress resultants N0 and N 0 are given by Eqs. (2.197) and (2.198),
respectively. Furthermore, as can be seen from Figure 2.52, the prebuckling stress
resultants N0 and N 0 are not constant but vary with  . As a matter of fact, the
presence of a stress-discontinuity in N 0 at the point B indicates that for a rigorous
solution one must use the bending theory to patch up the membrane solutions
for the torus and the spherical cap. Since, however, the bending solutions have a
boundary layer type behavior one may use, as a first approximation, the membrane
solution.
To solve the resulting variable coefficient linearized stability equations for the
rather complex meridional geometry of the torispherical shells one must rely on
numerical methods. Galletly and his coworkers have performed extensive experi-
mental and numerical studies using the BOSOR-4 and BOSOR-5 shell of revolution
codes (see [2.59] [2.63]) to provide buckling and collapse data for the design of
internally pressurized dished ends.

2.1.13 Influence of Nonlinear Effects

In the preceding chapters the linearized stability equations are used to obtain the
buckling loads of the structures considered. Although buckling is a nonlinear
phenomena for many applications the use of the linearized stability equations,
which are amenable to analytical treatment, yields results that are suitable for design
purposes. As has been pointed out in [2.2], there are three situations, however, in
which a nonlinear analysis is needed.

1. In the applications considered up to now, it is assumed that the prebuckling


deformation is rotation free and the primary equilibrium paths are governed
by membrane stress states. If, however, one wants to satisfy the boundary
conditions from the outset, then the prebuckling deformation of cylindrical and
Physical Concepts Types of Observed Behavior and Their Meaning 81

general shells contains rotation from the beginning of the loading process. In
these cases the linearized stability equations have variable coefficients which
must be solved for from the nonlinear equilibrium equations governing the
prebuckling state.
2. Up to now the determination of the buckling load consisted of solving
the linearized stability equations for the critical load, at which the primary
equilibrium path in the load-displacement plane is intersected by a secondary
equilibrium path. It has been shown in advanced texts on stability (see, for
instance, [2.6] and [2.36]) that equilibrium on the primary path becomes
unstable at such point and that structural behavior beyond the bifurcation point
is governed by conditions on the secondary path. There are cases where the
behavior of the structure can only be explained if the shape of the secondary
path is known. Such a knowledge is needed to explain why a flat plate develops
considerable postbuckling strength, for example, but a cylindrical shell under
axial compression buckles abruptly and even explosively.
In the following chapter Koiter’s linearized theory for initial postbuckling
behavior is presented. In Koiter’s work the shape of the secondary equilibrium
path near bifurcation (see Figure 2.34) plays a central role in determining
the influence of initial geometric imperfections. If the initial portion of the
secondary path has a positive slope (like for plates), then the structure can
develop considerable postbuckling strength and loss of stability of the primary
path does not result in structural collapse. However, when the initial portion
of the secondary path has a negative slope (like for cylindrical shells) then
in most cases buckling will occur violently and the magnitude of the critical
load is subject to the degrading influence of initial geometric imperfections.
Unfortunately, the information given by Koiter’s theory is limited to the imme-
diate neighborhood of the bifurcation point. Thus a nonlinear solution must be
carried out if the shape of the secondary equilibrium path in the more advanced
postbuckling region is needed.
3. Finally, in the most general case, when both geometric and/or material nonlin-
earities are included in the analysis, loss of stability occurs at a limit point
rather than at a bifurcation point. In such cases the critical load must be
determined through solution of the nonlinear equations of equilibrium.

In the following, examples illustrating the effects of nonlinear behavior are


considered.

a. Axially Compressed Cylindrical Shells

To solve for the axisymmetric prebuckling deformation shown in Figure 2.35a, one
must specialize the nonlinear equilibrium Eqs. (2.98) for axial symmetry. Assuming
w D hW C hw0 x 2.199
 
Eh2 1
fD  y 2 C R2 f0 x
cR 2
82 The Concepts of Elastic Stability

W D  /c is the uniform increase in radius due to the Poisson’s effect and
where
c D 31  2 , then substitution into Eqs. (2.98a) and (2.98b) yields for axial
compression only p D 0
w0iv C 4cR2 /h2 f000 C 4cR/h w000 D 0 (2.200a)
fi0v  cw000 D 0 (2.200b)
d
where  0 D R  . Equation (2.200b) can be integrated twice yielding
dx
f000 D cw0 C C
Q 1x C C
Q2 2.201
Q1 D C
where C Q 2 D 0 because of the circumferential periodicity condition (see, for
example, [2.64]). Substituting this expression into Eq. (2.200a) one gets
w0iv C 4cR/h w000 C 4c2 R2 /h2 w0 D 0 2.202
where it is assumed that the axial coordinate x is zero at the midpoint of the shell.
If at both edges identical boundary conditions are specified then the prebuckling
displacement must be symmetric about x D 0 and therefore only even functions
are included in the solution of Eq. (2.302) and the boundary conditions need only
to be enforced at x D L/2. Consequently the solution is
w0 x D C1 sinh ˛x sin ˇx C C2 cosh ˛x cos ˇx 2.203
where # #
p c p c
˛D 1 ; ˇ D 1C . 2.204
Rh Rh
For simply supported boundary conditions (w0 D W , w000 D 0 at x D L/2) the
constants of integration become
a1 a2
C1 D ; C2 D
 
 
L L  2
L L
a1 D W 2c/Rh cosh ˛ cos ˇ C 1  sinh ˛ sin ˇ 2.205
2 2 2 2
 
2
L L L L
a2 D W 2c/Rh 1  cosh ˛ cos ˇ  sinh ˛ sin ˇ
2 2 2 2
  
L L
 D 1  2 2c/Rh cosh2 ˛  sin2 ˇ .
2 2
This solution was first obtained by Föppl in 1926 [2.65]. As can be seen in
Figure 2.53 the disturbance due to the restraint at the cylinder edge spreads over
a large part of the cylinder as the axial load increases. When D 1, the displace-
ment pattern becomes purely sinusoidal (˛ D 0) and the lateral displacements grow
without bound.
Notice that thus the axisymmetric collapse load is identical with the critical
buckling load obtained from a bifurcation analysis with membrane prebuckling
Physical Concepts Types of Observed Behavior and Their Meaning 83

Figure 2.53 Prebuckling deformation due to end-constraint

(Eq. 2.110). However, the use of the rigorous prebuckling solution (Eq. 2.203)
may result in a lower eigenvalue. That is bifurcation buckling into asymmetric
modes may occur before the axisymmetric collapse load is reached. The possible
asymmetric buckling modes consists of deformation patterns in which the
lateral displacement varies harmonically in the circumferential direction. Thus if
84 The Concepts of Elastic Stability

one assumes
y
O D hWx
W O cos n 2.206
R
2
O D ERh Fx
f O cos n
y
c R
then upon substitution into the linearized stability equations separation with respect
to the circumferential space variable y is possible. The condition of continuity in
the circumferential direction will be satisfied if n, the number of circumferential
waves, is an integer. Upon substitution of Eqs. (2.203) and (2.206) into Eqs. (2.100)
one obtains
R2 00 R
O iv  2n2 W
W O 00 C n4 W
O C 4c O 00 C n2 w000 F
O C 4c  W
F O
h 2 h
R
C 4c2 n2 w0 WO D0 (2.207a)
h
h
O iv  2n2 F
F O 00 C n4 F O 00  c n2 w000 W
O  cW O D0 (2.207b)
R
where  0 D Rd/dx . These are linear differential equations with variable coef-
ficients. For a given n besides the trivial solution W O DF O D 0, solutions which
satisfy specified boundary conditions exist for particular values of . The lowest
of these values represents the critical load of the cylinder.
All the known solutions of this variable coefficients eigenvalue problem were
obtained by numerical methods. Stein in [2.66] obtained solutions for simply
supported boundary conditions N Ox D N O xy D W
O DW O 00 D 0 using an energy-based
finite difference approach. Independently Fischer in [2.67] presented similar solu-
O x D vO D W
tions for slightly different boundary conditions N O DW O 00 D 0. Almroth
in [2.68] published an extension of Stein and Fischer’s work using finite differ-
ence approximations to solve the stability Eqs. (2.207) for eight different sets
of boundary conditions. As can be seen from Table 2.1, which also includes
results from [2.69] where Hoff and Soong rigorously satisfied the same boundary
conditions for the linearized stability equations but used a membrane prebuckling
solution, the use of rigorous (nonlinear) prebuckling solution may in some cases
result in a significant decrease of the critical buckling load. Notice, however, that
the very low critical buckling loads for the SS1 and SS2 cases are caused not by
the use of the rigorous (nonlinear) prebuckling analysis, but rather by the weak
boundary support in the circumferential direction N O xy D 0 .

b. Bending of Cylinders Ovalization of the Cross-Section

When the stability of circular cylindrical shells under combined bending and axial
compression is discussed in Sub-section 2.1.11, it is explicitly stated there that
the solution is only valid for relatively short shells were the shell edges are held
circular. It is well known that bending of long thin-walled shells induces ovalization
Physical Concepts Types of Observed Behavior and Their Meaning 85

Table 2.1 Influence of prebuckling deformations and boundary conditions


on the axial buckling load (isotropic shell, R/h D 1000,
L/R D 1.0)
Boundary conditions Prebuckling solutions used
Membrane Rigorous
(Ref. 2.69) (Ref. 2.68)
c c n
SS1: Nx D Nxy D w D w,xx D 0 0.5 0.502 2
SS2: u D Nxy D w D w,xx D 0 0.5 0.503 2
SS3: Nx D v D w D w,xx D 0 1.0 0.844 26
SS4: u D v D w D w,xx D 0 1.0 0.867 27
C1: Nx D Nxy D w D w,x D 0 1.0 0.908 26
C2: u D Nxy D w D w,x D 0 1.0 0.926 27
C3: Nx D v D w D w,x D 0 1.0 0.910 26
C4: u D v D w D w,x D 0 1.0 0.926 27

of the cross-section, the so-called Brazier effect [2.70]. In such cases the prebuck-
ling deformation is obviously not rotation free and thus a bifurcation analysis based
on membrane prebuckling is not a rigorous solution.
To take the effect of ovalization into account when discussing stability consider
a long cylindrical shell subjected to a bending moment which causes the curvature
shown in Figure 2.54. Notice that due to this curvature the longitudinal tensile and
compressive stresses will have components directed towards the midplane of the
curved shell. The effect of these components is to flatten the cross-section of the
shell and this flattening results in a decrease of the resistance to the applied bending
moment. Thus, as shown in Figure 2.55, a plot showing the applied bending
moment M versus curvature 0 must have a decreasing slope. This leads to a
maximum (a limit point) and collapse of the shell results.
To calculate the value of the maximum bending moment at the limit point Brazier
considered the shell deformation to occur in two steps. First a long shell is thought
of bent into a circle with large radius b forming a toroidal shell, whereby it is
assumed that the deformation that occurs during this process is described by St.
Venant’s theory of bending. Next the cross-section is allowed to assume additional
displacements vO , wO so directed that the applied forces do no work. Thus these
additional displacements can be calculated from the condition that the strain energy
on the deformed body must be a minimum.
Considering a cross-section of the toroidal shell with a moving coordinate system
as shown in Figure 2.54 (notice that here w is positive inward) and assuming that
the displacement of the centerline of the cylinder is equal to zero, then St. Venant’s
(linear) solution of the bending problem yields the following displacements
v0 D  21 0 a2 sin 2.208
w0 D 21 0 a2 cos .
For a thin-walled shell the work due to shell-wall bending in the axial direction
may be omitted in comparison with the membrane strain energy. Hence the total
86 The Concepts of Elastic Stability

Figure 2.54 Bent cylindrical shell element notation and sign convention

Figure 2.55 Equilibrium paths for bent cylindrical shell


Physical Concepts Types of Observed Behavior and Their Meaning 87

strain energy per unit length is


 2  2
Eah Da
UD ε2x d C  2 d 2.209
2 0 2 0
where
εx D 0 [a  w cos  v sin ] 2.210
1
 D v, C w, . 2.211
a2
Using the assumptions that
v D v0 C vO 2.212
w D w0 C wO
and that the incremental field is inextensional
1
ε D Ov,  w O D0 2.213
a
one can express both εx and  as a function of v only. If, in addition, one uses
Brazier’s assumption that vO , wO − a then the total strain energy per unit length of
the toroid (Eq. (2.209)) becomes
 
Eah 2
UD 02 [a2  2av, C 0 a2 cos ] cos2  av sin 2
2 0
 2  2.214
h2 1
C v, C v,  d .
121  2  a2
The variational equation υU D 0 yields the following Euler equation
vvi C 2viv C v00 D N sin 2 2.215
where  0, and N D 18 02 a5 1  v2 /h2 . The general solution of this linear,
inhomogeneous ordinary differential equation can readily be found. The constants
of integration are evaluated by symmetry and continuity considerations in the
circumferential direction and by discarding of rigid body displacements. The final
solution is
N
vD sin 2 2.216a
36
N
wD cos 2 C 0 a2 cos . 2.216b
18
Notice that the term containing  is very small. If contains St. Venant’s displace-
ments (Eq. (2.208)) plus a rigid body motion which is immaterial. Substituting
these expressions into Eq. (2.214) and carrying out the integrals one obtains
 
3
2 Ea h 3 2 a4 2
U D 0 1  0 2 1    2.217
2 4 h
88 The Concepts of Elastic Stability

By Castigliano’s Second Theorem the bending moment acting at a section of the


toroid is  
υU 3 3 3 a4 2
MD D Ea h 0  0 2 1    . 2.218
∂0 2 h

The maximum value of M occurs when ∂M/∂0 D 0, that is when


2 h2 1
02 D c2 D . 2.219
9 a4 1   2
Thus at the limit point the maximum value of bending moment is
p
2 2 Eh2 a
Mc D p 2.220
9 1  2
Substituting Eq. (2.219) into Eq. (2.216) one obtains
1
vD a sin 2 2.221a
9
 # 
2  9h
w D a cos 2 C p cos . 2.221b
9 1  2 2 a
The form of the cross-section at this point is shown in Figure 2.56.
In recent years, thanks to considerable research effort sponsored mainly by the
off-shore industry, much has been learned about the elastic-plastic response and
the various instabilities which govern the behavior of circular shells under bending.
Some of this is discussed in Chapters 9 and 16, Volume 2 and more information
can be found in recent papers by Kyriakides and his co-workers ([2.71] and [2.72]).

c. Plastic Buckling

Up to now in all cases discussed it is assumed that instability occurs before any of
the fibers in the structure reaches the yield stress of the material. This assumption

Figure 2.56 Cross-sectional shape immediately before buckling (from [2.70])


Physical Concepts Types of Observed Behavior and Their Meaning 89

is valid for sufficiently slender structural elements, where at least one dimension is
relatively small in comparison with the others. For thicker components instability
may occur at load levels where at least in parts of the structure stresses do exceed
the proportional limit. In such cases the stability analysis must be based on the
nonlinear material behavior depicted in the stress-strain diagram of Figure 2.57.
One feature that causes a lot of difficulties is the fact that the unloading of the
material follows a different path than the loading. Thus, as seen in Figure 2.57,
if the material is stressed to a value of 0 and a positive (loading) increment of
strain υε is imposed, the resulting increment in stress is
υ D Et υε
where Et is the slope of the tangent to the stress-strain curve at the point under
consideration. For a negative (unloading) increment of strain the increment in
stress is
υ D Eυε
where E is Young’s modulus, the slope of the unloading curve (which is identical
to the slope of the linear elastic portion of the loading curve).
The inelastic buckling of columns under axial compression has been the subject
of extensive theoretical and experimental investigations for over a century. Thus
it can be used conveniently to illustrate the steps involved in the determination of
inelastic (or plastic) buckling loads.
One model of buckling is based on the assumption that the equilibrium of a
straight column becomes unstable when under the same axial load there are adja-
cent equilibrium positions infinitesimally close to the straight equilibrium position.
Then as the column undergoes a small lateral displacement w, the stresses on the
concave side increase according to the constitutive law of the compressive stress-
strain diagram (see Figure 2.57), whereas the stresses on the convex side decrease
according to Hooke’s law. If one assumes that after bending the cross-sections
remain plane and normal to the center line of the column, then one obtains the
stress distribution shown in Figure 2.58b. Notice that at every cross-section there
is a straight line, the axis of average stress, along which the stress 0 remains
unchanged.

Figure 2.57 Typical compressive stress-strain diagram


90 The Concepts of Elastic Stability

Figure 2.58 Inelastic buckling of a column notation and sign convention

After buckling, at any cross-section the moment of the stresses about the
centroidal axis of the cross-section must equal the moment of the applied load about
the same axis. Since the moment of the average stresses 0 about the centroidal
axis is zero, one can express this condition as (see [2.1], [2.7] and [2.73])
 
Eε  ε0 z dA C Et ε  ε0 z dA D Pw 2.222
A1 A2

where A1 and A2 are the parts of the cross-sectional area on the two sides of the
axis of average stress, which are subjected to a decrease and an increase of the
average compressive stress 0 , respectively. It is easy to show (see, for instance,
[2.1], p. 164) that
d2 w
ε  ε0 D z 2 . 2.223
dx
Thus upon substitution and regrouping Eq. (2.222) can be written as
d2 w
Er I C Pw D 0 2.224
dx 2
where
Er D EI1 C Et I2 /I 2.225
and  
I1 D z2 dA; I2 D z2 dA. 2.226
A1 A2

The last two integrals are the moments of inertia of A1 and A2 with respect to
the axis of average stress. In order to evaluate I1 and I2 it is necessary to locate
Physical Concepts Types of Observed Behavior and Their Meaning 91

the axis of average stress. This can be done by recalling that at any cross-section
the resultant of the stress distribution must be equal to the applied load. Since, by
definition, the average stress 0 D P/A, therefore this condition implies that
 
Eε  ε0  dA C Et ε  ε0  dA D 0. 2.227
A1 A2

By substituting ε  ε0  for from Eq. (2.223) and regrouping one obtains


 
E z dA C Et z dA D 0. 2.228
A1 A2

But these integrals are, respectively, the negative of the first moment of A1 and
the first moment of A2 with respect to the axis of average stress. Denoting these
quantities by S1 and S2 , respectively, Eq. (2.228) becomes
ES1 D Et S2 . 2.229
This expression can be used to locate the axis of average stress for a given average
stress 0 at which the tangent modulus is Et . Once this is done, the values of I1
and I2 can be computed from Eq. (2.226) and from Eq. (2.225) one can evaluate
the “reduced modulus” Er as a function of E and Et . Notice that Er is not only
a function of the average stress 0 (because of Et ) but also of the shape of the
cross-section. See [2.1] and [2.73] for sample calculations.
The solution of Eq. (2.224) gives the critical inelastic buckling load for a simply
supported column as
I
Pc D 2 Er 2 . 2.230
L
An expression that is similar to the Euler buckling load (compare with Eq. (2.3))
with the reduced modulus Er replacing the modulus of elasticity E.
The “reduced modulus” or “double-modulus” theory is often called the
Considère Engesser von Kármán theory after the scientists who around the turn
of the century had first proposed and developed it (see [2.74] [2.76]).
A second possible model of buckling is based on the assumption that the bending
of an initially straight column will begin as soon as the tangent modulus load is
exceeded. Assuming further, that at least initially, the straight column will start
to deflect laterally under increasing loading, Shanley in 1947 [2.11]) by way of a
simple model and careful experimentation clarified the significance of Engesser’s
tangent modulus load [2.75] and the role of the reduced modulus load of von
Kármán. It is now accepted that the tangent modulus load
I
Pc D 2 Et 2.231
L2
is the lowest possible bifurcation load at which the straight configuration loses its
uniqueness but not its stability.
Experimental determination of the inelastic buckling load shows that the
maximum column load will lie somewhere between the loads predicted by the
92 The Concepts of Elastic Stability

tangent modulus and the reduced modulus formulations. These two values can
thus be considered as the lower and the upper bounds of the critical inelastic
buckling load for axially compressed columns.
The uniaxial elastic-strain hardening plastic behavior of most structural materials
used in aerospace and off-shore applications is adequately described by the simple
three parameter strain-stress relationship proposed by Ramberg and Osgood [2.77];
   n
εD Cˇ 2.232
E E
where the parameters E, ˇ and n are material constants which must be obtained
by tests. Notice that for  D 0, Eq. (2.232) yields d/dε D E. This slope is equal
to the modulus of elasticity at the origin and it can be obtained from the experi-
mental curve in the usual manner. The remaining two parameters are determined
by requiring the empirical curve, given by Eq. (2.232), to coincide with the exper-
imental curve at secant moduli of 0.7 E and 0.85 E. Notice that above the propor-
tional limit the secant modulus is defined as the ratio stress divided by strain.
Recalling that when the stress is equal to 0.7 the strain is equal to 0.7 /E, and
when the stress is equal to 0.85 the strain is equal to 0.85 /E, then by substituting
in turn the coordinates of these points into Eq. (2.232) one obtains
 
3 E n1
ˇD 2.233
7 0.7
and
ln17/7
nD1C 2.234
ln0.7 /0.85 
respectively. Finally, one can calculate the tangent modulus Et from Eq. (2.232)
yielding
d E
Et D . 2.235
dε 3n n1
1C 0.7 /0.85 
7
The three parameters of the Ramberg Osgood method E, 0.7 , n are tabulated
for a wide variety of materials in [78].
As a further possible mode of buckling let us consider the rigid-perfectly plastic
behavior of the column with rectangular cross-section b ð h (b > h) depicted in
Figure 2.59. Under axial load a plastic hinge forms at the central cross-section. The
assumed stress distribution shown in Figure 2.60 can be interpreted as follows, the
central portion of width h, carries the axial load P while the outer portions provide
the reduced plastic moment MIP . Thus
P D y bh1 2.236
1
M0p D y bh2  h12  2.237
4
Eliminating h1 between these equations one obtains
M0p D Mp 1  P2 /Pp2  2.238
Physical Concepts Types of Observed Behavior and Their Meaning 93

Figure 2.59 Plastic mechanism of a simply supported column

Figure 2.60 Stress distribution at the plastic hinge

where
Mp D 41 y bh2 full plastic moment of the cross-section 2.239a
Pp D y bh squash load of the cross-section. 2.239b
From equilibrium considerations at the plastic hinge (see Figure 2.59)
M0p D Pwc . 2.240

Eliminating M0p between Eqs. (2.238) and (2.240) one obtains


 
  
2wc 2 2wc 
P D Pp 1C  . 2.241
 h h 

This equation gives the load carrying capacity of an axially compressed column
once a plastic hinge has formed. It is plotted in Figure 2.61 together with the elastic
94 The Concepts of Elastic Stability

Figure 2.61 Collapse behavior of an axially compressed column

response curve of a column with a half-wave sine initial imperfection of amplitude


hN1 . In addition the figure displays the actual response of the real column, which
begins to depart from the elastic curve at point B, when the first fiber reaches the
yield stress. The real column attains its maximum load carrying capacity at a limit
point, after which it decreases and approaches the theoretical rigid-plastic curve
asymptotically.
When using the rigid-perfectly plastic approximation it is assumed that there is
no strain hardening, since only in this case are the plastic hinges concentrated in a
very short length of the column. If strain hardening is of major concern then one
must rely on Hill’s bifurcation criterion for elastic-plastic solids (see [2.79] and
[2.80]). For an extensive review on plastic buckling the interested reader should
consult [2.81] by Sewell.

2.2 Mathematical Models for Perfect Structures


The applications presented so far serve a dual purpose. First, the reader is intro-
duced by means of relatively simple examples to the concept of structural stability.
Secondly, he acquires basic skills to solve those buckling problems that occur
frequently in practice.
When investigating the stability behavior of a structure under a given load one
is really concerned whether the corresponding equilibrium configuration is stable
or unstable. Thus at first, the analyst must solve for the equilibrium configuration
and then investigate whether the equilibrium state found is stable or unstable.
Referring to the load-displacement curves shown in Figure 2.34 each point on
a path represents an equilibrium position of the structure. From the form of the
curves it is obvious that the governing equilibrium equations are nonlinear. At
parts of the load-displacement curves the equilibrium is stable, at other parts
it is unstable. The critical load is defined as the smallest load at which the
Mathematical Models for Perfect Structures 95

equilibrium of the structure fails to be stable as the load is slowly increased


from zero.
The critical load may occur at the limit point of the fundamental equilibrium path,
that is, at the point where the load is a relative maximum. Another possibility for
reaching the critical load occurs when the primary (or fundamental) path emanating
from the origin is intersected by a secondary equilibrium path. At the point of
intersection, the so-called bifurcation point the equilibrium equations have multiple
solutions, one corresponding to each branch.
Thus the structural analyst must, in principle, always deal with two sets of
equations, one which governs equilibrium and the second which yields information
about the stability behavior. The equilibrium equations are often nonlinear, whereas
in most cases the stability equations used are linearized.
The fact that for stability investigations one must rely on nonlinear equilibrium
equations is due to the concept of instability used, namely that at the critical load
more than one equilibrium position exists. Since for linear theory of elasticity there
is a uniqueness proof (that is, for a given load there is one and only one solution),
obviously one cannot base the derivation of stability analysis on it.
In general, the nonlinear equilibrium equations can be derived either by estab-
lishing the equilibrium of forces and moments on a slightly deformed element, or
by using the stationary potential energy criterion (see, for example, [2.1] and [2.2]).
On the other hand, the linearized stability equations can be obtained either by the
method of adjacent equilibrium, or by the minimum potential energy criterion. In
the following the different approaches shall be illustrated by examples.

2.2.1 Static Versus Kinematic Approach

The energy criteria of equilibrium and stability, which state that a conservative
system is in equilibrium if its total potential energy is stationary, and the equilib-
rium is stable if its total potential energy is a minimum, is applied in the following
to finding the critical load of the prismatic column subjected to compressive end
loads P shown in Figure 2.1. Its total potential energy may be written (see, for
instance, [2.2])
 D U m C U b C p 2.242

where

EA L 2
Um D extensional energy D ε dx 2.243a
2 0 x

EI L 2
Ub D bending energy D  dx 2.243b
2 0 x
p D potential of the applied load D P[uL  uo]
 L
DP u,x dx 2.243c
0
96 The Concepts of Elastic Stability

and
2
εx D extensional strain of the centroidal axis D u,x C 21 w,x 2.244a
x D curvature of the centroidal axis D w,xx . 2.244b

The requirement that if the column is in equilibrium its total potential energy
must assume a stationary value yields the following variational problem:
 L
υ D υUm C Ub C p  D υ Fx, u, u,x , w, w,x , w,xx  dx D 0. 2.245
0

The condition that υ D 0 implies that the integrand (the functional F) in


Eq. (2.245) must satisfy the corresponding Euler equations of the calculus of
variation, which in this case are
∂F d ∂F
 D0 2.246a
∂u dx ∂u,x
∂F d ∂F d2 ∂F
 C 2 D 0. 2.246b
∂w dx ∂w,x dx ∂w,xx
Calculating the required partial derivatives, substitution and regrouping yields
d d
EAεx  D N D 0 2.247a
dx dx
d2 d
EIw,xx   Nw,x  D 0. 2.247b
dx 2 dx
The first of these equations can be integrated yielding N D constant D P. The
second equation becomes then for EI D constant

Elw,xxxx C Pw,xx D 0. 2.248

This is the equilibrium equation of an axially compressed perfectly straight


column. For any w that satisfies this equation and the specified boundary conditions
at x D 0 and x D L, the total potential energy  is stationary. Whether  is also
a relative minimum (that is  > 0) will next be investigated.
The character of the total potential energy  for a given equilibrium configu-
ration may be determined by examination of the change in total potential energy
 corresponding to an arbitrary infinitesimal virtual displacement of the struc-
ture from the given equilibrium position. In terms of a Taylor series expansion the
change in the total potential energy is
1 2 1
 D υ C υ  C υ3  C Ð Ð Ð 2.249
2! 3!
where the terms on the right are linear, quadratic, etc., respectively, in the infinites-
imal virtual displacements. Recalling that the first-order term vanishes identically
for equilibrium configurations, hence the sign of  is governed by the sign of
Mathematical Models for Perfect Structures 97

the second variation. For sufficiently small values of the applied load, it can be
shown that the second variation is positive definite. The critical load is defined as
the smallest load for which the second variation no longer is positive definite.
To obtain the expression for υ2  one assumes in Eqs. (2.242) (2.244) that

u D u0 C uO ; w D w0 C wO 2.250

where u0 , w0 denote the configuration whose stability is under investigation and uO ,


wO are infinitesimally small increments. Substituting into Eq. (2.242) and regrouping
yields the following expression for the second variation
    
1 2 1 L 2 2 3 2 2 2
υ D EA uO ,x C u0,x wO ,x C 2w0,x uO ,x wO ,x C w0,x wO ,x C EIw,
O xx dx
2 2 0 2
2.251
Recalling that for the undeflected form of the column
P
u0 D  x; w0 D 0 2.252
EA
hence Eq. (2.251) becomes
 L
υ2  D 2
fEAOu,x 2
C EIwO ,xx 2
 PwO ,x g dx. 2.253
0

This quadratic form is seen to be positive definite for sufficiently small values
of the applied load P. The critical value of P is the smallest load for which the
definite integral ceases to be positive definite.
The criterion for the limit of positive-definiteness for a continuous system is
attributed to Trefftz [2.82]. Considering Eq. (2.253), for a small value of P, υ2  >
O For large values of P, υ2  < 0 for some variations
0 for all nonzero variations uO , w.
O As P is increased from zero, a value is reached (say, P D Pc ) at which υ2 
uO , w.
is for the first time zero for at least one variation uO , w.O It is still positive for all
O Thus for P D Pc , υ2  assumes a stationary value for the
other variations uO , w.
particular set of variations uO , w.
O Then

υυ2  D 0. 2.254

Hence, on the basis of Trefftz criterion, the stability equations for the critical
load are given by the Euler equations for the functional in the second-variation
expression. For a functional of the form of the integrand in Eq. (2.253) the Euler
equations are given by Eq. (2.246), where
2 2 2
F D EAOu,x C EIwO ,xx  PwO ,x . 2.255

Calculating the required partial derivatives, substitution and regrouping yields


uO ,xx D 0 (2.256a)
EIwO ,xxxx C PwO ,xx D 0. (2.256b)
98 The Concepts of Elastic Stability

These are the uncoupled stability equations of the axially compressed column.
The variational approach yields also the natural boundary conditions that must be
satisfied in order for υυ2  D 0 to hold. Thus at x D 0, L
O D EAOu,x D 0 or
either N υOu D 0 2.257a
O D EIwO ,xx D 0 or υwO ,x D 0
either M 2.257b
either ElwO ,xxx C PwO ,x D 0 or υwO D 0 2.257a
The general solution of Eq. (2.256b) is
wO D C1 sin kx C C2 cos kx C C3 x C C4 2.258
where k 2 D P/EI. This solution must satisfy for the clamped-free column of
Figure 2.1 the following boundary conditions
at x D 0 : wO D wO ,x D 0 2.259a
at x D L : wO ,xx D 0; wO ,xxx C k 2 wO ,x D 0. 2.259b
This requirement leads to four homogeneous algebraic equations in the four
constants C1 , . . ., C4 . For a nontrivial solution to exist (that is, all four constants
are not identically equal to zero) the determinant of the coefficients of the Ci0 s must
vanish. Expansion of the determinant yields the following characteristic equation
4 EI
cos kL D 0 ! Pn D 2n  12 n D 1, 2, . . . . 2.260
4 L2
The smallest buckling load occurs for n D 1. Thus
2 EI
Pc D 2.261
4 L2
the value found by Euler in 1744. The buckling mode is
 x 
wO D C2 cos  1 . 2.262
2L
The fact that for the column buckling both the equilibrium Eq. (2.248) and the
stability Eq. (2.256b) are identical is an exception. These equations are usually
different. Notice also that the form of the stability equations depends on the
prebuckling (undeflected) solution used. This point will be discussed further in
Section 2.2.4b.
Turning now to the problem depicted in Figure 2.62, where the compressive load
P at the free end does not remain fixed in its direction but follows the deformation
of the body in some manner, then the work done by the end load P in reaching
the final position is path dependent and one is dealing with a nonconservative
force. That in such cases a stability based on the energy criteria may fail to yield
the correct answer can easily be demonstrated for the present case. The boundary
conditions at the free end now are
at x D L : wO ,xx D wO ,xxx D 0. 2.263
Mathematical Models for Perfect Structures 99

Figure 2.62 Column loaded by a follower force

Applying the boundary conditions specified by Eqs. (2.259a) and (2.263) to the
general solution for wO given by Eq. (2.258) yields the following system of linear
equations   
sin kL cos kL C1
D 0. 2.264
 cos kL sin kL C2
But here the determinant of the coefficient matrix is not equal to zero since
sin2 kL C cos2 kL D 1. 2.265
Hence the only admissible solution is the trivial one, namely C1 D C2 D C3 D
C4 D 0. This would imply that in this case the column does not buckle. This is
obviously incorrect.
For nonconservative problems one must always use the kinetic approach, where
one starts with the equations governing small free vibrations of the elastic structure
at some level of the external loading (treated as a fixed quantity) and then tries to
find at what level of the external loading the free vibrations cease to be bounded
in the small.
If  denotes the mass per unit volume then the equation of motion of the column
depicted in Figure 2.62 under a constant axial load P is
w,xxxx C k 2 w,xx D w
N ,tt 2.266
where
p PA
K2 D ; N D . 2.267
EI EI
Using separation of variables
Wx, t D Wx eiωt 2.268
one obtains
W,xxxx C k 2 W,xx  ω
N 2W D 0 2.269
an ordinary differential equation with constant coefficients, whose solution can be
written as
Wx D C1 sin ˛x C C2 cos ˛x C C3 sinh ˇx C C4 cosh ˇx 2.270
100 The Concepts of Elastic Stability

where
$ 
%
%
k & N 2
4ω
˛D p 1C 1C 4 2.271a
2 k
$ 
%
%
k & N 2
4ω
ˇD p 1 C 1 C 4 . 2.271b
2 k
Applying the boundary conditions specified by Eqs. (2.259a) and (2.263) one
obtains the following characteristic equation
˛4 C ˇ4   2˛2 ˇ2 cos ˛L C cosh ˇL C ˛ˇˇ2  ˛2  sin ˛L sinh ˇL D 0
2.272
But now from Eqs. (2.271a) and (2.271b)
˛4 C ˇ4 D k 4 C 2ω
N 2
˛2 ˇ2 D ω2 N 2.273
ˇ2  ˛2 D k 2
Substituting into Eq. (2.272) and regrouping yields
2 C  sin ˛L sinh ˇL C 22 1 C cos ˛L cosh ˇL D 0 2.274
where
P EI L4
D ; PE D  2 ; 2 D ω2 N
PE L2 4
$ 
# %%  2
& 
˛L D  1C 1C4 2.275
2
$ 
# %%  2
& 
ˇL D  1 C 1 C 4 .
2
The transcendental Eq. (2.274) can be solved repeatedly for the frequencies 1
and 2 by assigning different positive values to , starting from zero. From the
results plotted in Figure 2.63 one sees that the unloaded natural frequencies (and
the corresponding eigenfunctions) of the column change with increasing loading
. It is also clear that as long as < 2.0316 (the load at which the two frequencies
1 and 2 coalesce) the motion is oscillatory and hence stable. For > 2.0316 the
frequencies become a complex conjugate pair. From Eq. (2.268) it is evident that
the negative imaginary part results in unbounded oscillation and is hence unstable.
Thus the critical value for the follower force P shown in Figure 2.62, also
called Beck’s problem who was first to obtain the correct solution (see [2.83]), is
D P/PE D 2.0316, or
EI
Pc D 2.03162 2 . 2.276
L
Mathematical Models for Perfect Structures 101

Figure 2.63 Characteristic curve for the cantilever column loaded by a follower force

A comparison with Euler’s solution for a fixed load given by Eq. (2.1) indicates
that the same column can carry about an eight times larger follower load. For
more information about the stability of nonconservative structural configurations
the interested reader should consult [2.84] and [2.85].

2.2.2 Approximate Solutions of Bifurcation Problems

In most of the cases considered up to now the finding of the critical buckling load
has been reduced to the solution of a linearized eigenvalue problem. In general the
differential equations involved can be written as

Lw  Mw D 0 2.277

where L and M are linear, homogeneous differential operators of order 2p and


2q respectively, with p > q. Any solution w must satisfy Eq. 2.277 at every point
of the region R. Associated with the differential equation there are p boundary
conditions that the function w must satisfy at every point of the boundary C of the
region R. The boundary conditions are of the type

Bi w D 0 i D 1, 2, . . . , p 2.278

where the Bi are linear, homogeneous differential operators involving derivatives


normal to the boundary and along the boundary through order 2p  1.
The eigenvalue problem consists of finding the values of the parameter , for
which there are nonvanishing functions w which satisfy the differential Eq. (2.277)
and the boundary conditions specified by Eq. (2.278). Such parameters are called
eigenvalues (say, buckling loads) and the corresponding functions are called eigen-
functions (say, buckling modes).
Unfortunately, in general, the solution of the eigenvalue problem for continuous
systems is not a straightforward matter. Exact solutions have been found only for
uniform systems with relatively simple boundary conditions. In other cases one
must rely on approximate solutions.
102 The Concepts of Elastic Stability

Before turning to the presentation of the different methods that are available to
obtain approximate solutions, one may want to summarize those definitions that
are frequently used in the literature dealing with eigenvalue problems.
When speaking of trial functions one often distinguishes the following classes
(see [2.86]).
1. Admissible functions: These are arbitrary functions which satisfy all the
geometric boundary conditions of the eigenvalue problem and are 2p times
differentiable over the region R.
2. Comparison functions: These are arbitrary functions which satisfy all boundary
conditions (geometric and natural) and are 2p times differentiable over the
region R.
3. Eigenfunctions: These are the solutions one is trying to obtain which, of course,
satisfy all boundary conditions (geometric and natural) and the differential
equation of the eigenvalue problem.

The eigenvalue problem defined by Eqs. (2.277) and (2.278) is said to be self-
adjoint if, for any two arbitrary admissible or comparison functions w1 and w2 , the
statements
 
w1 Lw2  dR D w2 Lw1  dR 2.279a
R R
 
w1 Mw2  dR D w2 Mw1  dR 2.279b
R R

hold true. Whether a specified system is self-adjoint or not can be established by


means of integration by parts.
Further, if for any such comparison function w

wLw dR ½ 0 2.280
R

the operator L is said to be positive. The operator L is said to be positive definite


if the integral is zero only if w is identically zero. There is a similar definition with
respect to the operator M. If both L and M are positive definite, the eigenvalue
problem is said to be positive definite, in which case all eigenvalues i are positive.
For further details about the nature of the different types of eigenvalue problems
the interested reader may consult [2.24] and [2.86].

a. The Rayleigh Ritz Method

One of the approaches that can be used to obtain an approximation for the critical
buckling load of a structure without having to derive and solve the linearized
stability equations is the Rayleigh Ritz method. Its use is based on the Trefftz
Mathematical Models for Perfect Structures 103

criterion, which defines the critical load as the smallest load for which the second
variation of the potential energy assumes a stationary value.
To apply this method one assumes a solution in the form of a linear combination
of trial functions wi , which satisfy at least all the geometric boundary conditions
of the problem. Hence
n
wn D ai wi 2.281
iD1

where the wi are known, linearly independent functions of the spatial coordinates
over the region R and the ai are unknown coefficients to be determined. This
assumed form wn is substituted into the second variation of the potential energy
υ2  of the problem. After carrying out the integrals involved, the coefficients ai are
determined so as to render the expression for the second variation of the potential
energy υ2  stationary. The necessary condition for this to occur is that
∂ 2 ∂ 2 ∂ 2
υυ2  D υ υa1 C υ υa2 C Ð Ð Ð C υ υan D 0. 2.282
∂a1 ∂a2 ∂an
Since the variations υa1 , υa2 , . . ., υan are arbitrary nonzero quantities this condition
is satisfied if and only if
∂ 2
υ  D 0 i D 1, 2, . . . , n 2.283
∂ai
a set of homogeneous, linear algebraic equations. The simultaneous solution of
these equations constitutes a matrix eigenvalue problem which can be solved easily
by standard methods.
As an illustration of the application of the Rayleigh Ritz method to buckling
load calculations, consider the problem depicted in Figure 2.64 involving the effect
of non-uniform in-plane compressive loading on the stability of a simply supported
rectangular thin flat plate of width b and length a where
 y'  y'
Nx0 D  N0 C N1  N0  sin  D N1 ˇ C 1  ˇ sin  2.284
b b
and ˇ D N0 /N1 . Notice that ˇ D 1 implies uniform loading.
Such non-uniform in-plane loading is typical of flight vehicles exposed to thermal
heating. Thermal stresses, induced by non-uniform temperature fields acting on the

Figure 2.64 Plate subjected to non-uniform in-plane compressive loading


104 The Concepts of Elastic Stability

Figure 2.65 Distribution of in-plane stresses for ˇ D 1.75

structure, are self-equilibrating since they are not the result of externally applied
loads. Therefore the stress distribution over a cross-section of the structure must
have compressive as well as tensile stresses. Notice that for ˇ D 1.75 Eq. 2.284
yields the in-plane stress distribution shown in Figure 2.65, which closely approx-
imates the distribution of thermal stresses over the width of a panel considered
in [2.87].
For this case the second variation of the total potential energy is found to be
(see, for instance, p. 93 of [2.2])
  
1 2 C 2 2 1
υ D O x C vO,y C 2u,
u, O x vO,y C O y C vO,x  dx dy
u, 2
2 2 2
RO

1 2 2
C fNx0 wO ,x C Ny0 wO ,y C 2Nxy0 wO ,x wO ,y g dx dy
2
RO

D 2 2 2
C fwO ,xx C wO ,yy C 2wO ,xx wO ,yy C 21  wO ,xy g dx dy 2.285
2
RO

Since in this case the in-plane and the out-of-plane stability equations are uncou-
pled, and guided by the results for the buckling of a rectangular plate under uniform
axial loading (see Eq. (2.6)) one can obtain a solution by assuming the following
displacement functions (see [2.87])

uO D vO D 0 2.286
x y x y
wO D C11 sin i sin  C C13 sin i sin 3
a b a b
which satisfy the displacement boundary conditions of the problem

wO D wO ,xx D 0 at x D 0, a
2.287
wO D wO ,yy D 0 at y D 0, b

and hence are admissible.


Substituting these functions into the second variation of the total potential energy,
Eq. (2.285), which includes the prebuckling resultants (Nx0 from Eq. (2.284) and
Ny0 D Nxy0 D 0), one obtains after carrying out the integrals involved and some
Mathematical Models for Perfect Structures 105

regrouping the following expression


  
2   
 2   2  2
2 
ab i  2 i 3
υ2  D D C211 C C C213 C
4  a b a b 
   
ab i 2  2 2
' ab i 2
 N0 C11 C C13 C N1  N0 
4 a 4 a
 
8 1 2 9
ð  C211 C C11 C13  C213 2.288
 3 15 35
where D D Eh3 /121  2 .
By Trefftz’s criterion instability occurs whenever υυ2  D 0 . Assuming that
i D a/b and minimization with respect to the free parameters C11 and C13 yields
the following set of algebraic equations
    
 2 8 8
4D  N1 ˇ C 1  ˇ C11 C 1  ˇC13 D 0 (2.289a)
b 3 15
   2  
8 72
N1 1  ˇC11 C 100D  N1 ˇ C 1  ˇ C13 D 0 (2.289b)
15 b 15
where ˇ D N0 /N1 . These homogeneous equations constitute a standard eigenvalue
problem. Notice that for a single mode solution (if C13 D 0) the eigenvalue is
  2
N1 D k c D 2.290a
b
where
4
kc D . 2.290b
8
ˇC 1  ˇ
3
For the two mode solution expansion of the stability determinant yields the
following characteristic equation
  2   4
aN21 C bD N1 C cD2 D0 2.291
b b
where
    
8 72 8 2
aD ˇC 1  ˇ ˇ C 1  ˇ  1  ˇ2 2.292a
3 15 15
    
8 72
b D  100 ˇ C 1  ˇ C 4 ˇ C 1  ˇ 2.292b
3 15
c D 400. 2.292c
The critical buckling load is the smaller of the two roots. Hence
  2
N1 D k c D 2.293a
b
106 The Concepts of Elastic Stability

Figure 2.66 Buckling coefficients for non-uniform in-plane compressive loading

where p
b  b2  4ac
kc D . 2.293b
2a
Either Eqs. (2.290b) or (2.293b) can be used to calculate the buckling coefficient kc
for a simply supported rectangular plate loaded by a non-uniform in-plane compres-
sive loading. As can be seen from Figure 2.66 little improvement is obtained by
O Both equations yield kc D 4 for
using the second term in the assumed solution for w.
ˇ D 1, which agrees with the previously obtained result for a long simply supported
plate under uniform in-plane compression (see Eq. (2.7) and Figure 2.4).
To obtain the value for ˇ for which the resultant compressive load acting on
the plate is zero, one integrates the value of Nx0 given by Eq. (2.284) from y D 0
to y D b and sets the resulting force equal to zero. This yields the value of ˇ D
2  2 , which gives a buckling coefficient of kc D 6.8496 if one uses the
single mode solution, and a buckling coefficient of kc D 6.6519 if the two mode
solution is employed.

b. Galerkin’s Method

To obtain an approximate solution of an eigenvalue problem one can also employ


Galerkin’s method, so named after the Russian naval engineer who first proposed
it in 1915 (see [2.88]). In this method one attempts to find an approximate solution
of the governing differential equation directly. This is done by assuming a solution
in the form of a series of comparison functions

n
wn D ai w i 2.294
iD1

where the wi are known, linearly independent functions which satisfy all the
boundary conditions and are 2p times differentiable, whereas the ai are unknown
Mathematical Models for Perfect Structures 107

coefficients to be determined. In general the series solution will not satisfy the
differential equation defining the eigenvalue problem unless, by some coincidence,
the assumed series solution is composed of the eigenfunctions of the problem. Thus
upon substitution of the assumed solution in the differential equation
Lw  Mw D 0 2.295
an “error εn ” will be obtained so that
εn D Lwn   ^n Mwn  2.296
where ^n is the corresponding estimate of the eigenvalue . At this point one
requires that the “weighted error εn ” integrated over the region R be zero. As
weighting functions one uses the n comparison functions wi . These conditions can
be written as follows


εn wn  dR D 0 j D 1, 2, . . . , n. 2.297
∂aj
R

Consider now
 n  n

wn Lwn  dR D ai wj Lwi  dR D kij ai j D 1, 2, . . . , n
∂aj iD1 iD1
R R
2.298
where the coefficients kij are given by

kij D kji D wj Lwi  dR 2.299
R

and are symmetric if the operator L is self-adjoint. Similarly one can write
 n  n

wn Mwn  dR D ai wj Mwi  dR D mij ai j D 1, 2, . . . , n
∂aj iD1 iD1
R R
2.300
where the coefficients mij are given by

mij D mji D wj Mwi  dR 2.301
R

and are symmetric if the operator M is self-adjoint.


With Eqs. (2.297) through (2.301) one can reduce the solution of the original
continuous eigenvalue problem specified by (Eq. 2.295) to the following system
of n simultaneous equations

N
kij  ^n mij ai D 0 j D 1, 2, . . . , n 2.302
iD1
108 The Concepts of Elastic Stability

which are known as Galerkin’s equations. They represent a matrix eigenvalue


problem for an n-degree-of-freedom system which can be solved easily by standard
techniques.
To illustrate the use of Galerkin’s method for buckling load calculations consider
the case of an axially compressed imperfect cylindrical shell. If w is positive
outward and
x
wN D hN1 cos i 2.303
L
where i is an integer, then the nonlinear Eqs. (2.98) become (see [2.2])
1
Dr4 w C f,xx  ff,yy w,xx C wN ,xx   2f,xy w,xy C f,xx w,yy g D p (2.304a)
R
 
4 2 1
r f  Eh w,xy  w,xx C wN ,xx w,xy C w,xx D 0. (2.304b)
R
and the linearized stability Eqs. (2.100) assumes the form
1O O D0
Dr4 wO C f,xx  LNL f0 , w
O  LNL w0 C w,
N f 2.305a
R
O  Eh wO ,xx C EhLNL w0 C w,
r4 f N w
O D 0. 2.305b
R
Since the loading, the boundary conditions and the initial imperfection (see
Eq. (2.303)) are axisymmetric, therefore the prebuckling solution will also be
axisymmetric, namely
v
w0 D h C wŁ x 2.306a
c
Eh2 1 2
f0 D  y C fŁ x 2.306b
cR 2
where the term  /c represents the Poisson’s expansion and is needed to satisfy
the circumferential periodicity condition (see [2.64]). A substitution of Eq. (2.306)
into Eq. (2.304) and regrouping yields for p D 0
Ł 1 Ł Eh2 Ł Eh3 2 N
Dw,xxxx C f,xx C w,xx D  1 cos i x 2.307a
R cR cR i
Eh Ł
fŁ,xxxx  w D0 2.307b
R ,xx
where i D i/L. If one neglects the effect of the boundary conditions on the
prebuckling solution (as is usually done for this type of analysis) then one must find
only a particular solution of Eq. (2.307). The use of the method of undetermined
coefficients yields easily
N
WŁ x D h 1 cos i x 2.308a
ci 
Eh3 1 N
fŁ x D 2
1 cos i x 2.308b
2c ˛i ci 
Mathematical Models for Perfect Structures 109

where  
1 2 1
ci D ˛i C 2 2.309
2 ˛i
is the classical axisymmetric buckling load and ˛2i D i2 Rh/2c/L2 .
Specializing the linearized stability Eqs. (2.306) to the axisymmetric imperfec-
tion Eq. (2.303) and the axisymmetric prebuckling state Eq. (2.308) yields
1O Eh2 Eh2 N1
Dr4 wO C f,xx C wO ,xx C cos i x wO ,yy
R cR R ci 
2c N1 ci O ,yy D 0
C ˛2i cos i x f (2.310a)
R ci 
N
O  Eh wO ,xx  2c Eh ˛2i i ci cos i x wO ,yy D 0
r4 f (2.310b)
R R ci 
a set of homogeneous, linear partial differential equations with variable coefficients.
If one assumes that the buckling mode is represented by
x y
wO D Cm sin m cos n 2.311
L R
which satisfies simply supported boundary conditions at the shell edges
wO D wO ,xx D 0 at x D 0, L, 2.312
then the linearized stability equations can be solved as follows. First an exact partic-
O p of the compatibility Eq. (2.310b) is solved for by the method of
ular integral f
undetermined coefficients. This guarantees that a kinematically admissible displace-
ment field is associated with an approximate solution of the other equation. An
approximate solution of the equilibrium Eq. (2.310a) by Galerkin’s method is then
equivalent to an approximate minimization of the second variation of the poten-
tial energy by the Rayleigh Ritz method. This guarantees that the eigenvalue so
obtained is an upper bound to the actual buckling load. Straightforward calculation
yields the following characteristic equation
 ci  2  c,mn   C  ci  C1 ci C C2 N1 υiD2m
C C3  C4 υim  2ci N12 D 0 2.313
where the axisymmetric buckling load ci is given by Eq. (2.309) and the asym-
metric buckling load c,mn by Eq. (2.105). The coefficients are
 
c ˛2i ˇn2 ˛2m ˛2i  m c ˇn2
C1 D C ; C2 D
2 ˛2m ˛2m C ˇn2 2 ˛2im C ˇn2 2 2 ˛2m
 
c2 ˛4i ˇn4 1 1 c2 ˛4i ˇn4 1
C3 D 2 2
C 2
; C4 D 2 2
2 ˛m ˛iCm C ˇn2 2 ˛im C ˇn2 2 2 ˛m ˛im C ˇn2 2
2.314
110 The Concepts of Elastic Stability

Figure 2.67 Effect of axisymmetric imperfection on the buckling load of axially compressed
cylinders

where
Rh   2 Rh   2
˛2iCm D i C m2 ; ˛2im D i  m2 ;
2c L 2c L
  
Rh 1 2
ˇn2 D n2 ; c D 31  2 
2c R
and
υiD2m D generalized Kronecker delta D 1 if i D 2m
D 0 otherwise.
The solution of this problem was first carried out by Koiter [2.89]. Using an imper-
fection in the form of the classic axisymmetric buckling mode
#
N x L 2c
wN D h1 cos icl  where icl D 2.315
L  Rh
he found that the minimum buckling load occurred when i D 2m for some value
of n. To obtain the results shown in Figure 2.67 one must find the smallest value
of which for a given axisymmetric imperfection N1 amplitude satisfies Eq. 2.313.
The reduced circumferential wave number ˇn is a free parameter in this equation,
with the restriction that the actual wave number n must be an integer.

2.2.3 Computational Tools for Bifurcation Problems

The majority of stability problems that arise at present in practical structural appli-
cations cannot be solved analytically. It might be possible, that after a number of
simplifying assumptions have been introduced one is able to obtain an approximate
solution via the Rayleigh Ritz or the Galerkin methods discussed earlier. However,
Mathematical Models for Perfect Structures 111

in nontrivial applications these methods may require considerable analytical and


computational effort before an approximate solution can be obtained.
Thus the point is soon reached where one looks towards the supposedly easier
approach offered by todays general purpose computer codes. A word of caution
is appropriate here. One should not expect that complicated structural stability
problems involving thin-walled plate and shell components, where nonlinear effects
play an important role, can be solved routinely without much effort and thought by
any of the many codes that are currently available. A thorough understanding of
the shell and stability theory involved supplemented by a good working knowledge
of the computational algorithms used are the prerequisites the analyst must possess
in order to be able to arrive at the appropriate solutions. Otherwise the chances are
high that incorrect or unreliable solutions will be obtained.
A comprehensive review of the currently available computer codes with buckling
analysis capabilities is beyond the scope of this book. Interested readers should
consult [2.90] and [2.91]. The state-of-the-art of buckling load calculations for
shells of revolution with very general wall construction loaded by a general axisym-
metric load system will be described in the following section using one of the more
popular finite difference codes available.

a. The BOSOR-4 Branched Complex Shell of Revolution Code

Although the BOSOR-4 program [2.92] represents the codification of three distinct
analyses, namely:
1. a linear stress analysis for axisymmetric and nonsymmetric behavior of axisym-
metric shell systems submitted to axisymmetric and nonsymmetric loads;
2. a nonlinear stress analysis for axisymmetric behavior of axisymmetric shell
systems;
3. an eigenvalue analysis in which the eigenvalues represent buckling loads or
vibration frequencies of axisymmetric shell systems submitted to axisymmetric
loads.
In the following only the nonsymmetric bifurcation problem from a nonlinear
axisymmetric prebuckling state will be discussed.
The independent variables of the analysis are the meridional arc length s,
measured along the shell reference surface and the circumferential coordinate .
For the cases considered it is possible to eliminate the circumferential coordinate
because:
1. in the nonlinear prebuckling analysis of axisymmetric behavior of axisym-
metric shell systems is not present;
2. in the bifurcation (eigenvalue) analysis the buckling modes (eigenvalues) vary
harmonically around the circumference.
The advantages of being able to eliminate one of the independent variables is very
significant. The number of calculations performed by the computer for a given mesh
112 The Concepts of Elastic Stability

point spacing along the arc-length s is greatly reduced, leading to great savings
in computer time. The disadvantage is, of course, the restriction to axisymmetric
structures, though in [2.92] and [2.93] methods are described by which BOSOR 4
can be used to analyze nonsymmetric structures of prismatic form.
The analysis is based on energy minimization with constraint conditions. The
total potential energy of the system involves:

1. strain energy of the shell segments Us ;


2. strain energy of the discrete rings Ur ;
3. potential energy of the applied line loads and pressures p ;
4. “energy of constraint” of the constraint conditions Uc .

The components of energy and the constraint conditions are initially in integro-
differential forms. They are then expressed in terms of the shell reference surface
displacement components ui , vi , wi at the finite difference mesh points and the
Lagrange multipliers i . The integration is performed numerically by the trape-
zoidal rule.
In the nonlinear prebuckling analysis the energy expression has terms linear,
quadratic, cubic and quartic in the dependent variables. The cubic and quartic terms
arise from the “rotation squared” terms which appear in the constraint conditions
and in the kinematic expressions for the reference surface strains ε1 , ε2 and ε12 . To
satisfy the equilibrium condition the energy, now in an algebraic form, is minimized
with respect to the discrete dependent variables. The resulting set of nonlinear
algebraic equations are solved for the displacement components at the mesh points
by the Newton Rhapson method. Stress and moment resultants are calculated in a
straightforward manner from the constitutive equations and the strain-displacement
relations.
The results from the nonlinear axisymmetric prebuckling analysis are then used
in the eigenvalue analysis for buckling. The prebuckling meridional and circum-
ferential stress resultants N10 and N20 and the meridional rotation 0 appear as
known variable coefficients in the second variation of the total potential energy
expression which governs buckling. This expression is a homogeneous quadratic
form. The values of the variable load, which render the quadratic form stationary
with respect to infinitesimal variations of the dependent variables, represent buck-
ling loads. These eigenvalues are calculated from a set of linear homogeneous
equations.

Shell strain energy Consider the typical shell segment shown in Figure 2.68.
The strain energy in the shell wall can be written in the form [2.94]

1
Us D 2 bεc[C]fεg C 2bNT cfεg r d ds 2.316

where
dCe D shell wall stiffness matrix 2.317a
bNT c D bNT1 , NT2 , 0, MT1 , MT2 , 0c thermal stress and moment resultants
Mathematical Models for Perfect Structures 113

Figure 2.68 Typical shell segment notation and sign convention (from [2.94])

and the strain-displacement relations are


   u0 C w/R1 C 21  2 C  2  
ε1
 ε   vP /r C ur 0 /r C w/R2 C 12  2 C  2  
 2   
 ε   uP /r C rv/r0 C  
f"g D  12  D   2.317b
 1   0 
   P 
2 /r C r 0 /r
212 2/r P C r 0 /r C v0 /R2 
 D w0  u/R1 ; P  v/R2 ;
D w/r  D 21 uP /r  v0  r 0 v/r 2.317c
Dots indicate differentiation with respect to , primes indicate differentiation with
respect to s. Positive values of u, v, w,  and  are shown in Figure 2.86. The quan-
tities R1 and R2 are the meridional and circumferential principal radii of curvature.
For a similar expression of the strain energy of a discrete ring see [2.94].

Potential energy of mechanical loads Two types of loads are permitted in the
analysis:

1. surface tractions p1 , p2 and p3 ;


2. line loads and moments V, S, H and M, which act at ring centroids and at
shell segment boundaries.

These loads are shown in Figures 2.68 and 2.69.


The potential energy associated with the surface tractions is for “live” loads [2.94]
   
1 1 1
p1 D  p 1 u C p 2 v C p3 w  p 3 C w2
2 R1 R2
  
1 u2 v2
C p3 C C p03 uw r d ds 2.318
2 R1 R2
whereas the potential energy associated with line loads at a given ring station can
be written as 
p2 D  Vuc C Svc C Hwc C Mrc d . 2.319
114 The Concepts of Elastic Stability

Figure 2.69 Discrete ring with centroidal displacement and forces (from [2.94])

Since all energy expressions must be expressed in terms of the same dependent
variables, therefore it is necessary to replace the ring displacements uc , vc , wc by
equivalent expressions in terms of the shell reference surface displacements u, v,
w. These variable transformations and expressions for “energy of constraint Vc ”
are given in [2.94].
The total energy of the system is obtained by summing over all shell segments,
discrete ring stiffeners and junctures.

Formulation of the stability problem When one attempts to solve the bifurca-
tion buckling problem of a complex, branched, ring-stiffened shell structure under
various systems of loads it is convenient to consider some of these to be known
and constant (or “fixed”) whereas the remaining ones are assumed to be unknown
eigenvalue parameters (or “variable”).
The notion of “fixed” and “variable” systems of loads helps in the formulation
of a sequence of simple classical eigenvalue problems for the solution of problems
governed by “nonclassical” eigenvalue problems. To illustrate the different types
of instability behavior consider the shallow spherical cap under external pressure
shown in Figure 2.70.
Deep spherical caps fail by bifurcation buckling where nonlinear prebuckling
effects are not important. On the other hand very shallow caps fail by nonlinear
axisymmetric collapse or snap-through buckling at pn , not by bifurcation buck-
ling at pb or pnb . Finally, there is an intermediate range of cap geometries that

Figure 2.70 Stability behavior of an externally pressurized shallow spherical cap


Mathematical Models for Perfect Structures 115

buckle by bifurcation buckling where the critical pressures are affected by nonlinear
prebuckling behavior. The analysis of this intermediate class of spherical caps is
simplified by the concept of “fixed” and “variable” pressure. Figure 2.70 shows
the load-deflection curve of a shallow spherical cap in this intermediate range. To
calculate the nonlinear bifurcation pressure pnb it is useful to consider it composed
of two parts
pnb D pf C pv 2.320

where
pf D a known or “fixed” quantity;
pv D an undetermined or “variable” quantity.

The fixed portion pf is an initial guess or the result of a previous iteration. The
variable portion pv is the remainder, which can be determined from a reasonably
simple eigenvalue problem. It is clear from Figure 2.70 that if pf is fairly close
to pnb , then the behavior in the range p D pf C pv is reasonably linear. Thus
the eigenvalue pnb can be calculated by means of a sequence of linear eigenvalue
problems. This procedure results in finding ever smaller pv values which are added
to the pf results from the previous iterations.
To illustrate the reduction of the bifurcation stability analysis to the solution
of a matrix eigenvalue problem consider the shell strain energy and the potential
energy of the surface tractions given by Eqs (2.316) and (2.318). This total potential
energy, denoted by , is quadratic in the shell reference surface displacement
components u, v, w and can be written in the form

 D 21 bεc[C]fεg C 2bNT cfεg C bdc[P]fdgr d ds 2.321

where

p/R1 0 p0
[P] D 0 p/R2 0 2.322a
p0 0 p1/R1 C 1/R2 
bdc D bu v wc 2.322b

and the other matrices have been defined earlier.


All expressions are referred to the undeformed surface of the shell. Next the
energy is expanded in a Taylor series about some equilibrium position by letting
u D u0f C u0v C uO
v D v0f C v0v C vO 2.323
w D w0f C w0v C wO

where uO , vO , wO are infinitesimal variations from the equilibrium state given by


u0f C u0v , vf0 C vv0 , w0f C w0v . Substituting Eq. (2.323) into the total potential energy
116 The Concepts of Elastic Stability

as given by Eq. (2.321) one obtains after some regrouping

 D   0 D υ C 12 υ2  C Ð Ð Ð

where υ contains all first order terms in the variations and υ2  contains all the
second order terms. Because the system is in equilibrium the first variation υ is
zero. The stability behavior is then governed by the second variation υ2 , which
after some regrouping can be put into the following form

υ2  D bε1 c[C]fε1 g C 2bε2 c[C]fε0 g C fNT g C bυc[P]fυgr ds d 2.324

where

bε0 c, bε1 c, bε2 c D zero, first and second order terms in the variations uO , vO , wO
bυc D bOu vO wc
O

and
   1  2 C  2  
u00 C w0 /R1 2 0 0
 vP 0 /r C u0 r 0 /r C w0 /R2   1 2 0 
   2  0 C 0  
 uP 0 /r C rv0 /r 0
  
f"0 g D f"01 g C f"02 g D   C  0 0 
 00    0 

 P 0 /r C 0 r 0 /r   
0
2P 0 /r C 0 r /r C v0 /R2 
0 0
0
2.325b
 
uO 0 C w/R
O 1 C 0 O C 0 O
 . 
 vO /r C uO r /r C w/R O 2 C 0 O C 0 O 
0
 uO /r C rOv/r C 0 O C 0 O
. 0 

f"1 g D   2.325c
O 0 
 
 O . /r C r O 0 /r 
2O /r C r /r C vO /R2 
. O 0 0

 1
O 2 C O 2  
2 
 1 O2 
C O 2  
 2
 

f"2 g D  O O 
. 2.325d
 0 
 
0
0
Next the prebuckling strain vectors f"01 g and f"02 g are divided into “fixed” and
“variable” parts
v
f"01 g D f"01
f
g C f"01 g; f"02 g D f"ff02 g C f"f02v g C f"02
vv
g 2.326
Mathematical Models for Perfect Structures 117

where
 1 f 2 f 2 
2 [0  C 0  ]
 1 f 2 f 2 
 2 [ 0  C 0  ] 
 f f 

f"ff02 g D  0 0  2.327a

 0 
 0 
0
   v 2 
0f 0v C 0f 0v 1 v 2
2 [0  C 0  ]
 f v f v  1 v 2 v 2 
 0 0 C 0 0   2 [ 0  C 0  ] 
   
 0f 0v C 0v 0f 
f"f02v g D  vv
f"02 
gD 0v 0v . 2.327b
 
 0   0 
 0   0 
0 0
The linear infinitesimal strain vector f"1 g can be divided into three components
f"1 g D f"11 g C f"1f g C f"1v g 2.328
where
 uO 0 C w/R
O 1 
 vO . /r C uO r 0 /r C w/R O 2 
 
 uO . /r C rOv/r0 
f"11 g D  0  2.329a
 O 
 O . /r C r O 0 /r 
2O . /r C O r 0 /r C vO 0 /R2 
   
0f O C 0f O 0v O C 0v O
 f O C  f O   vO v 
 0 0   0 C 0 O 
  f O C O f   v v

f"1 g D  0
f 0  0 O C O 0  .
f"1v g D  2.329b
 
 0   0 
 0   0 
0 0
Finally, the “pressure-rotation” matrix [P], Eq. (2.322a), and the “thermal load”
vector bNT c, Eq. (2.322b), can also be considered split into “fixed” and “variable”
parts
[P] D [Pf ] C [Pv ]; bNT c D bNTf c C bNTv c. 2.330
With these definitions one can rewrite part of the integrand of Eq. (2.324) as
follows
2b"2 c[C]f"0 g C fNT g C 2["2 ]fN0f g C fN0v g C fN0vv g 2.331
where
fN0f g D [C]f"f01 g C f"ff02 g C fNTf g 2.332a
118 The Concepts of Elastic Stability

fN0v g D [C]f"v01 g C f"f02v g C fNTv g 2.332b


fN0vv g D [C]f"vv
02 g 2.332c

Notice that the expressions given by Eq. (2.331) actually represent a quadratic
form. To indicate this explicitly the following change in notation is introduced.

2b"2 cfN0f g C fN0v g C fN0vv g D bc[N0f ] C [N0v ] C [N0vv ]fg 2.333

where
bc D bO O c
O 2.334a
   v 
Nf1 f
N12 0 N1 v
N12 0
[Nf0 ] D  N12
f
Nf2 0  [N0v ] D  Nv
12 N2v 0 
0 0 N1f C N2f  0 0 N1v C N2v 
2.334b
 
Nvv
1
vv
N12 0
[N0vv ] D  Nvv N2vv 0 .
12
vv vv
0 0 N1 C N2 
Assuming now that the “variable” parts are proportional to a scalar quantity ,
then upon substitution and regrouping the second variation υ2  from Eq. (2.324)
can be written 
υ2  D A1 C A2 C 2 A3 r ds d 2.335

where
A1 D bε11 C εf1 c[C]fε11 C εf1 g C bυc[Pf ]fυg C bc[N0f ]fg 2.336a
A2 D 2bε1v cCfε11 C ε1f g C bυc[Pv ]fυg C bc[N0v ]fg 2.336b
A3 D bε1v c[C]fε1v g C bc[Nvv
0 ]fg. 2.336c

In this expression the dependent variables uO , vO , wO are functions of the arc length s
and the circumferential coordinate . Additional details describing the contributions
of discrete rings and constraint conditions to υ2  are given in [2.62]. The -
dependence can be eliminated from the analysis by the following Fourier series
uO s,  D uO n s sin n
vO s,  D vO n s cos n 2.337
O  D wO n s sin n .
ws,

Upon substitution into Eq. (2.335) and carrying out the -integration will result in
an expression where the circumferential wave number n appears as a parameter and
where the corresponding expressions A1n , A2n , A3n are now functions of s only.
Mathematical Models for Perfect Structures 119

Figure 2.71 Finite-difference discretization: the “finite-difference element” (from [2.92])

To eliminate the s-dependence and to reduce the second variation of the total
potential energy υ2  to an algebraic form the finite-difference discretization shown
in Figure 2.71 is used. Notice that the “Ou” and “Ov” points are located halfway
between adjacent “w”O points. The “energy” contains up to first derivatives in uO and
vO and up to second derivatives in w.
O Hence, the “shell energy density” evaluated at
the center of the length  (the point labeled E) involves the following seven points
bqi c D bwO i1 , uO i , vO i , wO i , uO iC1 , vO iC1 , wO iC1 c. 2.338
The “energy” per unit circumferential length is simply the “energy” per unit area
multiplied by the length of the finite-difference element i , which is the arc length
of the reference surface between the adjacent uO or vO points. Thus
Ei D bqi c[B]T [C][B]fqi gi 2.339
where the matrices [B] and [C] represent the kinematic relation and the constitutive
law, respectively. In [2.95] it is shown that this formulation yields a (7 ð 7) stiffness
matrix corresponding to a constant-strain, constant-curvature-change finite element
that is incompatible in normal displacement and rotation at its boundaries but that
in general yields very rapidly converging results with increasing density of nodal
O
points. Notice that two of the w-points lie outside of the element.
Summing over all the finite-difference elements of length i the second variation
of the total potential energy υ2  can be written as
υ2  D bqc[K1 ] C [K2 ] C 2 [K3 ]fqg. 2.340
By Trefftz’s criterion instability occurs whenever υυ2  D 0. Minimization of
Eq. (2.340) with respect to the dependent variables uO i , vO i , wO i and the Lagrange
multipliers results in the following eigenvalue problem
[K1 ] C [K2 ] C 2 [K3 ]fqg D 0. 2.341
The eigenvalues of this “quadratic” eigenvalue problem are extracted by means of
the method of inverse power iterations with spectral shifts. For details the interested
reader should consult [2.62].
120 The Concepts of Elastic Stability

To demonstrate the capabilities of a modern shell of revolution code and to


illustrate the need for using rigorous nonlinear prebuckling analysis the stability
behavior of the very thin cylinder under axial compression from [2.92] is consid-
ered. Using the dimensions of this shell (radius R D 500 in., thickness h D 1 in.,
length L D 2000 in., Young’s modulus E D 107 psi and Poisson’s ratio v D 0.3)
one obtains from Eq. (2.111) its classical buckling load as

Nc D c h D 12104 lb/in.

In Figure 2.72 the discrete model of the same shell used for the BOSOR-4 runs is
shown. Notice that the cylinder is treated as being symmetric about the midlength,
and the 1000 in. half cylinder is divided into two segments: a 200 in.-long edge
zone segment with 83 mesh points, and an 800 in.-long interior segment with
99 mesh points. Simple support conditions are applied at the edge, and symmetry
conditions at the midlength. The sequence of wave number and load search, carried
out automatically by BOSOR-4 and described with some detail in [2.92], finally
yielded a critical buckling load of 10274 lb/in. and a buckling pattern consisting of
n D 18 full waves in the circumferential direction. The prebuckling displacement
at the predicted critical load and the axial dependence of the buckling mode are
shown in Figure 2.72.
Notice that the use of a rigorous prebuckling and buckling analysis resulted
in a 17.8 percent decrease of the predicted buckling load when compared with
the classical result of Eq. (2.111), which is based on a membrane prebuckling
analysis. It may be of interest to point out that the edge-buckling type behavior
here encountered might be missed if one does not use a fine enough mesh.

Figure 2.72 Buckling of an axially compressed cylinder (from [2.92])


Mathematical Models for Perfect Structures 121

b. Finite Element Formulation of Bifurcation Problems

A discussion of the computational tools available for bifurcation problems would


be incomplete without mentioning the very popular finite element method. In the
following the energy criteria of equilibrium and stability will be discussed in the
form proposed by Zienkiewicz [2.96].
If a conservative system is described by n generalized coordinates qi , i D
1, 2, . . . , n, then to total potential energy  of the system can be written
 D q1 , q2 , . . . , qn . 2.342
Equilibrium is satisfied if
υ D 0 2.343
which implies the following set of n nonlinear algebraic equations;
∂
D0 i D 1, 2, . . . , n. 2.344
∂qi
The equilibrium configuration is stable if the total potential energy is a relative
minimum, i.e.
∂2 
υ2  D ∂qi ∂qj > 0 i, j D 1, 2, . . . , n 2.345
∂qi ∂qj
where repeated indices indicate summation. Notice that in this case the associated
positive definite matrix Vij D υ2 /∂qi ∂qj has all positive eigenvalues r . If the
matrix Vij evaluated at an equilibrium point has any negative eigenvalues then
the total potential energy function  attains local maxima in the directions of the
corresponding eigenvectors and the system is in a state of unstable equilibrium.
The transition from stable to unstable equilibrium occurs when at least one
eigenvalue, say 1 , becomes zero. The matrix Vij is then singular and the corre-
sponding point on an equilibrium path is called a singular (or critical) point.
Singular points indicate either that there is a bifurcation of the equilibrium path
into other, stable or unstable branches or that a limit point has been reached. It
is therefore important to detect and calculate singular points in addition to stable
points on an equilibrium path.
Turning now to the finite element formulation, let the displacements at any point
within an elastic body be defined as a column vector fug, then
fug D [H]fqg 2.346
where the components of [H], the shape functions, are so chosen as to give the
appropriate nodal displacements when the coordinates of the corresponding nodes
are inserted and fqg contains all the nodal displacements. Notice that this and the
following expressions are to be interpreted as applying to the whole structure under
consideration.
With the displacements at all points within the body known one can proceed to
calculate the generalized strains (extensional strains and curvatures), which can be
122 The Concepts of Elastic Stability

written in matrix notation as

f"g D [B̄]fqg D [B0 ] C [BL ]fqg 2.347

where [B0 ] is the matrix obtained from the linear infinitesimal strain analysis and
[BL ] contains the contributions of the nonlinear strain components. Notice that [B0 ]
is independent of fqg whereas [BL ] is usually a linear function of fqg (see [2.96],
p. 414).
Next, assuming general linear elastic behavior, the relationship between stresses
and strains will be of the form

[] D [C]fεg 2.348

where [C] is the elasticity matrix containing the appropriate material properties.
Finally, following [2.96] the variational form of the overall equilibrium condition
can be written as 
υ D fυεgT fg dv  fυqgT fP̄ g D 0 2.349
v

where the column vector fP̄ g contains all the external nodal forces due to the
imposed loads and the integral is carried out over all the elements of the structure
under consideration. It is easily seen that the first term of this equation represents
the variation of the strain energy U of the structure while the second term is
the variation of the potential of the applied loads p . Using Eq. (2.347) one can
rewrite Eq. (2.349) as
 
T
υ D fυqg [B̄]fg dV  fP̄ g D 0 2.350
v

The stability criterion involves the second variation of the total potential energy.
Computing it one gets
  
2 T T T
υ  D υυ D fυqg υ[B̄] fg dV C [B̄] υfg dV 2.351
v v

N D 0.
Notice since fP̄g is independent of fqg, therefore υfPg
With the help of Eqs. (2.347) and (2.348) it is straight forward to rewrite
Eq. (2.351) as
  
υ2  D fυqgT υ[B̄]T fg dV C [B̄]T [C][B̄]fυqg dV . 2.352
v v

The first term of this equation can generally be written as (see [2.96] and [2.97]
for details) 
υ[B̄]T fg dV D [K ]fυqg 2.353
v

where [K ] is a symmetric matrix which depends on the stress level and is called
the initial stress or geometric matrix. Finally, substituting for [B̄] from Eq. (2.347)
Mathematical Models for Perfect Structures 123

and regrouping, the second variation of the total potential energy can be written in
the following quadratic form

υ2  D fυqgT [KT ]fυqg > 0 2.354

where [KT ] D [K ] C [K0 ] C [KL ] is the tangent stiffness matrix and

[K0 ] D [B0 ]T [C][B0 ] dV 2.355
v

[KL ] D [B0 ]T [C][BL ] C [BL ]T [C][B0 ] C [BL ]T [C][BL ] dV 2.356
v

Notice that [K0 ] represents the usual small displacements stiffness matrix, whereas
the matrix [KL ] is due to the large displacements and is variously known as the
initial displacement or large displacement matrix.
Thus, when the finite element discretization is employed, Eq. (2.354) represents
the stability criterion of an equilibrium configuration. From the theory of quadratic
forms one knows that a stable equilibrium configuration is ensured if the tangent
stiffness matrix
[KT ] D [K ] C [K0 ] C [KL ] 2.357

has no negative eigenvalues. A critical point is reached when [KT ] has at least
one zero eigenvalue. Thus the stability of an equilibrium configuration can be
determined by solving the eigenvalue problem

[KT ]fXr g D r fXr g 2.358

at the current equilibrium state, where r is the rth eigenvalue and fXr g is the
corresponding eigenvector.
Notice that the computation of the critical point must be done in two steps. First,
the equilibrium configuration associated with a given load level P is computed.
Next, the stability of this configuration is examined by calculating the eigenvalues
of [KT P], the tangent stiffness matrix evaluated at the load P.
This method of determining the stability of a conservative system is very accu-
rate, however it can be computationally expensive because it involves the solution
of a quadratic eigenvalue problem for the critical load (see also Eq. (2.341)).
Cheaper methods of estimating the critical load are available. These methods are
usually referred to as linearized buckling analyses, where the critical load is calcu-
lated based on a linear extrapolation of the behavior of the structure at a small
load level.
Considering Eqs. (2.355), (2.356) and (2.357) one observes that only the matrices
[K ] and [KL ] depend on the load level P. As a first approximation these matrices
can be assumed to be only linearly proportional to the applied load. Then the
tangent stiffness matrix at some level P can be approximated as
P
[KT P] D [K0 ] C [K P] C [KL P] 2.359
P
124 The Concepts of Elastic Stability

where the initial stress matrix [K ] and the large displacement matrix [KL ] are
both evaluated at a small load level P. If one assumes further that the critical
load can be approximated by  Ð P, then the condition for a singular point (i.e.,
a singular tangent stiffness matrix) becomes a standard matrix eigenvalue problem.
Once the lowest eigenvalue 1 is found, the critical buckling load is equal to
1 Ð P and the buckling mode is given by the eigenvector fX1 g.
An additional simplification is frequently used. It is based on the argument that
at the low load level P the displacements fug are so small that one can neglect
the contribution of the large displacement matrix [KL ]. This leads to the classical
“initial” stability problem
[K0 ] C r [K P]fXr g D 0 2.360
frequently used for investigating the stability of structures consisting of struts,
plates and shells.
One must realize here that strictly speaking this approach can only give phys-
ically significant answers if the elastic solution based on the small displacement
stiffness matrix [K0 ] yields such deformations that the large displacement matrix
[KL ] is identically zero. Zienkiewicz warns explicitly in [2.96] and [2.97] that this
only happens in a very limited number of practical situations (such as a perfectly
straight column under axial load). Thus in real engineering applications the stability
problem should ultimately always be investigated by using the full tangent stiff-
ness matrix. That is, the step-by-step formulation given by Eq. (2.358) should be
employed.
There are many commercially available finite element codes with buckling
analysis capabilities such as NASTRAN [2.98], ADINA [2.99], MARC [2.100],
ANSYS [2.101], and ABAQUS [2.102], just to name a few. A comprehensive
review of these and other codes is obviously beyond the scope of this book. Inter-
ested readers should consult [2.90] and [2.91] for further information.

References

2.1 Timoshenko, S.P. and Gere, J.M., Theory of Elastic Stability, McGraw-Hill, New
York, 1961.
2.2 Brush, D. O. and Almroth, B. O., Buckling of Bars, Plates and Shells, McGraw-Hill
Book Company, New York, 1975.
2.3 Allen, H.G. and Bulson, P.S., Background to Buckling, McGraw-Hill, New York,
1980.
2.4 Simitses, G.J., An Introduction to the Elastic Stability of Structures, Prentice-Hall,
Englewood Cliffs, New Jersey, 1976.
2.5 Dym, C.L., Stability Theory and Its Applications to Structural Mechanics, Noord-
hoff Int. Publishing, Leyden, 1974.
2.6 Thompson, J.M.T. and Hunt, G.W., A General Theory of Elastic Stability, John
Wiley & Sons, London, 1973.
2.7 Gerard, G., Introduction to Structural Stability Theory, McGraw-Hill, New York,
1962.
References 125

2.8 Pflüger, A., Stabilitätsprobleme der Elastostatik, Springer Verlag, Berlin, 1975.
2.9 Euler, L., De curvis elasticis, Leonhard Euler’s Elastic Curves, translated and anno-
tated by W.A. Oldfather, C.A. Ellis, and D.M. Brown, reprinted from Isis, 20, (58),
1933, The St. Catherine Press, Bruges, Belgium.
2.10 Rivello, R.M., Theory and Analysis of Flight Structures, McGraw-Hill, New York,
1969.
2.11 Shanley, F.R., Inelastic Column Theory, Journal of the Aeronautical Sciences, 14,
(5), May 1947, 261 268.
2.12 Bryan, G.H., On the Stability of a Plane Plate under Thrusts in its Own Plane with
Applications to the Buckling of the Sides of a Ship, Proc. London Math. Soc., 22,
1891, 54 67.
2.13 Gerard, G. and Becker, H., Handbook of Structural Stability, Part 1: Buckling of
Flat Plates, NACA TN 3781, 1957.
2.14 Coan, J.M., Large-Deflection Theory for Plates with Small Initial Curvatures Loaded
in Edge Compression, ASME Journal of Applied Mechanics, 18, (2), June 1951,
143 151.
2.15 Hu, P.C., Lundquist, E.E. and Batdorf, S.B., Effect of Small Deviations from Flat-
ness on Effective Width and Buckling of Plates in Compression, NACA TN 1124,
1946.
2.16 Notenboom, R.P., Finite Strip Elements in Thin Plate Buckling Analysis, Report
LR-642, Delft University of Technology, Faculty of Aerospace Engineering, Delft,
September 1990.
2.17 van der Neut, A., The Interaction of Local Buckling and Column Failure of Thin-
Walled Compression Members, in: Proceedings 12th IUTAM Congress in Stanford,
California, 1968, M. Hetényi and W.G. Vincenti, eds., Springer Verlag, Berlin-
Heidelberg-New York, 1969, 389 399.
2.18 Koiter, W.T. and Kuiken, G.D.C., The Interaction between Local Buckling and
Overall Buckling on the Behaviour of Built-Up Columns, Report WTHD-23, Delft
University of Technology, Delft, 1971.
2.19 Nethercot, D.A., Elastic Lateral Buckling of Beams, in Beams and Beam Columns,
R. Narayanan, ed., Applied Science Publishers, London and New York, 1983, 1 34.
2.20 Rockey, K.C., El-Gaaly, M.A. and Bagchi, D.K., Failure of Thin-Walled Members
under Patch Loading, Journal of the Structural Division, ASCE Proceedings, 98,
(ST12) Dec. 1972.
2.21 Rockey, K.C. and Bagchi, D.K., Buckling of Plate Girder Webs under Partial Edge
Loadings, International Journal of Mechanical Sciences, 12, (1), 1970, 61 76.
2.22 Chen, W.F. and Atsuto, T., Theory of Beam Columns, McGraw-Hill, N.Y., 1977.
2.23 Bodner, S.R., On the Conservativeness of Various Distributed Force Systems,
Journal of the Aeronautical Sciences., 25, (2), February 1958, 132 133.
2.24 Langhaar, H.L., Energy Methods in Applied Mechanics, John Wiley & Sons, New
York, 1962.
2.25 Washizu, K., Variational Methods in Elasticity and Plasticity, 2nd ed., Pergamon
Press, Oxford, 1975.
2.26 Bresse, M., Cours de Méchanique Appliquée, 2nd ed., 1866, 333 334.
2.27 Lévy, M., Memoir Sur un Nouveau Cas Intégrable du Probléme de L’Élastique et
l’Une de ses Applications, Journal des Mathématiques Pure et Appliqueés, Ser. 3,
10, 1884, 5 42.
2.28 Boresi, A.P., A Refinement of the Theory of Buckling of Rings under Uniform
Pressure, ASME Journal of Applied Mechanics, 22, (1), March 1955, 95 103.
126 The Concepts of Elastic Stability

2.29 Singer, J. and Babcock, C.D. Jr., On the Buckling of Rings Under Constant Direc-
tional and Centrally Directed Pressure, ASME Journal of Applied Mechanics, 37,
(1), March 1970, 215 218.
2.30 Hurlbrink, E., Festigkeits-Berechnung von röhrenartigen Körpern, die unter
äusserem Druck stehen, Schiffbau, 9, (14), 1908, 517.
2.31 Timoshenko, S.P., Buckling of a Uniformly Compressed Circular Arch, Bull. Poly-
tech. Inst., Kiev, 1910.
2.32 Schreyer, H.L. and Masur, E.F., Buckling of Shallow Arches, Journal of the
Engineering Mechanics Division, ASCE Proceedings, 92, (EM4), August 1966,
1 19.
2.33 Kerr, A.D. and Soiter, M.T., The Linearization of the Prebuckling State and its
Effect on the Determined Instability Loads, ASME Journal of Applied Mechanics,
36, December 1969, 775 783.
2.34 Koiter, W.T., General Equations of Elastic Stability for Thin Shells, Proceedings
of the Symposium on the Theory of Shells to Honor Hamilton Donnell, University
of Houston, Houston, Texas, 1967, 187 228.
2.35 Donnell, L.H., Stability of Thin-Walled Tubes under Torsion, NACA Report
No. 479, 1933.
2.36 Koiter, W.T., On the Stability of Elastic Equilibrium, Ph.D. Thesis (in Dutch), TH-
Delft, H.T. Paris, Amsterdam, 1945. English translation issued as NASA TT F-10,
1967.
2.37 Lorenz, R., Achsensymmetrische Verzerrungen in dünnwandigen Hohlzylinder,
Zeitschrift des Vereines Deutscher Ingenieure, 52, 1908, 1706 1713.
2.38 Thielemann, W., Schnell, W. and Fischer, G., Beul und Nachbeulverhalten
Orthotroper Kreiszylinderschalen unter Axial und Innendruck, Zeitschrift für
Flugwissenschaften, 8, (10/11), 1960, 284 293.
2.39 Becker, H. and Gerard, G., Elastic Stability of Orthotropic Shells, Journal of the
Aerospace Sciences, 29, (5), 1962, 505 512.
2.40 Hedgepeth, J.M. and Hall, D.B., Stability of Stiffened Cylinders, AIAA Journal, 3,
(12), December 1965, 2275 2286.
2.41 Singer, J., Baruch, M. and Harari, O., On the Stability of Eccentrically Stiffened
Cylindrical Shells under Axial Compression, International Journal of Solids and
Structures, 3, 1967, 445 470.
2.42 Tasi, J., Effect of Heterogeneity on the Stability of Composite Cylindrical Shells
under Axial Compression, AIAA Journal, 4, (6), June 1966, 1058 1062.
2.43 Meck, H.R., A Survey of Methods of Stability Analysis of Ring Stiffened Cylinders
Under Hydrostatic Pressure, Transactions of ASME, Journal of Engineering for
Industry, 87B (3), 1965, 385 390.
2.44 Cheng, S. and Ho, B.P.C., Stability of Heterogeneous Aelotropic Cylindrical Shells
under Combined Loading, AIAA Journal, 1, (4), April 1963, 892 898.
2.45 Southwell, R.V. and Skan, S.W., On the Stability under Shearing Forces of a Flat
Elastic Strip, Proc. Roy. Soc. London, Ser. A, 105, 1924, 582 607.
2.46 Schwerin, E., Die Torsions-Stabilität des dünnwandigen Rohres”, Z. Angew. Math.
Mech. (ZAMM), 5, 1925, 235 243.
2.47 Yamaki, N., Elastic Stability of Circular Cylindrical Shells, North-Holland Series
in Applied Mathematics and Mechanics, Amsterdam, 1985.
2.48 Flügge, W., Die Stabilität der Kreiszylinderschale, Ingenieur Archiv, 3, 1932,
463 506.
References 127

2.49 Seide, P. and Weingarten, V.I., On the Buckling of Circular Cylindrical Shells
under Pure Bending, ASME Journal of Applied Mechanics, 28, (1), March 1961,
112 116.
2.50 Potters, M.L., A Matrix Method for the Solution of a Linear Second Order Differ-
ence Equation in Two Variables, Report M.R. 19, Mathematisch Zentrum, Amster-
dam, 1955.
2.51 Budiansky, B. and Radkowski, P.P., Numerical Analysis of Unsymmetrical Bending
of Shells of Revolution, AIAA Journal, 1, (8), August 1963, 1833 1842.
2.52 Stephens, W.B., Starnes, J.H. and Almroth, B.O., Collapse of Long Cylindrical
Shells under Combined Bending and Pressure Loads, AIAA Journal, 13, (1), January
1975, 20 25.
2.53 Almroth, B.O., Brogan, F.A., Miller, E., Zele, F. and Peterson, H.T., Collapse
Analysis for Shells of General Shape: User’s Manual for the STAGS-A Computer
Code, Report AFFDL-TR-71-8, Air Force Flight Dynamics Lab, Wright-Patterson
AFB, 1973.
2.54 Hutchinson, J.W., Imperfection Sensitivity of Externally Pressurized Spherical
Shells, ASME Journal of Applied Mechanics, 34, (1), March 1967, 49 55.
2.55 Flügge, W., Stresses in Shells, Springer Verlag, Berlin, 1962.
2.56 Stein, M. and McElman, J.A., Buckling of Segments of Toroidal Shells, AIAA
Journal, 3, (9), September 1965, 1704 1709.
2.57 Hutchinson, J.W., Initial Post-Buckling Behavior of Toroidal Shell Segments, Inter-
national Journal of Solids and Structures, 3, 1967, 97 115.
2.58 Galletly, G.D., Torispherical shells A Caution to Designers, ASME Journal of
Engineering Industry, 81, (1), February 1959, 51 66.
2.59 Galletly, G.D. and Radhamohan, S.K., Elastic-Plastic Buckling of Internally Pres-
surized Thin Torispherical Shells, ASME Journal of Pressure Vessel Technology,
101, 1979, 216 225.
2.60 Galletly, G.D. and Blachut, J., Torispherical Shells under Internal Pressure Fai-
lure due to Asymmetric Plastic Buckling or Axisymmetric Yielding, Proceedings
Institution of Mechanical Engineers, 199 (C3), 1985, 225 238.
2.61 Galletly, G.D., Design Equations for Preventing Buckling in Fabricated Torispher-
ical Shells Subjected to Internal Pressure, Proceedings Institution of Mechanical
Engineers, 200, (A2), 1986, 127 139.
2.62 Bushnell, D., Stress, Stability and Vibration of Complex Branched Shells of
Revolution Analysis and User’s Manual for BOSOR 4, NASA CR-2116, 1972.
2.63 Bushnell, D., BOSOR5 Program for Buckling of Elastic-Plastic Complex Shells
of Revolution Including Large Deflections and Creep, Computers and Structures,
6, 1976, 221 239.
2.64 Arbocz, J. and Hol, J.M.A.M., Koiter’s Stability Theory in a Computer Aided Engi-
neering (CAE) Environment, International Journal of Solids and Structures, 26,
(9/10), 1990, 945 973.
2.65 Föppl, L., Achsensymmetrisches Ausknicken zylindrischer Schalen, S.-B. Bayr.
Akad. Wiss., 1926, 27 40.
2.66 Stein, M., The Influence of Prebuckling Deformations and Stresses on the Buckling
of Perfect Cylinders, NASA TR R-190, 1964.
2.67 Fischer, G., Über den Einfluss der gelenkigen Lagerung auf die Stabilität
dünnwandiger Kreiszylinderschalen unter Axiallast und Innendruck, Zeitschrift für
Flugwissenschaften, 11, (3), 1963, 111 119.
128 The Concepts of Elastic Stability

2.68 Almroth, B.O., Influence of Edge Conditions on the Stability of Axially


Compressed Cylindrical Shells, NASA CR-161, February 1965.
2.69 Hoff, N.J. and Soong, T.C., Buckling of Circular Cylindrical Shells in Axial Com-
pression, International Journal of Mechanical Sciences, 7, 1965, 489 520.
2.70 Brazier, L.G., On the Flexure of Thin Cylindrical Shells and Other “Thin” Sections,
Proc. Royal Society, Series A, 116, 1926, 104 114.
2.71 Kyriakides, S. and Ju, G.T., Bifurcation and Localization Instabilities in Cylindrical
Shells under Bending I. Experiments, International Journal of Solids and
Structures, 29, (9), 1992, 1117 1142.
2.72 Ju, G.T. and Kyriakides, S., Bifurcation and Localization Instabilities in Cylindrical
Shells under Bending II. Predictions, International Journal of Solids and
Structures, 29, (9), 1992, 1143 1171.
2.73 Ostgood, W.R., The Double-Modulus Theory of Column Action, Civil Engineering,
5, (3), March 1935, 172 175.
2.74 Considère, A., Resistance des pieces comprimées, Congrès Int. des Procédés de
Construction, 3, Annexe, Librairie Polytechnique, Paris, 1891, 371 378.
2.75 Engesser, Fr., Über Knickfragen, Schweizerische Bauzeitung, 26, (4), July 1895,
24 30.
2.76 von Kármán, Th., Die Knickfestigkeit gerader Stäbe, Physikalische Zeitschrift, 9
(4) 1908, 136 140..
2.77 Ramberg, W. and Osgood, W.R., Description of Stress-Strain Curves by Three
Parameters, NACA TN 902, July, 1943.
2.78 Bruhn, E.F., Analysis and Design of Flight Vehicle Structures, Tri-State Offset Co.,
Cincinnati, Ohio, 1965.
2.79 Hill, R., A General Theory of Uniqueness and Stability in ElasticnPlastic Solids,
J. Mech. Phys. Solids, 6, 1958, 236 249.
2.80 Hill, R., Bifurcation and Uniqueness in Nonlinear Mechanics of Continua,
(Muskhelishvili Volume) Soc. Indust. Appl. Math., Philadelphia, 1961, 153 164.
2.81 Sewell, M.J., A Survey of Plastic Buckling, in Stability, H. Leipholz, ed.,
Chapter 5, 1972, 85 197.
2.82 Trefftz, E., Zur Theorie der Stabilität des elastischen Gleichgewichts, ZAMM, 13,
1933, 160 165.
2.83 Beck, M., Die Knicklast des einseitig eingespannten, tangential gedrückten Stabes,
ZAMP, 3, 1952, 225 228.
2.84 Ziegler, H., Principles of Structural Stability, Blaisdell Publishing Co., Waltham,
Massachusetts, 1968.
2.85 Bolotin, V.V., Nonconservative Problems of the Theory of Elastic Stability, A Perg-
amon Press Book, The MacMillan Co., New York, 1963.
2.86 Meirovitch, L., Analytical Methods in Vibrations, The MacMillan Co., London,
1967.
2.87 Van der Neut, A., Buckling Caused by Thermal Stresses, in High Temperature
Effects in Aircraft Structures, N.J. Hoff, ed., Pergamon Press, New York, 1958,
215 246.
2.88 Galerkin, B.G., Beams and Plates, (in Russian), Vestnik Inzhenerov, 1, (19), 1915,
897 908.
2.89 Koiter, W.T., The Effect of Axisymmetric Imperfections on the Buckling of Cylin-
drical Shells under Axial Compression, Koninkl. Ned. Akad. Wetenschap. Proc. B66,
1963, 265 279.
References 129

2.90 Pilkey, W., Saczalski, K. and Schaeffer, H. eds., Structural Mechanics Computer
Programs, Surveys, Assessments and Availability, University of Virginia Press,
Charlottesville, VA., 1974.
2.91 Noor, A.K., Belytschko, T. and Simo, J.C. eds., Analytical and Computational
Models of Shells, CED ASME, 3, 1989.
2.92 Bushnell, D., Stress, Stability and Vibration of Complex Branched Shells of Revo-
lution, Computers and Structures, 4, Pergamon Press, 1974, 399 435.
2.93 Bushnell, D., Stress, Buckling and Vibration of Prismatic Shells, AIAA Journal, 9,
(10), October 1971, 2004 2013.
2.94 Bushnell, D., Analysis of Ring-Stiffened Shells of Revolution under Combined
Thermal and Mechanical Loading, AIAA Journal, 9, (3), March 1971, 401 410.
2.95 Bushnell, D., Finite-Difference Energy Models versus Finite-Element Models: Two
Variational Approaches in One Computer Program, Proceedings ONR International
Symposium for Numerical and Computer Methods in Structural Mechanics, Urbana,
Illinois, September 1971.
2.96 Zienkiewicz, O.C., The Finite Element Method in Engineering Science, 2nd Edition,
McGraw-Hill, London, 1971.
2.97 Zienkiewicz, O.C., The Finite Element Method, 3rd Edition, McGraw-Hill Book
Co. (UK), London, 1977.
2.98 NASTRAN, The MacNeal-Schwendler Corporation, 815 Colorado Blvd, Los
Angeles, California, 90041, U.S.A.
2.99 ADINA, ADINA R&D, Inc., 71 Elton Ave., Watertown, Massachusetts, 02172,
U.S.A.
2.100 MARC, MARC Analysis Research Corporation, 260 Sheridan Ave., Palo Alto,
California, 94306, U.S.A.
2.101 ANSYS, Swanson Analysis Systems, Inc., P.O. Box 65, Johnson Road, Houston,
Pennsylvania, 15342, U.S.A.
2.102 ABAQUS, Hibbitt, Karlsson & Sorensen, Inc., 100 Medway Str., Providence,
Rhodes Island, 02906, U.S.A.
3
Postbuckling Behavior of
Structures

3.1 Introduction
The chances for a successful correlation between the test data and the appli-
cable theoretical results will be greatly enhanced if one tries to take into account
already at the planning stage all those factors that may affect the outcome of the
experiments.
In the last 50 years or so extensive experimental and theoretical research
programs have been carried out in the aerospace, (sub) marine, pressure vessel
and off-shore industries trying to establish a reliable design basis for buckling
sensitive applications. It has been found that, in these cases, great care must be
taken in defining the boundary conditions adequately, one has to check whether
inelastic effects will occur, and one has to investigate whether the buckling load is
sensitive to the unavoidable initial imperfections always present in real structures.
Depending on the application, initial imperfections could have different mean-
ings. Unwanted load eccentricities by columns, slight deviations from flatness by
plate assemblies or minute waviness along the generator of a cylindrical shell are all
examples of initial (geometric) imperfections. Theoretical and experimental inves-
tigations have shown that the degree to which the presence of initial imperfections
will affect the occurrence of the bifurcation buckling load depends on the partic-
ular 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.
As a general result one can state that in order to characterize the buckling
behavior of a slender, thin-walled structure one must investigate both its (bifurca-
tion) buckling and its postbuckling behavior under the specified external loading.
In the following, typical characteristic postbuckling behaviors will be illustrated
using different structural elements.
The exact solution of the postbuckling behavior of a perfect column is known. As
pointed out in [2.2] and as can be seen in Figure 2.2 the postbuckling curve of an

Buckling Experiments: Experimental Methods in Buckling of Thin-Walled Structures: Basic Concepts, Columns, Beams and
Plates – Volume 1. J. Singer, J. Arbocz and T. Weller Copyright © 1998 John Wiley & Sons, Inc.
132 Postbuckling Behavior of Structures

axially compressed perfectly straight column is tangent to the horizontal line at the
bifurcation point P/Pc D 1.0, where the lateral deflection is zero. Notice that the
solution curves for columns with small initial imperfections closely approximate
the perfect curve. Thus one can expect a good agreement between the theoretical
predictions and the experimental results. However, it must be remembered that the
postbuckling curves shown in Figure 2.2 are only valid up to the proportional limit
of the material.
From the equilibrium paths for initially perfectly flat and for slightly imperfect
plates shown in Figure 2.7 it is evident, that plates subjected to in-plane compres-
sion will carry additional load after buckling if the unloaded edges are supported.
Further it can be seen, that the buckling of imperfect plates appears to be so
gradual that it becomes difficult to decide at precisely what load the buckling may
be said to occur. Notice that also in this case the solution curves for small initial
imperfections follow closely the theoretical curve for the perfect plate both in the
prebuckling and in the postbuckling regions. Thus once again one can expect a
close agreement between test results and theoretical predictions.
Quite another type of behavior appears to predominate when one considers the
correlation between theoretical and experimental results for axially compressed
cylindrical shells as attempted by Flügge [2.48], Lundquist ([9.53], Volume 2),
Donnell ([9.54], Volume 2) and others. As can be seen in Figure 3.1 (from [9.103],
Volume 2) the tests reveal a wide scatter in the experimental results, with exper-
imental buckling loads for very thin shells (R/h > 1000) as low as 20 percent
of the theoretical values. The reason for this behavior becomes evident if one
considers the postbuckling equilibrium paths for axially compressed cylinders
calculated by von Kármán and Tsien [3.1] using a nonlinear theory. As can be
seen in Figure 3.2, their results show that the postbuckling equilibrium path drops
sharply downward from the bifurcation point. Although von Kármán and Tsien
did not analyze initially imperfect shells, their results suggest that equilibrium

Figure 3.1 Test data for isotropic cylinder under axial compression (from [9.103], Volume 2)
Introduction 133

Figure 3.2 Theoretical postbuckling curves for axially compressed cylindrical shells

paths for shells with initial imperfections might have the form as indicated by
the dashed curve in Figure 3.2. This conjecture was confirmed by the well known
analysis of initially imperfect cylindrical shells presented by Donnell and Wan in
1950 [3.2].
Rigorous confirmation of the influence of initial imperfections was given by
Koiter. Thanks to his pioneering work [2.36], first published in 1945, and the
efforts of many investigators since then, the theory of imperfection sensitivity
of elastic and inelastic structures is well developed, and today one has a thorough
understanding of the principal factors that must be considered for a reliable predic-
tion of the buckling load. What often is missing, however, are the experimental
data (read information about the initial imperfections present in the structure and
precise definition of the boundary conditions) needed for a successful prediction.
See Chapters 10 and 11, Volume 2 for further details.
In Koiter’s theory the initial postbuckling behavior plays a central role. When
the initial portion of the secondary path emanating from the bifurcation point
has a positive slope, considerable postbuckling strength can be developed by the
structure, and loss of stability on the primary path does not result in structural
collapse. On the other hand, when the initial portion of the secondary (postbuckling)
path has a negative slope, the buckling is sudden, explosion-like and the magnitude
of the critical collapse load is subject to the influence of initial imperfections.
Koiter’s theory is exact in the asymptotic sense, that is, it is exact at the bifurcation
point itself and a close approximation for postbuckling configurations near the
bifurcation point.
Summarizing, in order to obtain an estimate of the critical load levels of imper-
fect structures one can rely either on the predictions of an asymptotic analysis
or one can choose to consider the results of a general nonlinear analysis. In the
following both approaches will be described in more detail.
134 Postbuckling Behavior of Structures

3.2 Asymptotic Imperfection Sensitivity Analysis


The Koiter type asymptotic analysis consists basically of a perturbation expansion
about the lowest (critical) eigenvalue of the structure. That is, one is interested
in the variation of  with  in the vicinity of the bifurcation point  D c of the
perfect structure, where  is the loading parameter and  is the suitably normalized
amplitude of the buckling mode. If the structure possesses a unique buckling mode
associated with the lowest buckling load, then its buckling and initial postbuckling
behavior can be represented by

D 1 C a C b 2 C Ð Ð Ð 3.1
c
where a and b are the first and the second postbuckling coefficients, respectively.
Figure 3.3 illustrates the case when a < 0 and b > 0, whereas Figure 3.4a and 3.4b
show the cases when a D 0, and b > 0 or b < 0, respectively. Notice that in all
the figures initially, along the prebuckling branch the buckling displacement  of
the perfect structure is identically zero for increasing load  until the bifurcation
load c is reached. Also the name bifurcation buckling gives a fitting description
of the transition of the state of the structure from the fundamental equilibrium path
to the buckled path (in either direction) at  D c .
To answer the question, what shall be the behavior of the structure when it
is subjected to a load that is increased slowly from zero, one has to introduce

Figure 3.3 Asymmetric equilibrium paths for a < 0 and b > 0

Figure 3.4 Symmetric equilibrium paths for a D 0 (case a: b > 0; case b: b < 0)
Asymptotic Imperfection Sensitivity Analysis 135

a suitably chosen initial imperfection into the mathematical model. Thus, if one
assumes a small, stress free, initial imperfection of amplitude N then one can
describe the variation of  with  in the vicinity of  D c by the following
expression [3.3].
  c  D ac  2 C bc  3 C Ð Ð Ð  ˛c N  ˇ  c N C Ð Ð Ð 3.2
where ˛ and ˇ are the so-called first and second imperfection form factors. Notice
that, as can be seen from Figure 3.4, this expression is chosen so as to have the
correct limiting behavior, namely
 
lim lim  D c and lim  D 0 if N 6D 0. 3.3
!0 N
!0 !0

If the initial imperfection is assumed to have the shape of the critical buckling
mode and one uses a membrane prebuckling analysis then ˛ D ˇ D 1 and Eq. (3.2)
reduces to  
 
1  C a 2 C b 3 C Ð Ð Ð D N 3.4
c c
the form originally proposed by Koiter [2.36].
As can be seen from Figures 3.3 and 3.4 the shape of the secondary equilibrium
path plays a central role in determining the influence of the initial imperfections.
When the initial portion of the secondary path slopes upward then the structure
can develop considerable postbuckling strength, and the loss of stability of the
primary (fundamental) path does not result in structural collapse. However, when
the initial portion of the secondary path slopes downward, then in most cases
buckling will occur violently and the magnitude of the critical load c of the
real (imperfect) structure is lower than the bifurcation buckling load c of the
corresponding idealized (perfect) structure.
Notice also that in the case of asymmetric equilibrium paths the sign of the
initial imperfection plays an important role. In Figure 3.3 a positive N produces
an imperfect sensitive configuration, with the buckling load of the real (imperfect)
structure s less than c , the bifurcation buckling load of the perfect structure. On
the other hand, a negative N has no degrading effects so far as elastic buckling is
concerned.
For the cases with symmetric equilibrium paths, as can be seen from Figure 3.4
the sign of the initial imperfection is immaterial. Whether the buckling load of the
structure is imperfection sensitive or not is governed by the sign of the second post-
buckling coefficient b. Notice that in these cases the first post-buckling coefficient
a is identically equal to zero.
What makes the use of asymptotic methods so attractive is that the postbuckling
coefficients a and b are properties of the perfect structure. Hence their computation
does not involve the shape and the size of the expected initial imperfections. With
the knowledge of the postbuckling coefficients one can make qualitative predictions
about the nature of the experimental results.
Thus, if the postbuckling path of the loaded structure has a limit point, then the
buckling load s is sensitive to initial imperfections. In this case the experimentalist
136 Postbuckling Behavior of Structures

must expect that the test results will be in general lower than the predictions based
on the stability analysis of the perfect structure. Furthermore, the results of repeated
buckling tests are probably going to exhibit noticeable scatter.
On the contrary, if the postbuckling path of the loaded structure is monotonically
increasing, then initial elastic buckling will not result in a collapse of the structure.
It can be loaded further and one says that the structure has additional postbuckling
strength. In this case the test results will, in general, agree quite well with the
theoretical predictions of the stability analysis of the perfect structure. Also the
scatter of the results of carefully executed repeated buckling tests should be slight.

3.2.1 Initial Postbuckling Behavior of Columns

To investigate the initial postbuckling behavior of an axially compressed, simply


supported, slender column one must first develop an asymptotic expression for the
total potential energy  valid in the neighborhood of the critical load. It has been
shown that the small parameter  involved can be taken as the amplitude of the
buckling mode (see [3.4]). Recalling from Chapter 2, Subsection 2.2.1
   L
EA L 2 EI L 2
D εx dx C x dx C P u,x dx 2.242
2 0 2 0 0

where
εx D u,x C 12 w,2x . 2.44a

It has been shown by Dym ([2.5], pp. 71 72) that if the column is assumed to
be incompressible during its bending from the straight line configuration, then the
deformed length of a line element dx Ł D dx, the undeformed length of the same
line element. Thus one must require
1 C u,x 2 C w,2x D 1 3.5a
1 1
or u,x D  w,2x  w,4x  Ð Ð Ð 3.5b
2 8
 
w,xx 1 2 3 4
x D  D w,xx 1 C w,x C w,x C Ð Ð Ð . 3.5c
1  w,2x 2 8

Thus the potential energy of an axially compressed, incompressible εx D 0,


slender column can be written
   
1 L 1 L P 4
D EIw,xx Pw,x  dx C
2 2
EIw,xx w,x  wx dx C Ð Ð Ð
2 2
3.6
2 0 2 0 4
To determine the characteristic form of  for a particular equilibrium configuration
w D w0 one examines the change in the total potential energy  corresponding
to an arbitrary virtual displacement w1 of the structure. Thus let

w D w0 C w1 . 3.7
Asymptotic Imperfection Sensitivity Analysis 137

Substitution and regrouping yields


 D   0 D υ C 12 υ2  C 1 3
3! υ  C 1 4
4! υ  C ÐÐÐ 3.8
where for an initially straight column w0 D 0 and
υ D 0  equilibrium condition 3.9a
 L
1 2 1
υ D 2
EIw1,xx  Pw1,x
2
 dx 3.9b
2 2 0
1 3
υ D0 3.9c
3!
  
1 4 1 L P 4
υ D 2
EIw1,xx 2
w1,x  w1,x dx 3.9d
4! 2 0 4
....
The stability equation for the determination of the classical buckling load may
now be obtained from the second variation expression υ2  by Trefftz’ criterion
υυ2  D 0 3.10
yielding for simply supported boundary conditions w1 D w1,xx D 0 at x D 0, L
Euler’s problem (see Eq. (2.2)). The lowest eigenvalue is given by
EI
Pc D
2 3.11a
L2
and the corresponding buckling mode is

x
w1 D wO D C sin . 3.11b
L
An equation for the secondary (postbuckling) equilibrium path may be obtained
from the expression for  by application of the stationary potential energy crite-
rion. However, the resulting differential equation will be nonlinear in the finite,
incremental displacement component w1 .
On the other hand, for points on the postbuckling equilibrium path sufficiently
close to the bifurcation point the incremental displacement component w1 is of the
form of the classical buckling mode w. O Thus, by limiting the range of validity of
the postbuckling analysis to a sufficiently small neighborhood of the bifurcation
point, one can assume that the small finite displacement component w1 is of the
form of the buckling mode. Thus using Eq. (3.11b) an approximate expression
for the total potential increment  can be obtained by evaluating the integrals
indicated. Thus
 L  4  2 
2
2
x 2
2
x
υ D
2
EIC sin  PC cos dx
0 L L L L
   
L 
2
2
D EI  P C2 3.12
2 L L
138 Postbuckling Behavior of Structures


 
6 
L
x
x P 4 
4
x
υ4  D 12 EIC4 sin2 cos2  C cos4 dx
0 L L L 4 L L
 
3L 
4 
2
D 4EI  3P C4 3.13
8 L L
Hence
1 2 1 1
2
4
 D υ  C υ3  C υ 4  C Ð Ð Ð D Pc L1   2 C Pc L
2 3! 4! 4 64
 
P
ð 43 4 C Ð Ð Ð 3.14
Pc

where  D P/Pc , Pc D EI


/L2 and  D C/L close to the bifurcation point is a
small quantity. Notice that here c D 1 and hence because of Eq. (3.1) close to the
bifurcation point the following expression holds
P 
D D 1 C a C b 2 C Ð Ð Ð . 3.15
Pc c
Thus upon substitution and regrouping Eq. (3.14) becomes


2
4
 D Pc L1   2 C Pc L 4 C O 5 . 3.16
4 64
Equilibrium along the postbuckling path implies that

υ D  Ð υ D 0. 3.17
∂
Hence for υ 6D 0


2
2
 D 1   C 2 Pc L D 0 3.18
∂ 8 2

this implies for  6D 0 that



2 2
1C  D0 3.19
8
or recalling that here c D 1, one can also write


2
D 1 C 2 . 3.20
c 8
Comparing this expression with Eq. (3.1) one concludes that the axially compressed
simply supported slender column has a stable, symmetrical postbuckling behavior
with

2
a D 0 and b D D 1.2337 > 0.
8
Asymptotic Imperfection Sensitivity Analysis 139

Figure 3.5 Axial load vs lateral deflection for a column

As can be seen in Figure 3.5 the asymptotic solution based on Eq. (3.20) compares
favorably with the rigorous postbuckling solution of the elastica of [2.5] for values
of  < 0.3.

3.2.2 Initial Postbuckling Behavior of Plates

In the following a formal procedure for obtaining the equations governing the
buckling and postbuckling states is presented. This procedure was developed by
Budiansky and Hutchinson [3.5] based on the original work by Koiter [2.36]. For
simplicity the derivation will be presented for the case of isotropic elastic plates
under in-plane edge loads (see Figure 2.3 for the sign convention used).
Using an Airy stress function such that Nx D f,yy , Ny D f,xx and Nxy D f,xy
the nonlinear von Kármán Donnell type governing equations are
Dr4 w  f,yy w,xx  2f,xy w,xy C f,xx w,yy  D 0 3.21a
r f
4 2
Ehw,xy  w,xx w,yy  D 0. 3.21b
For the sign convention shown in Figure 2.3 one has the following strain-
displacement relations
εx D u,x C 12 w,2x x D w,xx
εy D v,y C 12 w,2y y D w,yy 3.22
xy D u,y C v,x C w,x w,y xy D 2w,xy .
For isotropic plates the constitutive equations can be written as
Nx D Cεx C εy  Mx D Dx C y 
140 Postbuckling Behavior of Structures

Ny D Cεy C εx  My D Dy C x  3.23


1   1  
Nxy D C xy Mxy D D xy
2 2
where
Eh Eh3
CD and D D 3.24
1  2 121  2 
are the extensional and bending stiffnesses, respectively.
Assuming that the eigenvalue problem for the buckling load Nc will yield a
unique buckling mode w1 with the associated stress function F1 , a solution
valid in the initial postbuckling region is sought in the form of the following
asymptotic expansions

D 1 C a C b 2 C Ð Ð Ð 3.25a
c
w D W0 C W1 C  2 W2 C Ð Ð Ð 3.25b
f D F0 C F1 C  2 F2 C Ð Ð Ð 3.25c
where W1 will be normalized with respect the plate thickness h and W2 is
orthogonal to W1 in some appropriate sense.
A formal substitution of this expansion into the governing equations and
regrouping by powers of the small parameter  generates a sequence of equations
for the functions appearing in the expansions.
Notice that by assuming that the unloaded edges are free to expand, the following
membrane prebuckling state

x D F,yy D N0
N0 D Ny0 D 0
0 0
Nxy 3.26a
u0 D εx0 x v0 D εy0 y w0 D 0 3.26b
where
N0 N0
εx0 D εy0 D 3.27
hE hE
satisfies the governing equations of the zero order state identically. Further, the
equations governing the first order state are reduced to the following set of
linearized stability equations
Dr4 W1 C N0 W,xx
1
D0 3.28a
r4 F1 D 0. 3.28b
These uncoupled, constant coefficient partial differential equations admit separable
solutions of the form
x y
W1 D Wmn sin m
sin n
3.29a
a b
F1 D 0. 3.29b
Asymptotic Imperfection Sensitivity Analysis 141

Notice that the expression assumed for W1 satisfies simply supported boundary
conditions at all edges. Substitution into Eq. (3.28a) yields the characteristic
equation with the eigenvalues

2
N0,mn D D kmn 3.30
b2
where  2
mb a
kmn D C n2 . 3.31
a mb
Notice that the critical buckling load Nc is obtained when for a given plate
aspect ratio a/b the plate buckling coefficient kmn assumes its minimum value. As
discussed in Chapter 2 the minimum value of kmn occurs for n D 1 and different
integer values of m, which depend on the specified plate aspect ratio a/b. See
Figure 2.4 for further details.
P
Introducing the classical plate buckling load Nc as
 2
p p 4
2 E h
Nc D c h D Ðh 3.32
121   b
2

then in Eq. (3.25a)


N0 Nc
D p and c D p 3.33
Nc Nc
Notice that in this case, for sufficiently long (say a/b > 1) simply supported plates
c D 1.
The equations governing the postbuckling or second order fields are
Dr4 W2 C Nc W,xx
2
D0 3.34a
Eh 2  m
2  n
2
r4 F2 D EhW,xy
1 1
W,xy W,xx
1
yy  D
W,1 W
2 mn a b
 x y
ð cos 2m
C cos 2n
. 3.34b
a b
These equations admit separable solutions of the form
W2 D 0 3.35a
x y
F2 D A1 cos 2m
C A2 cos 2n
3.35b
a b
where
Eh  n a 2 2
A1 D Wmn 3.36a
32 m b
 
Eh m b 2 2
A2 D Wmn . 3.36b
32 n a
142 Postbuckling Behavior of Structures

General expressions for the postbuckling coefficients a and b have been derived
by Budiansky and Hutchinson [3.5]. Alternative derivations of a and b are
presented, among others, in [3.3], [3.4] and [3.6]. For the case under consideration
these expressions reduce to
 
3
aD F,xx W,y W,y 2F,xy W,x W,y CF,yy W,x W,x
1 1 1 1 1 1 1 1 1
dS
O
2 S
3.37
  
1
bD 2 1
F,xx y W,y 2F,xy W,x W,y CW,y W,x
W,1 2 1 1 2 1 2
3.38
O S

C F,yy W,x W,x
1 1 2
dS C xx W,y W,y 2F,xy W,x W,y
F,2 1 1 2 1 1
S

C F,yy
2
W,1 1
x W,x dS

where

O D
 P xx
F, 0
W,1
y W,1
y 2 P
F, 0
xy W,1
x W, 1
y C P
F, 0
yy W,1
x W,1
x dS 3.39
S
∂ 
P D . 3.40
∂
Notice that for the membrane prebuckling state specified by Eq. (3.26a)
P 0
F, P 0
xx D F,xy D 0;
P yy
F, 0 p
D Nc . 3.41

Thus Eq. (3.39) reduces to



2 ab
O D Nc

p
W2mn m 3.42
a 4
Since by Eq. (3.29b) F1 D 0 and by Eq. (3.35a) W2 D 0, the evaluation of the
postbuckling coefficients is greatly simplified. In this case Eqs. 3.37 and 3.38 yield
aD0 3.43
 2

1 Eh
2 2 mb  a 2
bD p W C n4 . 3.44
Nc 16 b2 mn a mb

Assuming that the buckling mode given by Eq. (3.29a) is normalized to one by
p
the wall-thickness h, then Wmn D h. Substituting for Nc from Eq. (3.32) one gets
 

3 mb 2  a 2
b D 1  2
Cn 4
. 3.45
16 a mb

Thus a simply supported isotropic elastic plate under in-plane edge load has
a stable, symmetrical behavior. For a square plate a/b D 1 and the initial
Asymptotic Imperfection Sensitivity Analysis 143

Figure 3.6 Uni-axial compression vs lateral deflection for a square plate

postbuckling behavior is described by



D 1 C b 2 3.46
c
where using D 0.3
3
bD 1  2  D 0.34125 > 0. 3.47
8
As can be seen in Figure 3.6 for  < 1.0 (say) the asymptotic solution based on
Eq. (3.46) agrees closely with the solution of the geometrically nonlinear theory
of plates of [2.14].

3.2.3 Initial Postbuckling Behavior of Shells

The imperfection sensitivity of eccentrically stiffened cylindrical shells under


combined axial compression, external (or internal) pressure and torsion can also be
investigated within the context of Koiter’s theory. Figure 3.7 displays the notation
and sign convention used. Notice that the out-of-plane displacement w is taken to
be positive outward.
The nonlinear Donnell type equations appropriate for eccentrically stiffened
cylindrical shells have been derived by different authors (see, for instance, [3.7],
[3.8] and [3.9]). Written in terms of w and f (an Airy stress function) these
equations are
1 1
LH f  LQ w D w,xx  LNL w, w 3.48a
R 2
1
LD w C LQ f D  F,xx C LNL f, w C p 3.48b
R
144 Postbuckling Behavior of Structures

Figure 3.7 Notation and sign convention for eccentrically stiffened shells

where the linear differential operators are


LD   D Dxx  ,xxxx C Dxy  ,xxyy C Dyy  ,yyyy
LH   D Hxx  ,xxxx C Hxy  ,xxyy C Hyy  ,yyyy 3.49
LQ   D Qxx  ,xxxx C Qxy  ,xxyy C Qyy  ,yyyy
and the nonlinear differential operator is
LNL S, T D S,xx T,yy  2S,xy T,xy C S,yy T,xx . 3.50
Subscripts following a comma denote partial differentiation. These equations are
applicable to ring- and stringer stiffened shells if the stiffener properties are
“smeared out” to arrive at effective bending, stretching and eccentricity coupling
stiffnesses for the skin-stiffener combination. The parameters Dxx , H,xx , Q,xx , . . .,
etc. are listed in [3.10].
To calculate the postbuckling coefficients a and b for the case where a unique
buckling mode W1 and F1 corresponds to the critical (lowest) buckling load c ,
one begins by assuming a solution valid in the initial postbuckling regime in the
form of the asymptotic expansions given by Eq. (3.25). A formal substitution of
this expansion into Eq. (3.48) and regrouping by powers of  generates a sequence
of linear equations for the functions appearing in the expansion.
If one neglects the effect of the prebuckling edge constraints, then the following
membrane prebuckling solution
W0 D hW C Wp C Wt  3.51a
 
Eh 1 2 1
F D
0
 y  pN e x  N xy
2
3.51b
cR 2 2
satisfies the governing equations of the zero order state identically. The quan-
tities W , Wp and Wt are evaluated by enforcing the periodicity condition (see
[3.6] for details). Furthermore,  is the nondimensional axial load parameter  D
cR/Eh2 N0 , pN e is the nondimensional external pressure pN e D cR2 /Eh2 pe  and
N is the nondimensional torque parameter N D cR/Eh2 Nxy , positive counter-
clockwise.
Asymptotic Imperfection Sensitivity Analysis 145

The set of equations for W1 and F1 yields the following classic eigenvalue
problem
1 1
LH F1   LQ W1  D W, 3.52a
R xx
1 Eh2  
LD W1  C LQ F1  D  F,1  W,1
2N
 W, 1
C N
p e W,1
3.52b
R xx cR xx xy yy

In Eq. (3.52b) the user can select the eigenvalue c to be the critical value of
either the normalized axial load , or the normalized external pressure pN e or the
normalized torque N . The remaining two load parameters are then held fixed.
Approximate solutions of this standard eigenvalue problem have been presented
by many authors. See, for instance, Hutchinson and Amazigo [3.7], Seggelke and
Geier [3.11], Block et al. [3.12] and Khot and Venkayya [3.13], just to name a
few.
From the next higher-order terms in the expansion one gets the governing
equations for W2 and F2
1 2
LH F2   LQ W2   W, D W,xy
1 1
W,xy W,xx
1 1
W,yy 3.53a
R xx
1 2 Eh2
LQ F2  C LD W2  C F, C 2
W,xx 2N W,xy
2
CpN e W,yy
2
R xx cR
D LNL F1 , W1  3.53b

whereby one of the three load parameters , pN or N is the eigenvalue c selected


above when solving the classic eigenvalue problem.
The same general expressions for the postbuckling coefficients a and b, quoted
earlier as Eqs. (3.37) and (3.38), are also valid for this case. It is easily verified that
the first postbuckling coefficient a is identically zero for an asymmetric buckling
mode. Therefore, it is necessary to solve for W2 and F2 in order to calculate
the second postbuckling coefficient b. It is shown in [3.7] that Eqs. (3.53a) and
(3.53b) admit separable solutions of the form
 
 1
x y 1
x
W2 D h Aj sin j
C cos 2n Bj sin j
3.54
 L R jD1 L
jD1
 
3  1 1 
Eh x y x
F2 D Cj sin j
C cos 2n Dj sin j
3.55
2c  jD1
L R jD1
L

where c D 31  2 . The coefficients are determined by the Galerkin procedure.
Notice that each individual term in the series in Eqs. (3.54) and (3.55) satisfies
simply supported boundary conditions W2 D W,2 xx D F
2
D F,2
xx D 0 at x D
0, L. Finally, the postbuckling coefficient b is calculated by evaluating the integrals
indicated in Eq. (3.38) to obtain
146 Postbuckling Behavior of Structures
 
c Q  1
 1
1
 j 
bD N j C Dj  C 2
2FB N j C Cj 
2FA 3.56

c jD1,3,... j jD1,3,...
4m2  j2 

Q depends on the choice of c .


where the value of 
Thus for axial compression (c D , both pN e and N fixed)

Q D 2˛m C ˛p  ˇn .
2 2
 3.57
˛2m C ˛2p 

Notice that for a specified internal pressure pN e D pN i  the eigenvalue must be
replaced by O c D c  21 pN i . Further for external lateral pressure (c D pN e , both 
and N fixed)
Q D ˛m C ˛p 2 3.58

whereas for hydrostatic pressure (c D pN e ,  D 21 pN e and N fixed)

Q D 2˛m C ˛p 2 ˇn2
 . 3.59
2 ˛m C ˛p  C 2ˇn
1 2 2 2

Finally for torsional loading (c D N , both  and pN e fixed)

Q D ˛m C ˛p  ˇn
2
 3.60
˛m  ˛p
where

n 2 Rh
˛2m D m C K
L R 2c

n  Rh
˛2p D m  K 3.61
L R 2c
 n 2 Rh
ˇn2 D .
R 2c
Notice that in these expressions m, n are integers denoting the number of half waves
and the number of full waves in the axial and in the circumferential directions,
respectively, whereas K is a real number called Khot’s skewedness parameter
[3.13] denoting the inclination of the nodal lines of the buckling pattern with
respect to the axis of the shell.
The series in Eq. (3.56) can be evaluated numerically to any degree of accuracy
desired. This solution was first obtained by Hutchinson and Amazigo in 1967 (see
[3.7]) using an asymmetric imperfection in the form of the critical buckling mode
x n
N D hN2 sin m

W cos y  K x 3.62
L R
where in the absence of torsional loading K D 0. The coefficients Aj , Bj , Cj and
Dj are listed in [3.14].
Asymptotic Imperfection Sensitivity Analysis 147

Knowing b one can use Eq. (3.4) and the condition for the occurrence of a limit
point
d
D0 3.63
d

to obtain a relation between the limit load s of the imperfect structure and the
bifurcation load c of the perfect structure. Notice that for a D 0 straightforward
calculations yield the formula
 
s 3/2 3 p s
1 D 3bjN2 j for b < 0 3.64
c 2 c

first presented by Koiter in 1945 [2.36]. Thus, for symmetrical bifurcation (a D 0)


if the second postbuckling coefficient b is negative, the equilibrium load decreases
following buckling and the buckling load of the real (imperfect) structure s is
expected to be imperfection sensitive.
Since the sixties many papers have been published dealing with the imperfection
sensitivity of different shells of revolution loaded by various types of external loads.
Following the standard set by the Harvard group under Budiansky and Hutchinson
it has become a widely accepted practice to display the results of such investigations
in the form shown in Figure 3.8. By plotting the normalized buckling load c and
the corresponding second
p postbuckling coefficient b versus Batdorf’s Z-parameter,
where Z D L /Rh 1  2 , it is possible to display both the critical buckling
2

load and a measure of its imperfection sensitivity for a wide range of possible
shell configurations in a single figure.
When trying to assess the imperfection sensitivity of the critical buckling load
one must remember, that this form of the so-called b-factor method can only be
applied in cases of symmetric bifurcation and when just a single non-axisymmetric
buckling mode is associated with the critical buckling load. To estimate the degree
of imperfection sensitivity as a function of the magnitude of the postbuckling
coefficient b, one can use the curves shown in Figure 3.9. For the sake of calibration
a curve taken from Koiter [2.36], showing the effect of axisymmetric imperfections
on the buckling load of axially compressed isotropic cylinders, is also included in
Figure 3.9. For authoritative reviews and for more detailed results the interested
reader should consult [3.15] and [3.16].
Recent investigations [3.17] and [3.18], have shown that the trends predicted
by a Koiter type asymptotic imperfection-sensitivity analysis are reliable, if in the
calculation of the field functions needed to evaluate the postbuckling coefficients
one employs rigorous nonlinear prebuckling analysis and satisfies the appropriate
boundary conditions exactly. Furthermore, it can be stated that with the availability
of computer codes like DISDECO [3.18], SRA [3.19] and FASOR [3.20] and with
the current generation of high-speed desk-top workstations, it has become feasible
for all structural engineers to use Koiter’s Imperfection Sensitivity Theory in every
day design practice (see also Sub-section 3.3.5 on this topic).
148 Postbuckling Behavior of Structures

Figure 3.8 Classical buckling load and imperfections sensitivity of simply supported, stringer
stiffened cylinders under axial compression (As /ds h D 0.506, Els /Dds D 2.69,
es /h D 1.71, GJs /Dds D 3/40)

3.2.4 Experimental Verification

The best known experimental verification of the predictions of the asymptotic


imperfection sensitivity theory is due to Roorda [3.21]. He tested the two-bar frame
shown in Figure 3.10. Since a real structure is never totally free of imperfections
the vertical leg tended to bend to the right or to the left as soon as the load
was applied. Roorda found that by placing the load at a distance q0 D 0.0013L
Asymptotic Imperfection Sensitivity Analysis 149

Figure 3.9 Variation of buckling load with imperfection amplitude for various values of b

to the right of the centerline of the vertical leg he could practically eliminate this
tendency. Provisions were then made to apply the load at any distance q to the
right or the left of the centerline of the vertical leg. In the test program the frame
loaded by a slightly eccentric load applied at q 6D q0 represented an equivalent
slightly imperfect frame, where the load eccentricity q  q0  played the role of
N
the initial imperfection L.
In the tests the rotation A of joint A was measured optically and it was used
as the displacement parameter in plots of the equilibrium paths (thus  D A ). In
Figure 3.11 results of two of the tests for the smallest eccentricity jq  q0 j that
could be achieved are displayed where  D P/Pc and
EI
Pc D 1.406
2 3.65
L2

Figure 3.10 Two-bar frame subjected to eccentric load


150 Postbuckling Behavior of Structures

Figure 3.11 Comparison of theoretical and experimental equilibrium paths (from [3.21])

is the classical buckling load of the two-bar frame shown in Figure 3.10. Negative
values of A represent counter-clockwise rotations of joint A. For small values of A
one can see the excellent agreement between the experimental points and the asym-
metric equilibrium paths predicted by the asymptotic theory. For counter-clockwise
rotations the equilibrium path exhibits a limit point at about s D 0.99. Experi-
mental limit loads s for values of q < q0 , which produces a counter-clockwise
initial rotation, were obtained by applying the load at other locations. Figure 3.12
displays a comparison between the experimental limit loads s plotted as a function
of the load eccentricity ratio q/L with the locus of the limit point predicted by the
asymptotic theory
1  s  C 4a Ň 
N sD0 3.66

where Ň is an equivalent imperfection form factor. From the theoretical solution


of [2.2] one obtains

Figure 3.12 Comparison of theoretical and experimental limit point loci (from [3.21])
Asymptotic Imperfection Sensitivity Analysis 151

a D 0.380 and Ň D 0.871. 3.67


Using these numerical values in Eq. (3.66) to obtain the solid line in Figure 3.12
one sees that once again the agreement is excellent. The curve is seen to have
the characteristic parabolic form with a vertical tangent at N D q  q0 /L D 0.
Thus in this figure the ratio N D q  q0 /L represents an equivalent imperfection
parameter.
Turning now to more complicated structures it soon becomes evident that there
are less results available and that due to the mathematical complexities of the
asymptotic theory extrapolation of the published results to more general structural
configurations requires considerable theoretical and practical insight and experi-
ence.
Considering again the postbuckling equilibrium paths in Figure 3.2 for an axially
compressed initially perfect cylindrical shell and one of the possible curves for a
slightly imperfect shell, one can conclude the following

1. The bifurcation buckling load of the perfect structure represents the ultimate
load carrying capability of the structure.
2. The collapse load of the imperfect structure may be considerably lower than
the bifurcation buckling load of the perfect structure.
3. The collapse loads of nominally identical structures may vary widely due to
the random nature of the initial deviations (imperfections) from the perfect
shape.

That the scatter in buckling loads for nominally identical cylindrical shells is
caused by the small unintentional differences in the initial shape can also be
deduced from the results displayed in Figure 3.13. The upper part contains the
results of a large number of buckling experiments under hydrostatic pressure
marked by circles and a solid line depicting the theoretical buckling pressure for
perfect cylinder. Notice that for Z-values between 10 and 100 there are larger
deviations between the experimental results and the theoretical predictions than
for higher Z-values. This trend is also predicted by the second postbuckling coef-
ficient b, plotted in the lower part, since for decreasing values of Z the value of b
becomes more negative thus indicating increasing imperfection sensitivity. In the
upper part of Figure 3.13 the following normalized external pressure is used
L2R Eh3
pN D pe where D D . 3.68

2 D 121  2 
The degree of imperfection sensitivity depends not only on the shell geometry
and on the boundary conditions used but also on the external load applied to the
shell. In the upper part of Figure 3.14 the results of a large number of buckling
experiments under torsion are shown marked by circles. Also included is a solid
line depicting the theoretical buckling coefficient for perfect cylinders. Notice that
the second postbuckling coefficient b, plotted in the lower part, predicts that for all
values of Z larger than, say five, the buckling loads are sensitive to initial imper-
fections. These predictions are for the most part confirmed by the test results of the
152 Postbuckling Behavior of Structures

Figure 3.13 Comparison of theoretical and experimental values for isotropic cylinders
subjected to hydrostatic pressure

upper part. Unfortunately there is little experimental data available for shell geome-
tries with Z-values between 10 and 50 where maximum imperfection sensitivity
is predicted. In the upper part of Figure 3.14 the following normalized torsional
buckling coefficient is used
L2 h
Kt D 2 c 3.69

D
where c D Nxy /h.
Asymptotic Imperfection Sensitivity Analysis 153

Figure 3.14 Comparison of theoretical and experimental values for isotropic cylinders
subjected to torsion

Considering the results of the correlation studies presented, one sees that for
those structures that are insensitive (or not very sensitive) to initial imperfections
there is a good agreement between the predictions of the asymptotic theory and
the available experimental results. Even for imperfection sensitive structures like
thin-walled shells the asymptotic theory appears to predict the trend correctly. Thus
a question comes automatically to one’s mind: “Why is the asymptotic imperfec-
tion sensitivity theory not used more often in practical applications?” Partially
the answer may lie in the fact that whereas in many cases the predictions of the
154 Postbuckling Behavior of Structures

asymptotic theory have been confirmed by subsequent full nonlinear solutions to


be accurate up to imperfection amplitudes of the order of one shell wall thick-
ness, there are also examples where the range of validity of the asymptotic theory
is too small to be of any practical value as has been shown for oval cylinders
[3.22] and elliptical cones [3.23] under axial compression. Another reason why
the asymptotic imperfection sensitivity theory is not used more often is because
very little is known about the deviations from the nominal shape of real life struc-
tures such as aircraft fuselages, launch vehicle shells, submarine hulls, silos, large
containment vessels, off-shore shells etc. It is encouraging to see that the need for
detailed initial imperfection surveys on large scale and full scale structures and the
establishment of Initial Imperfection Data Banks is being recognized by a growing
number of investigators. It is the authors opinion that the existence of extensive
data on characteristic initial imperfection distributions classified according to fabri-
cation processes is one of the prerequisites for better and more reliable buckling
load prediction. This point is discussed in more detail in Chapter 10, Volume 2.

3.3 Direct Solutions of the Nonlinear Stability Problem


As has been pointed out earlier, in many cases one can use Koiter’s theory to
make a fairly accurate estimate of the initial postbuckling behavior of real (read
imperfect) structures. Unfortunately, the validity of the information provided by
the asymptotic approach is strictly speaking restricted to the immediate neighbor-
hood of the corresponding bifurcation point. Thus, in order to establish the range
of validity of the asymptotic analysis, or whenever the shape of the (secondary)
equilibrium path in the more advanced postbuckling region is needed, one must
solve the nonlinear stability problem. Such will be the case, for instance, if both
geometric and material nonlinearities are included in the analysis.

3.3.1 Elastic Postbuckling Behavior of Columns

The equation governing the large displacement deformations of elastic columns


can best be derived by using the stationary potential energy criterion. For slender
columns one can assume that the column is incompressible during its bending from
the initial straight line configuration. Thus its potential energy may be written

 D U b C p 3.70

where Ub and p are the bending energy and the potential of the applied load
(see Eqs. (2.243b) and (2.243c)), respectively. As can be seen from Figure 3.15
and as has been described in detail in [2.5] on pp. 71 72, in view of the condition
of incompressibility
u,x D cos   1. 3.71
Direct Solutions of the Nonlinear Stability Problem 155

Figure 3.15 Geometry of deformed column

Further, the true curvature of the deformed curve is


1 d
x D D . 3.72
R dx
Thus Eq. 3.70 becomes
 L  2  L
EI d
D dx C P cos   1 dx. 3.73
2 0 dx 0

A straightforward use of the variational statement of equilibrium

υ D 0

yields the following two-point boundary value problem


d2 
EI C P sin  D 0 for 0  x  L 3.74a
dx 2
d
M D EI D 0 at x D 0, L. 3.74b
dx
The solution of this equation is rather involved because of the nonlinearity inherent
in the term sin . Solutions in terms of complete elliptic integrals of the first kind
Kq have been presented in [2.1] and [2.5]. For a simply supported column one
gets  
P 2
/2 d 2
D  D Kq 3.75
PE
0 1  q2 sin2 

where
˛
q D sin
2
156 Postbuckling Behavior of Structures


q sin  D sin 3.76
2

2
PE D EI
L2
and ˛ is the unknown slope at the column ends. One can also calculate the
maximum deflection in terms of P/PE yielding
 w 2  wx D L/2 2  2 2  P 
max
D D q2 . 3.77
L L
PE
For details of the solution the interested reader should consult [2.5].
Equations [3.75] and [3.77] can be used to plot the curve labelled “exact” solution
in Figure 3.5.
Considering this solution curve, notice that it is tangent to the horizontal line
at /c D 1, where the deflection is zero  D 0. Thus an increase in the axial
compression P (or ) corresponding to a small increment in the buckling deflection
w1 (or ) is a small quantity of second order. This explains why when one uses
the linearized stability equations (Eqs. (2.2) or (2.256b)) to calculate the critical
buckling load Pc , the buckling deflection w1 D wO is found to be indefinite. Notice
that, as indicated also in Figure 2.2, the solution curve represented by Eq. (3.77)
can only be used up to the proportional limit of the material. Beyond this limit
the resistance of the column to bending diminishes and in order to obtain the
proper postbuckling solution curve the inelastic behavior of the material must be
accounted for.

3.3.2 Plastic Postbuckling Behavior of Columns


In Chapter 2 the plastic buckling case of an axially compressed column has been
formulated as a bifurcation problem leading to the well-known “reduced modulus”
buckling formula of Considére Engesser von Kármán (see Eq. (2.230)). To inves-
tigate the postbuckling behavior of the column, the problem must be reformulated
as a response problem. This will be done at some length following basically von
Kármán’s approach [2.76]. It serves as a lucid account of how one can combine the
results of carefully done experiments with physical insight to arrive at an useful
engineering solution of a pressing problem.
Using the sign convention shown in Figure 3.15 one sees that if the applied
compressive force has a slight eccentricity e, one has bending and axial compres-
sion acting simultaneously. Assuming in the plastic analysis that plane sections
remain plane also during plastic deformation results in a linear distribution of the
normal strain. Thus the strain at any point is
z
ε D ε0 C 3.78
R
where ε0 is the strain caused by the centrally applied axial load P. Furthermore,
in the plastic analysis the normal stress distribution is given by the stress-strain
curve of the material used, shown here in Figure 3.16.
Direct Solutions of the Nonlinear Stability Problem 157

Figure 3.16 Stress-strain diagram

Notice that the position of the neutral axis does not necessarily coincide with
the centroidal axis. Its position is determined by the values of ε1 and ε2 , the
elongation and contraction of the extreme fibers and the value of ε0 . By using the
static equilibrium equations

 dA D P 3.79
A

z dA D M D Pυ0 3.80
A
one can calculate the position of the neutral axis and the radius of curvature R.
Introducing the notation
h 1 h2 h
 D ε1  ε2 D C D 3.81
R R R
and recalling from Eq. (3.78) that dz D Rdε one can rewrite Eq. (3.79) as
 h1  ε1 
h ε1
P D b  dz D bR  dε D b  dε. 3.82
h2 ε2  ε2
Dividing by the cross-sectional area bh one obtains the average compressive stress
c as 
P 1 ε1
c D D  dε. 3.83
bh  ε2
Notice that the integral in this expression represents the shaded area under the
stress-strain curve in Figure 3.16. Equation (3.83) can be used to calculate the value
of ε2 corresponding to any assumed value of ε1 provided the axial load P is known;
or one can assume both ε1 and ε2 and calculate the corresponding value of P.
Using the same notation one can rewrite also Eq. (3.80) as
 h1  ε1   
12 ε1 I
MDb z dz D bR 2
ε  ε0  dε D 3
ε  ε0  dε 3.84
h2 ε2  ε2 R
where for a rectangular cross-section I D bh3 /12. As can be seen, the integral
in this expression represents the first moment of the shaded area for the given
158 Postbuckling Behavior of Structures

stress-strain diagram with respect to the vertical axis A-A in Figure 3.16. Since
the ordinates in the figure represent stresses and the abscissas represent strains the
integral has the same dimensions as the modulus E. Thus Eq. (3.84) can be put in
the form
I
M D EQ 3.85
R
where  ε1
12
Q D
E ε  ε0  dε. 3.86
3 ε2

The magnitude of E Q for a given material is function of  D ε1  ε2 . By varying


ε1 and ε2 in such a manner that c calculated from Eq. (3.83) remains constant,
one obtains EQ as a function of  for any given value of c . The resulting relations
were calculated by Timoshenko and Gere for a given structural steel (see [2.1],
p. 170) and are presented here in Figure 3.17. Using these curves with Eq. (3.85)
one can calculate the bending moment M as a function of  D ε1  ε2 D h/R for
any given value of c . The resulting curves are displayed in Figure 3.18.
Next, using the curves of Figure 3.18 the shape of the deflection curve for an
eccentrically loaded column can be obtained from Eq. (3.83) by numerical inte-
gration. The details of this method are discussed at some length in von Kármán’s
paper of 1910 (see [2.76]) and shall not be repeated here. The calculations are best
carried out in terms of dimensionless ratios. Thus finally one obtains for specified

Figure 3.17 Q vs  for specified values of c (from [2.1])


Equivalent plastic modulus E
Direct Solutions of the Nonlinear Stability Problem 159

Figure 3.18 Equivalent plastic bending moment vs  for specified values of c (from [2.1])

values of c and eccentricity ratio e/h, the deflection wmax /h as a function of the
normalized length L/h of the columns. Several curves of this type, calculated by
von Kármán [2.76] for steel having a yield stress of about 45 000 psi using various
values of c and e/h D 0.005, are shown in Figure 3.19. Notice that instead of
values of L/h, values of the slenderness ratio L/ are used as ordinates in this
figure. From the points of intersection of horizontal lines and the curves, desig-
nated as points M, N and Q, one obtains a relation between the direct compressive
stress c and the deflection wmax for a given slenderness ratio L/ and an assumed
load eccentricity e.
Cross-plotting these values von Kármán obtained the curves shown in
Figure 3.20 using different values of load eccentricity and a slenderness ratio

Figure 3.19 Variation of deflection with slenderness ratio for specified values of c (from
[2.76])
160 Postbuckling Behavior of Structures

Figure 3.20 Postbuckling curves for an imperfect column including material nonlinearities
(from [2.76])

of L/ D 75. Notice that in this figure for any eccentricity the load initially
increases with increasing deflection. However, contrary to the elastic case, where
the curves with different load eccentricities approach the same critical buckling
load asymptotically (see also Figure 2.2), in the plastic buckling case the maximum
load carrying capacity of the column is noticeably decreased even for small values
of load eccentricity. Thus the critical buckling load of an axially compressed
column becomes imperfection sensitive if buckling occurs at stress levels higher
than the yield point of the material, whereby the load eccentricity e is the initial
imperfection. Notice further, that beyond the limit load a column can carry for a
given slenderness ratio and assumed load eccentricity, increase in deflection will
proceed with a diminishing of the load.

3.3.3 Postbuckling Behavior of Plates

Thin-walled plates are widely used in all branches of engineering technology. Their
popularity arises from the fact that a slender compressed plate is able to support
loads greater than that which causes the plate to buckle. The postbuckling strength
exhibited by thin-walled panels often leads to the use of structural elements with
compound cross-sections which operate within the postbuckling range under certain
loading conditions. In order to be able to exploit these characteristics of plate
structures optimally the designer must be aware of their postbuckling behavior.
To get an initial indication, one can use the results of the initial postbuckling
theories based on Koiter’s work. However, to verify the predictions of the asymp-
totic theory and to investigate the load carrying capacity of plates in the deep
postbuckling range one must solve the nonlinear equations directly.
Direct Solutions of the Nonlinear Stability Problem 161

a. Perfect Plates

The nonlinear von Kármán Donnell type equilibrium equations of a flat isotropic
plate under edge compression are (see, for instance, [2.2], p. 87)
Nx,x C Nxy,y D 0 3.87a
Nxy,x C Ny,y D 0 3.87b
D4 w  Nx w,xx C 2Nxy w,xy C Ny w,yy  D 0. 3.87c
By introducing the constitutive and kinematic relations (Eqs. (3.22) and (3.23)),
the in-plane equilibrium equations can be written in terms of the displacements u
and v as
1   1 C  1  
u,xx C u,yy C v,xy D w,x w,xx  w,x w,yy
2 2 2
1 C 
 w,y w,xy 3.88a
2
1   1 C  1  
v,xx C v,yy C u,xy D w,y w,yy  w,y w,xx
2 2 2
1 C 
 w,x w,xy 3.88b
2
For a square plate a D b the out-of-plane displacement w is given by
x y
w D W11 sin
sin
3.89
a a
an expression that satisfies simply supported boundary conditions at all edges. After
substituting for w on the right hand side of Eqs. (3.88a) and (3.88b) one can use
the method of undetermined coefficients to obtain the following particular integrals
(see Figure 2.3 and [2.5])
x x y
u D ε0 x C u20 sin 2
C u22 sin 2
cos 2
3.90
a a a
y x y
v D v0 y C v02 sin 2
C v22 cos 2
sin 2
3.91
a a a
where
 

W11 2
u20 D v02 D  1   a 3.92a
16 a
 

W11 2
u22 D v22 D a . 3.92b
16 a
The terms ε0 x and v0 y have been added in Eqs. (3.90) and (3.91) in order to allow
for constant in-plane displacements at the shell edges. Notice that ε0 is the applied
end-normal-strain at x D a. The total potential energy of the system can be written
(see, for instance, [2.2], p. 84)
 D Um C Ub C p 3.93
162 Postbuckling Behavior of Structures

where
 a 
C a 1   2
Um D εx C εy C εx εy C
2 2
xy dx dy
2 0 0 2
    
C a a 1 2 2 1 2 1 2
D u,x C w,x C u,x C w,x u,y C w,y
2 0 0 2 2 2

1  
C u,y C v,x C w,x w,y u,y C v,x C w,x w,y  dx dy 3.94a
2
  
D a a 2 1   2
Ub D x C y C x y C
2
xy dx dy
2 0 0 2
 
D a a
D w,xx C w,yy 2 dx dy 3.94b
2 0 0
and C and D are the extensional and bending stiffnesses, respectively (see
Eq. (3.24)). Notice that use has been made of the fact that if w D 0 at all four
edges of the plate the contribution of the Gaussian curvature term to the bending
strain energy is zero. Furthermore the potential of the external load is
 a a
 D aN0 [ua  u0] D N0 u,x dx dy. 3.94c
0 0

Substituting for u, v and w from Eqs. (3.89) (3.91), carrying out the integration
and regrouping, one obtains
 
1 Ea2 h
2 W11 2
D ε0  2 ε0 v0 C v0 C 1 C v0  ε0 
2 2
2 1  2 4 a
     


4 h 2 W11 2
4 W11 4
C C [4  1  2 ]  N0 a2 ε0 3.95
12 a a 64 a

For equilibrium the total potential energy  must be stationary; that is, its first vari-
ation must equal zero. Thus, remembering that the edge displacement jux D ajD
ε0 a is prescribed, it follows that
∂ ∂
υ D υv0 C υW11 D 0 3.96
∂v0 ∂W11
∂ ∂
implying that D D 0. This yields the following equations
∂v0 ∂W11
 

2 W11 2
v0 D ε0  1 C  3.97
8 a
  2  2   2 

W11
W11
2 h
 ε0  D 0. 3.98
a 4 a 31  2  a
Direct Solutions of the Nonlinear Stability Problem 163

The last equation obviously has two solutions. Either


 
W11
D0 3.99
a
or  2   2 

2 W11
2 h
 ε0  D 0. 3.100
4 a 31  2  a

Notice that this last equation implies that real solutions can exist only if the
prescribed end-normal-strain ε0 > εc , where
 2

2 h
εc D . 3.101
31   a
2

Thus one can distinguish three regimes. Initially, when the applied normal strain
ε0 < ε c
there is no normal deflection only shortening of the plate. This is the prebuckling
state where  
W11
D 0.
a
 
W11
Buckling occurs, that is 6D 0, whenever ε0 D εc . Finally, in the postbuck-
a
ling region, whenever
ε0 > ε c
the normal deflection is given by
 
W11 2p
Dš ε0  εc . 3.102
a

Calculating the corresponding stress fields one obtains for the prebuckling state

x D Ehε0 ;
N0 Ny0 D Nxy D 0.
0
3.103
At buckling, when ε0 D εc
 2
4
2 E h p
N1 D Nc D Ehεc D  h D Nc . 3.104
x
121  2  a
Finally, in the postbuckling region (where ε0 > εc )
   
Eh ε0 ε0 y
Nx D x h D εc 2 C  1 1  cos 2
. 3.105
2 εc εc a
Evaluating the average value of Nx one obtains for a square plate a D b

  h a Eh
Nx ave D hx ave D x dy D  ε0 C εc . 3.106
a 0 2
164 Postbuckling Behavior of Structures

Plotting from this equation x jave vs ε0 in Figure 3.21 one sees that the plane
continues to carry increased loading beyond buckling at one-half of the rate prior
to buckling. Thus at buckling the plate has lost only a part of its load carrying capa-
bility. This implies that a redistribution of the normal stress takes place. Looking
at Figure 3.22, where using Eq. (3.105) the distribution of the axial stress x as a
function of y is plotted for different values of end-shortening ratios ε0 /εc one sees
that after buckling the additional loading is carried by parts of the plate near the
edges. It is interesting that the center of the plate carries only the critical stress c
at which buckling occurs, even as the plate is being further compressed.
For efficient design the postbuckling strength of plates must be taken into
account. This has led to the concept of effective width beff , which denotes that
portion of the plate width b that is actively carrying the applied load. If one iden-
tifies that ratio beff /b with the ratio of the average stress after buckling for a given
value of ε0 , to that of the stress carried by the unbuckled plate at the same value
ε0 , then one obtains a convenient expression to calculate beff .

Figure 3.21 Average stress vs prescribed normal strain for a buckled plate

Figure 3.22 Redistribution of normal stresses in a buckled plate


Direct Solutions of the Nonlinear Stability Problem 165

Notice that since, by definition



x jave 1 E beff
D  ε0 C εc  D 3.107
x0 Eε0 2 b
therefore  
b εc
beff D 1C . 3.108
2 ε0
Thus, for the square plate at buckling where ε0 /εc D 1, beff D a; that is, the whole
plate is carrying the uniformly distributed applied load. On the other hand, in the
deep postbuckling region, say at ε0 /εc D 10, beff D 0.55a; that is, only a fraction
of the whole plate is actively carrying the applied load.
For design purposes it is convenient to express the results of the postbuckling
analysis in terms of an effective width beff over which the stress is considered to
be uniform, as shown in Figure 3.23b. Hence
Px D hbeff max 3.109
where max is the maximum stress at the plate edges y D 0, b. A widely used
approximate expression for beff is ([3.24], Eq. 7)
 
c 1/2
beff D b 3.110
max
where c is the classical critical stress for the given boundary condition. Using
the square plate considered above, and assuming that at the plate edges max /c D
ε0 /εc D 10, Eq. (3.110) yields
beff D 0.316a.

Figure 3.23 Alternate stress distribution in a buckled plate (from [2.2])


166 Postbuckling Behavior of Structures

More accurate expressions to calculate beff are to be found in [3.25] [3.27]. A


detailed description and a historical review of the concept of “effective width” is
presented in Chapter 8.

b. Imperfect Plates

To account for the effect of a small initial curvature (read, imperfection) one can
use the von Kármán-Donnell type imperfect plate equations
Dr4 w  ff,xx w,yy C wN ,yy   2f,xy w,xy C wN ,xy 
C f,yy w,xx C wN ,xx g 3.111
Eh
r4 f D fw,xx w,yy C 2wN ,yy   2w,xy w,xy C 2wN ,xy 
2
C w,yy w,xx C 2wN ,xx g 3.112
where the shape of the small initial imperfection is given by
x y
N y D W
wx, N mn sin m
sin n
3.113
a b
Assuming that the out-of-plane displacement w can be represented by the affine
expression
x y
wx, y D Wmn sin m
sin n
3.114
a b
then upon substituting and regrouping the compatibility equation (3.112) becomes
Eh  m
2  n
2  x y
r4 f D Wmn Wmn C 2W N mn  cos 2m
C cos 2n
.
2 a b a b
3.115
This equation admits a particular solution of the form
1 x y
fx, y D  N0 y 2 C A1 cos 2m
C A2 cos 2n
3.116
2 a b
where
Eh  na 2
A1 D N mn 
Wmn Wmn C 2W 3.117a
32 mb
 
Eh mb 2
A2 D Wmn Wmn C 2WN mn  3.117b
32 na
and N0 is the applied uniformly distributed compressive edge load.
Substituting the above expressions for w, N w and f in the out-of-plane equilibrium
Eq. (3.111) yields the residue εx, y; Wmn , which is then minimized by Galerkin’s
procedure. Evaluation of the integral involved
 a b
x y
εx, y; Wmn  sin m
sin n
dx dy D 0 3.118
0 0 a b
Direct Solutions of the Nonlinear Stability Problem 167

yields after some regrouping the following nonlinear algebraic equation


 
m
2  n
2 2  m
2
D C Wmn  N0 Wmn C WN mn 
a b a
  
Eh  m
2  n
2 m b 2  n a 2
C C N mn 
Wmn Wmn C W
16 a b na mb
N mn  D 0.
ð Wmn C 2W 3.119

Introducing the notation


p
N0 D c h N mn D hN
W Wmn D h 3.120

where  2
p 4
2 E h
c D 3.121
121  2  b
makes it possible to rewrite Eq. (3.119) as
  
3 mb 2  a 2
pc,mn   C 1  2  C n4 N C 2
 C  N D N
16 a mb
3.122
where  2
1 1 mb 2 a
c,mn D kmn D
p
Cn . 3.123
4 4 a mb

For a square plate Eq. (3.122) is then used to plot the curves for N > 0 shown
in Figure 3.6. More accurate solutions using a truncated double Fourier series
N w and f have been obtained by Lévy [3.28], Hu et al. [2.15]
representation for w,
and Coan [2.14].

3.3.4 Postbuckling Behavior of Circular Cylindrical Shells

In modern engineering design stiffened or unstiffened shells play an important role


when it comes to weight critical applications, since these thin-walled structures
exhibit very favorable strength over weight ratios. Unfortunately, they are also
prone to buckling instabilities.
In the following ways to obtain the equilibrium paths for initially perfect
and imperfect circular cylindrical shells subjected to axial compression will be
discussed.

a. Perfect Shells

In order to arrive at their pioneering results depicted in Figure 3.2, describing the
postbuckling behavior of axially compressed perfect isotropic cylindrical shells,
168 Postbuckling Behavior of Structures

von Kármán and Tsien [3.1] had to solve the nonlinear governing equations of the
problem. They employed the stationary potential energy criterion to derive a set
of three nonlinear algebraic equations in terms of the unknown amplitudes f0 , f1
and f2 of the out-of-plane displacement w. To obtain a plausible form for w von
Kármán and Tsien relied heavily on the known experimental results. They based
their analysis on the following assumed displacement function
  
w 1 1 x y 1 x 1 y
D f0 C f1 C f1 cos m cos n C cos 2m C cos 2n
R 4 2 R R 4 R 4 R
1  x y 
C f2 cos 2m C cos 2n . 3.124
4 R R
Notice that the term f0 C 1/4f1  allows for the shell to expand radially. Further,
by letting f0 D f2 D 0, f1 D 1, Eq. 3.124 reduces to

w 1 1 x y 1 x 1 y
D C cos m cos n C cos 2m C cos 2n
R 4 2 R R 4 R 4 R
mx C ny mx  ny
D cos2 cos2 3.125
2R 2R
which is the well-known diamond shaped pattern observed at large values of the
wave amplitude in the stable postbuckling region.
Notice further, that by setting
f0 C 14 f1 D 0; 1
4 f1 C 21 f2 D 0 and 21 f1 D 1.
Equation (3.124) reduces to
w x y
D cos m cos n 3.126
R R R
which corresponds to the well known checkerboard type buckling pattern obtained
at the bifurcation point of the classical linearized stability theory valid, strictly
speaking, for infinitesimal values of the wave amplitude. With other values of the
parameters f0 , f1 and f2 , wave patterns intermediate between these two limits can
be obtained. Notice that the wavelengths in the axial and circumferential direction
are 2
R/m and 2
R/n, respectively.
It is important to remember that with the appearance of high-speed photography
it was possible to show that indeed this latter incipient buckling mode plays an
important role at the beginning of the buckling process ([9.73], Volume 2).
Von Kármán and Tsien defined failure as the transition from the stable prebuck-
ling to the stable postbuckling configuration. Such a transition would occur at the
lowest bifurcation point along the prebuckling path or earlier, if through external
disturbances enough energy was imparted to the system to overcome the “energy
barrier” represented by the unstable portion of the postbuckling path. Since it is
difficult to determine the magnitudes of these disturbances in advance, it has been
suggested to use the minimum of the postbuckling curve as a safe design value in
practical engineering, thus lending theoretical support to the Lower Bound Design
Philosophy.
Direct Solutions of the Nonlinear Stability Problem 169

Von Kármán and Tsien’s work [3.1] on the postbuckling behavior of perfect
cylindrical shells has been followed by refinements and extensions by many inves-
tigators ([3.29] [3.32] and [9.189], Volume 2). In general all investigators used
the same method for finding equilibrium configurations in the postbuckling range.
That is, initially the potential energy of the system was expressed in terms of
finite displacements and then the equations governing the equilibrium configura-
tions were found by the application of the principle of stationary potential energy. It
was felt that by using a sufficiently refined expression for the out-of-plane displace-
ment w, the finite displacement analysis would yield an accurate prediction of the
postbuckling minimum load LB (see Figure 3.2). Thus using
1 1
w 
x
y
D ajk cos j cos k 3.127
R j,kD0
x y

where
L
R
x D , y D and j C k D even
m n
terms have been added successively to the displacement function until no significant
change occurred in the magnitude of the minimum postbuckling load. The system
of simultaneous nonlinear equations obtained by minimizing the total potential
energy with respect to the generalized coordinates x , y and the aij of the shell
were solved by the Newton Raphson iterative method. In Figure 3.24 the result
of Kempner [3.31], who kept the three coefficients a20 , a11 and a02 , is shown as
case I with LB D 0.301. Almroth’s result [3.32], who kept the nine coefficients
a20 , a11 , a02 , a40 , a31 , a22 , a13 , a60 and a33 , is shown as case II with LB D 0.108.
In a 1966 paper Hoff et al. [3.33] have suggested that if the number of terms in
the expression for the radial displacement w (Eq. (3.127)) approaches infinity, then
the minimum of the postbuckling equilibrium curve approaches zero, whereas at
the same time also h/R ! 0. This mathematical limit is naturally of little use in
engineering applications.

Figure 3.24 The search for the postbuckling minimum


170 Postbuckling Behavior of Structures

Figure 3.25 Experimental postbuckling curves (from [34]) (R/h D 394, L/R D 2.3)

Experimental results published by Esslinger and Geier [3.34] have shown conclu-
sively that axially compressed, thin-walled, finite length cylinders possess a low
but nonzero minimum postbuckling load-carrying capacity (see Figure 3.25). It has
been argued to use this value as a possible design load on the grounds that the
cylinder would always support at least this much load, and that even the presence of
initial imperfections would not reduce the critical buckling load below this value.
However, because of the often very low postbuckling minima, for weight sensitive
applications this approach will definitively result in technically unacceptable solu-
tions. Also for other applications the practical use of this idea is questionable, since
up to now there are no universally accepted postbuckling lower-bound buckling
loads for axially compressed cylindrical shells. Nevertheless, in the past few years
there have been new attempts, [3.35] and [3.36], to revitalize the lower-bound
design approach as a possible value for residual strength after damage. However,
it remains to be seen whether significant advances will be achieved.

b. Imperfect Shells

In order to account for the effect of small, stress free deviations of the shell mid-
surface from the circular cylindrical shape, for orthotropic shells under combined
axial compression, lateral pressure and torsion one must solve the following
Donnell Mushtari Vlasov type imperfect shell equations
1 1
LH f  LQ w D w,xx  LNL w, w C 2w
N 3.128a
R 2
1
LQ f  LD w D  f,xx C LNL f, w C w
N Cp 3.128b
R
where both the initial geometric imperfection wN and the out-of-plane displacement
w are taken to be positive outward. The linear and nonlinear operators have been
defined earlier (see Eqs. 3.49 3.50).
Direct Solutions of the Nonlinear Stability Problem 171

If we assume that the initial radial imperfection is given by


x x n
wN D hN1 cos i
C hN2 sin m
cos y  K x 3.129
L L R
then any equilibrium state of the circular cylindrical shell under combined axial
compression , lateral pressure pN and torsion N can be represented by
x x n
w D hW C Wp C Wt  C h1 cos i
C h2 sin m
cos y  K x 2.130a
L L R
 
Eh 1 2 1 O
fD  yO  pN e x 2  N xy C f 3.130b
cR 2 2
where the quantities W , Wp and Wt are (as mentioned earlier) evaluated by
enforcing the periodicity condition and

O D   RO p.
N 3.131

The prescribed value of RO depends on the loading case considered.


Notice that because of the sign convention used (w and wN are positive outward)
for internal pressure pN D pN i , whereas for external pressure pN D pN e . Recall further
that  is the nondimensional axial load parameter  D cR/Eh2 N0 , pN e is the
nondimensional external pressure pN e D cR2 /Eh2 pe , pN i is the nondimensional
internal pressure pN i D cR2 /Eh2 pi  and N is the nondimensional torque parameter
N D cR/Eh2 Nxy , positive counter-clockwise. Thus it can easily be shown that
for an axially compressed pressurized cylindrical shell the nondimensional axial
load parameter is
O D   12 pN i 3.132

In the absence of torsional loading (if N D 0) Khot’s skewedness parameter K , a


real number which was introduced to denote the inclination of the nodal lines of
the buckling pattern with respect to the axis of the shell [3.13], is identically equal
to zero.
An approximate solution of the nonlinear governing equations is obtained as
follows. First, the compatibility equation (3.128a) is solved exactly for the stress
function fO in terms of the assumed radial displacement w and the specified initial
N In this solution, only the effect of initial imperfections on the buck-
imperfection w.
ling load is of interest. Hence only a particular solution of Eq. (3.128a) needs to
be considered. Second, the equation of equilibrium, Eq. (3.128b) is solved approx-
imately by substituting therein for f, w and w, N and then applying Galerkin’s
procedure. This procedure yields the following set of nonlinear algebraic equations
in terms of the unknown amplitudes 1 and 2

ci  ^1 C D1 1 C D2 2 C D3 1 2 C D4 22 C D5 1 22 D ^N1 3.133


^c  ^2 C D6 2 C D7 1 C D8 12 C D9 1 2 C D10 22 C D11 12 2 C D12 23
D ^ C K^ N2 3.134
172 Postbuckling Behavior of Structures

where K^ is defined below. The coefficients D1 through D12 are listed in [3.37].
These equations describe the prebuckling, buckling and postbuckling behavior
of perfect N 1 D N2 D 0 and imperfect shells N 1 6D 0, N2 6D 0 under combined
loading. Solutions are obtained by Riks’ path following technique [3.38] whereby
one of the three load parameters O or pN e or N  is selected as the variable load ^.
The remaining two load parameters are assigned fixed values. Further, ci is the
axisymmetric buckling load

2N
1 1 C ˛ Q  2
ci D N xx C
˛2i D i xx
3.135
2 N xx
˛2i H

and ^c is the asymmetric buckling load for the specified values of m, n and K .
For ^ D , both pN c and N fixed
 
1 1 TN 3,m,n
2 TN 4,p,n
2
^c D 2 TN 1,m,n C TN 2,p,n C C  2pN e ˇn2
˛m C ˛2p 2 TN 5,m,n TN 6,p,n

 2N ˛m  ˛p ˇn 3.136

1
K^ D 2 N
p e ˇ 2
C 2N
 ˛m  ˛p ˇn 3.137
˛2m C ˛2p n

whereas for ^ D pN e , both O and O fixed


 
1 1 TN 3,m,n
2 TN 4,p,n
2
^c D 2 TN 1,m,n C TN 2,p,n C C O 2m C ˛2p 
 ˛
2ˇn 2 NT5,m,n TN 6,p,n

 2N ˛m  ˛p ˇn 3.138

1 2
K^ D O m C ˛2p  C 2N ˛m  ˛p ˇn
˛ 3.139
2ˇn2

and for ^ D N , both O and O fixed


 
1 1 TN 3,m,n
2 TN 4,p,n
2
^c D TN 1,m,n C TN 2,p,n C C
2˛m  ˛p ˇn 2 TN 5,m,n TN 6,p,n

O m C ˛p   2pN e ˇn
 ˛ 2 2 2
3.140

1
K^ D O 2m C ˛2p  C 2pN e ˇn2 g.
f˛ 3.141
2˛m  ˛p ˇn
The coefficients used are listed in [3.14].
Direct Solutions of the Nonlinear Stability Problem 173

The 2-modes solution was first employed by Hutchinson in 1965 for isotropic
shells [3.39] and was extended by Arbocz in 1973 to orthotropic shells [3.10]. It
is available as one of the computational modules in DISDECO [3.18]. Hutchinson
has restricted his imperfections to the form of the classic buckling modes. In the
present analysis the imperfections are quite general. However, as has been shown
in [3.39] and [3.10] in order to activate the nonlinear interaction between the
axisymmetric and the asymmetric modes the condition i D 2m must be satisfied.
The form of the 2-modes simplified imperfection model of Eq. (3.129) is dictated
by the results of a 1976 paper by Arbocz and Sechler [3.40], in which the effect
of different boundary conditions was investigated.
Figure 3.26 displays traces of the response curve in the load () vs amplitude
of the asymmetric displacement (2 ) plane for the axially compressed stringer
stiffened shell AS-2 of [1.25]. Notice that initially both the perfect shell and the
shells with axisymmetric imperfection have zero asymmetric deflection until the
bifurcation point is reached. Following bifurcation the axial load initially decreases
with increasing asymmetric deflection 2 .
On the other hand, as can be seen from Figures 3.27 and 3.28, the buckling
behavior of a shell with asymmetric imperfection only or with both axisym-
metric and asymmetric imperfections is characterized by the occurrence of a limit
point.

Figure 3.26 Theoretical Postbuckling curves for axisymmetric imperfections


174 Postbuckling Behavior of Structures

Figure 3.27 Theoretical Postbuckling curves for asymmetric imperfections

Figure 3.28 Theoretical Postbuckling curves for both axisymmetric asymmetric imperfections
Direct Solutions of the Nonlinear Stability Problem 175

The comparison of the predicted imperfection sensitivity by the asymptotic


Koiter type analysis with the results obtained by the 2-modes solution of the
nonlinear governing equations reveals excellent agreement for axisymmetric imper-
fections only (see Figure 3.29), and good agreement for asymmetric imperfections
only (see Figure 3.30). The explanation for the increasing differences between
the two approaches in the case of increasing asymmetric imperfections shown in
Figure 3.30 is given by the fact that the perturbation approach used for the b-factor
N whereas the 2-modes solution keeps terms up
method neglects terms of order ( )
to and including order ( N 2 ).
Finally in Figure 3.31 the effect of both axisymmetric and asymmetric imper-
fections on the buckling load of unpressurized cylinders is displayed.

3.3.5 Concluding Remarks


The reader may have asked himself how it is that the authors of a book on buckling
experiments spend so much space on covering the stability theory of thin-walled

Figure 3.29 Critical bifurcation load vs axisymmetric imperfection amplitude (curves obtained
by “axibif” and “twomod” are practically identical)
176 Postbuckling Behavior of Structures

Figure 3.30 Critical limit load vs asymmetric imperfection amplitude (for N D 0)

structures. This has to do with the conviction of the authors that theory and
experiments must go hand-in-hand if the sometimes very complicated structural
stability problems of everyday practice are to be solved successfully. Before doing
experiments it is especially important to carry out the initial stability analysis of
the perfect structure with great care and accuracy. In addition, if one finally has
found the lowest (critical) buckling load and the corresponding buckling mode (or
modes) it is necessary to investigate whether it has a stable or unstable postbuck-
ling behavior. With the advanced computational software and hardware currently
available a final accurate check using a full nonlinear calculation based on the
measured initial imperfections (see Chapter 10, Volume 2 for further details) and
employing careful modeling of the experimental boundary conditions (see also
Chapter 11, Volume 2) must become a standard practice. Only then can one assert
with reasonable accuracy what will be the expected behavior of the real (hence
imperfect) structure under the sometimes manifold loading conditions it may be
exposed to during its functional life time.
References 177

Figure 3.31 Critical limit load vs asymmetric imperfection amplitude (for N D


6 0)

References

3.1 von Kármán, Th. and Tsien, H.S., The Buckling of Thin Cylindrical Shells Under
Axial Compression, Journal of the Aeronautical Sciences, 8, 1941, 303 312.
3.2 Donnell, L.H. and Wan, C.C., Effect of Imperfections on Buckling of Thin Cylinders
and Columns under Axial Compression, ASME Journal of Applied Mechanics, 17,
(1), March 1950, 73 83.
3.3 Cohen, G.A., Effect of a Nonlinear Prebuckling State on the Postbuckling Behavior
and Imperfection Sensitivity of Elastic Structures, AIAA Journal, 6, (8), August 1968,
1616 1619.
3.4 Fitch, J.R., The Buckling and Post-Buckling Behavior of Spherical Caps under
Concentrated Load, International Journal of Solids and Structures, 4, 1968, 421 446.
3.5 Budiansky, B. and Hutchinson, J.W., Dynamic Buckling of Imperfection Sensitive
Structures, in: Proceedings 11th IUTAM Congress in Munich, 1964, H. Görtler ed.,
Springer Verlag, Berlin-Heidelberg-New York, 1964, 636 651.
178 Postbuckling Behavior of Structures

3.6 Arbocz, J. and Hol, J.M.A.M., ANILISA Computational Module for Koiter’s Imper-
fection Sensitivity Theory, Report LR-582, Delft University of Technology, Faculty
of Aerospace Engineering, Delft, The Netherlands, January 1989.
3.7 Hutchinson, J.W. and Amazigo, J.C., Imperfection-Sensitivity of Eccentrically Stiff-
ened Cylindrical Shells, AIAA Journal, 5, (3), March 1967, 392 401.
3.8 Geier, B., Das Beulverhalten versteifter Zylinderschalen. Teil 1: Differentialgle-
ichungen, Zeitschrift für Flugwissenschaften, 14, July 1966, 306 323.
3.9 Singer, J., Personal Communication, 1969.
3.10 Arbocz, J., The Effect of Initial Imperfections on Shell Stability, in: Thin-Shell Struc-
tures, Theory, Experiment and Design Y.C. Fung and E.E. Sechler eds., Prentice
Hall, Englewood Cliffs, N.J., 1974, 205 245.
3.11 Seggelke, P. and Geier, B., Das Beulverhalten versteifter Zylinderschalen. Teil 2:
Beullasten, Zeitschrift für Flugwissenschaften, 15, December 1967, 477 490.
3.12 Block, D.L., Card, M.F. and Mikulas, M.M. Jr., Buckling of Eccentrically Stiffened
Orthotropic Cylinders, NASA TN D-2960, 1965.
3.13 Khot, N.S. and Venkayya, V.B., Effect of Fiber Orientation on Initial Postbuckling
Behavior and Imperfection Sensitivity of Composite Shells, Report AFFDL-TR-70-
125, Air Force Flight Dynamics Laboratory, Wright-Patterson Air Force Base, Ohio.
3.14 Arbocz, J., The Effect of Initial Imperfections on Shell Stability An Updated Review,
Report LR-695, Delft University of Technology, Faculty of Aerospace Engineering,
Delft, The Netherlands, September 1992.
3.15 Koiter, W.T., Elastic Stability and Postbuckling Behavior, Proceedings Symposium
on Nonlinear Problems, University of Wisconsin Press, Madison, 1963, 257 275.
3.16 Hutchinson, J.W. and Koiter, W.T., Postbuckling Theory, Applied Mech. Rev., 23,
1970, 1353 1366.
3.17 Arbocz, J., Comparison of Level-1 and Level-2 Buckling and Postbuckling Solutions,
Report LR-700, Delft University of Technology, Faculty of Aerospace Engineering,
Delft, The Netherlands, November 1992.
3.18 Arbocz, J. and Hol, J.M.A.M., Shell Stability Analysis in a Computer Aided
Engineering (CAE) Environment, in: Proceedings 34th AIAA/ASME/ASCE/AHS/ASC
Structures, Structural Dynamics and Materials Conference, April 19 22, La Jolla,
California, 1993, 300 314.
3.19 Cohen, G.A., User Document for Computer Programs for Ring-Stiffened Shells of
Revolution, NASA CR-2086, March 1973.
3.20 Cohen, G.A., FASOR A Second Generation Shell of Revolution Code, Computers
& Structures, 10, 1979, 301 309.
3.21 Roorda, J., Stability of Structures with Small Imperfections, Journal Eng. Mech.
Div., ASCE, 91, (EM1), 1965, 87 106.
3.22 Kempner, J. and Chen, Y.N., Buckling and Postbuckling of an Axially Compressed
Oval Cylindrical Shell, Proceedings 70th Anniversary Symposium on the Theory of
Shells to Honor Lloyd Hamilton Donnell, University of Houston, Houston, Texas,
1967, 141 183.
3.23 Almroth, B.O., Brogan, F.A. and Marlowe, M.B., Collapse Analysis for Elliptic
Cones, AIAA Journal, 9, (1), January 1971, 32 37.
3.24 van der Neut, A., Postbuckling Behavior of Structures, NATO AGARD Report 60,
1956.
3.25 Koiter, W.T., The Effective Width at Loads far in Excess of the Critical Load for
Various Boundary Conditions, (in Dutch), NLL Report S287, Amsterdam, 1943.
References 179

3.26 Cox, H.L., The Buckling of a Flat Plate under Axial Compression and its Behavior
after Buckling, Aeronautical Research Council, R. & M. 20201, 1945.
3.27 Vilnay, O. and Rodney, K.C., A Generalized Effective Width Method for Plates
Loaded in Compression, Journal of Constructional Steel Research, 1, (3), May 1981,
3 12.
3.28 Levy, S., Bending of Rectangular Plates with Large Deflections, NACA TN 846,
May 1942. (Also available as NACA TR 737, 1942.)
3.29 Legget, D.M.A. and Jones, R.P.N., The Behavior of a Cylindrical Shell under Axial
Compression when the Buckling Load has been Exceeded, ARC R. & M. 2190, 1942.
3.30 Michielsen, H.F., The Behavior of Thin Cylindrical Shells after Buckling under
Axial Compression, Journal Aeronautical Sciences, 15, 1948, 738 744.
3.31 Kempner, J., Postbuckling Behavior of Axially Compressed Cylindrical Shells,
Journal Aeronautical Sciences, 21, 1954, 329 342.
3.32 Almroth, B.O., Postbuckling Behavior of Axially Compressed Circular Cylinders,
AIAA Journal, 1, (3), March 1963, 630 633.
3.33 Hoff, N.J., Madsen, W.A. and Mayers, J., Postbuckling Equilibrium of Axially
Compressed Circular Cylindrical Shells, AIAA Journal, 4, (1), January 1966,
126 133.
3.34 Esslinger, M. and Geier, B., Buckling and Postbuckling Behavior of Thin-Walled
Circular Cylinders, Deutsche Luft- und Raumfahrt FB 69 99, 1966.
3.35 Croll, J.G.A., Towards Simple Estimates of Shell Buckling Loads, Der Stahlbau, 44,
1975, 243 248 and 283 285.
3.36 Wittek, U. and Krätzig, W.B., Ein Masstab für die Beurteilung der Imperfektions-
unempfindlichkeit allgemeiner Schalen, Schalenbeultagung Meersburg, 1976, Sond-
erheft der DFVLR, 1976, 170 182.
3.37 Arbocz, J., Potier-Ferry, M., Singer, J. and Tvergaard, V., Buckling and Post-
Buckling, Lecture Notes in Physics No. 288, Springer-Verlag, Berlin, Heidelberg,
1987, 83 142.
3.38 Riks, E., The Application of Newton’s Method to the Problem of Elastic Stability,
ASME Journal of Applied Mechanics, 39, 1972, 1060 1066.
3.39 Hutchinson, J.W., Axial Buckling of Pressurized Imperfect Cylindrical Shells, AIAA
Journal, 3, (8), August 1965, 1461 1466.
3.40 Arbocz, J. and Sechler, E.E., On the Buckling of Stiffened Imperfect Shells, AIAA
Journal, 14, (11), November 1976, 1611 1617.
4
Elements of A Simple
Buckling Test A Column
under Axial Compression

4.1 Columns and Imperfections


In order to appreciate the nature of the experimental approach, let us first consider
a simple buckling test a column under axial compression. The column is not
only the earliest and classic example of elastic instability and postbuckling studies,
dating back to Euler’s work in 1744 ([2.9] and [4.1]), but is also the element that
has been the subject of the most extensive experimental and theoretical studies
since the experimental investigations of Petrus van Musschenbroek in 1729 [4.2].
Musschenbroek had discovered by experiment that the buckling load was inversely
proportional to the square of the length of the column, a result later deduced
theoretically by Euler. Salmon’s 1920 treatise [4.3] summarizes most of the column
studies till 1920, lists nearly 400 references, and shows how experiment and theory
advanced hand in hand.
If one studies the experimental work on buckling of columns carried out at the
turn of the century, one is amazed at its high quality. Maybe one should not be
surprised, since it was performed by some of the giants of mechanics, like von
Kármán ([4.4] and [2.76]) and Prandtl [4.5], who in their doctoral dissertations
combined outstanding analysis with outstanding experiments.
It is certainly illuminating to review such a classic set of tests, for example von
Kármán’s 1907 10 experiments [4.4]. In the introduction, the main purpose of
the work is stated to be the experimental proof of the formulae for the buckling
strength of shorter columns, which buckle inelastically. Von Kármán points out,
however, that the experiments presented an opportunity to investigate the influence
of exact centering of the columns, as well as the postbuckling behavior after the
peak load has been exceeded.

Buckling Experiments: Experimental Methods in Buckling of Thin-Walled Structures: Basic Concepts, Columns, Beams and
Plates – Volume 1. J. Singer, J. Arbocz and T. Weller Copyright © 1998 John Wiley & Sons, Inc.
182 Elements of A Simple Buckling Test A Column under Axial Compression

One can generalize von Kármán’s comments on the influence of centering of


the columns, to the influence of imperfections in columns, of which eccentricity of
loading is only one. These imperfections can be divided according to their effect
into three groups, more or less along the lines suggested by Salmon [4.3]:

1. Eccentricity of Loading: Eccentricity of the load.


Variations in the modulus of elasticity.
Inequality of areas and shape of
cross-sections.
Nonhomogeneity of material.
2. Initial Curvature: Initial curvature.
Variations in the modulus of elasticity.
3. Reduction in Strength of Material: Residual stresses.
Nonhomogeneity of material.
Flaws and local defects, like voids and
delaminations in composites.

Eccentricity of loading is the imperfection that attracted most of the attention


of the investigators in the 19th century and in the beginning of the 20th. Both
methods of estimation and determination from experiment were extensively tried,
and besides the non-central application of the load, the variations in modulus
were considered to be of primary importance. Based on their experiments and the
accumulated experience, many investigators proposed likely values for eccentricity
of loading ε2 D L/70 to L/150, where L is the length of the column.
The influence of initial curvature was seriously considered only at the turn of
the century. To assess this influence, Salmon collected experimentally observed
initial curvatures and deduced from his nearly randomly chosen points (Figure 42
of [4.3]) an empirical probable initial deflection ε1 D L/750. The effects of initial
curvature are similar to those of eccentricity of loading and hence the two effects
can be combined into an equivalent initial deflection, ε D ε1 C ε2 .
The third imperfection, reduction in strength of the material, was of concern
already to investigators at the beginning of this century, but the great influence of
residual stresses on the strength of columns and plates in compression has only
been elucidated in recent decades (see [4.6] or [4.7]). The effects of flaws and local
defects, and in particular those of voids and delaminations in composites, are still
the subject of intensive studies.

4.2 Von Kármán’s Experiments


Returning now to the von Kármán experiments, he emphasized the importance of
load eccentricity, by comparing the load-deflection curves of his rather precisely
centered columns with those of Tetmajer [4.8] and those of Kirsch [4.9] as can be
seen in Figure 4.1 (reproduced from [4.4]). It can be clearly seen in the figures
that the very careful centering of von Kármán’s columns, yielded load-deflection
Von Kármán’s Experiments 183

Figure 4.1 Load-deflection curves of von Kármán’s columns compared with those of Tetmajer
and of Kirsch (from [4.4])

curves that closely approximate those of an ideally perfect column; whereas the
load-deflection curve of a typical Kirsch column, with noticeable load eccentricity,
differs significantly in behavior, and even the careful tests of Tetmajer exhibit some
eccentricity of loading effects, appearing as deflections already at the lower loads.
The figure also indicates why columns with appreciable load-eccentricity will yield
collapse loads that are significantly below the Euler load the buckling load of an
ideal centrally loaded column.
Von Kármán’s tests at the University of Göttingen were carried out in a 150-
ton hydraulic compression testing machine with a 1000 mm long working section,
permitting slender specimens with a convenient cross section of the order of 20 ð
30 mm2 . In the tests, the testing machine was not loaded beyond 20 percent of its
capacity. This ensured a relative rigidity of the test setup, and eliminated possible
eccentricity of loading resulting from bending of the two pillars of the machine.
Careful compression tests on very short specimens to evaluate the compressive
material properties, preceded the buckling tests. This was not the usual procedure at
the time and only became standard, as a “stub-column test procedure”, in the sixties.
Von Kármán designed special end fixtures (Figure 4.2) which facilitated accurate
placing of the centerline of the column on the loading line. This fixture permitted
readjusting of the position of the column under load, that led to more accurate
centering. In describing these very neat fixtures, he gave credit to Considère [2.74]
for first using readjustable end fixtures and to Prandtl [4.5] for the achievement of
a “near theoretical” behavior, of negligible deflection before the buckling load is
approached, by extremely careful centering of his test columns.
Von Kármán pointed out two sources of error connected with the attachment of
the test specimens. The first error is inclusion of the practically rigid end fixtures
184 Elements of A Simple Buckling Test A Column under Axial Compression

Figure 4.2 Von Kármáns end fixtures for column tests (from
[4.4])

in the effective length of the tested columns. This error may be assessed by an
approximate evaluation of the equivalent length of an entirely elastic column. Von
Kármán calculated the necessary length correction and showed it to be small, less
than 0.3 percent even for rigid ends extending over 10 percent of the total length.
The calculation was for elastic buckling. For inelastic buckling the correction is
even smaller, since then the deflection curve deviates from the elastic sine curve
by limiting its curvature to the middle position of the columns and the ends remain
nearly straight, coinciding with the tangential rigid ends.
The second error can be due to the friction of the knife edges in their bases,
which would increase the measured buckling load and perhaps make them exceed
the theoretical Euler load, which is an upper bound. From his results in the elastic
range, which generally deviated from the Euler load by less than 1 1.5 percent,
von Kármán concluded that this error also was not significant.
These apparently minor details are pointed out here, since they signify some
elements of the methodology of careful buckling experiments: readjustment of
specimen positions in end fixtures under load, assessment and compensation for end
fixture rigidity and consideration of secondary effects like friction and justification
of their neglect by comparison with well established earlier results.
Von Kármán continued many of his tests well into the postbuckling region (see
Figure 4.3), since he realized that the postbuckling behavior is of importance for the
understanding of the buckling behavior. The results (see Figure 4.3) show that for
the long columns, the theoretically predicted elastic behavior of constant load with
The Basic Elements of a Buckling Experiment 185

Figure 4.3 Postbuckling load-deflection curves for von Kármán’s long and short columns
(from [4.4])

increasing deflection is indeed confirmed experimentally (for example in column


no. 1). For short columns, however, the behavior differs and the load decreases
very significantly with increasing deflection, due to inelastic effects (for example
in column no. 6). These inelastic effects were the prime interest in von Kármán’s
thesis, and indeed have since been the main topic of experimental studies of the
buckling of columns.

4.3 The Basic Elements of a Buckling Experiment


In this chapter a simple buckling test was considered of a centrally compressed
column. One should remember, that a simple column, a perfectly straight, uniform
column, loaded centrally through frictionless pin-ends, is an idealized compression
member, and though it is the basic theoretical element for the study of buckling
and postbuckling of structures, it is not found in real structures! As has been stated
by many investigators (for example, [2.7], [4.7], [4.10] or [4.11]), the strength of
practical columns depends on the initial geometric imperfections (usually called
initial out-of-straightness in columns), eccentricities of load, transverse loads, the
boundary conditions, local buckling (if the column is thin-walled or built-up), the
homogeneity of the material and residual stresses. Many tests on columns did not
isolate these various effects and hence fairly wide scatter bands resulted for column
curves. The modern test procedures and a discussion of these effects are presented
in Chapter 6.
Having studied the classic column tests of von Kármán and briefly looked at the
different imperfections that affect a column test, one can discuss and summarize
the basic elements of a buckling test. The first question one has to address is “what
is the aim of the buckling experiment?” Is the aim to explore the physical behavior
186 Elements of A Simple Buckling Test A Column under Axial Compression

near, at and after buckling, or is the aim only verification of the theory derived for
a perfect structure, a perfect column in our simple buckling test? Is it verification
of an exact theory for the perfect structure or an approximate one, that predicts the
behavior of a simplified model; or is the aim to verify a theory for an “imperfect”
structure, closer to the real one?
One has to remember that, as was discussed in Chapter 2, classical
buckling bifurcation buckling occurs only in ideal perfect structures. Real
imperfect structures begin to deform from the initiation of loading, and buckling in
an engineering sense occurs when the lateral (“escaping”) deformations grow at an
increasing rate. This difference between the theoretical bifurcation behavior of an
ideal column and the behavior of real columns is very clearly shown in Figure 4.1,
especially for the Kirsch column. But even von Kármán’s carefully made, centered
and tested columns show to some extent the behavior of real columns, though they
approach that of the theoretical model very closely.
Hence, if the aim of the experiments is to verify the theory for a perfect struc-
ture, methods are needed that correlate the test results on real structures with
the predictions for an idealized structure. The most widely used method is that
proposed by R.V. Southwell in 1932 for simply supported columns [4.12], often
called the Southwell Plot. In this method, the deflection w is plotted versus w/P
and the slope of the resulting straight line yields the buckling load of the corre-
sponding perfect column. Details of the Southwell method, its limitations and its
extensions will be discussed later. One should remember that the main value of
methods like the Southwell plot is to provide a correlation between experiments on
real imperfect structures and theoretical predictions, i.e. to facilitate the verification
of theory.
But experiments aim not only to verify theories, they explore the physical
behavior near buckling, at buckling and in the postbuckling range and they also
yield empirical data upon which design guidelines can be based.
The design of a proper experiment is determined by its main purpose. If it aims
at verification of theory, the experiment should be carried out under as “perfect”
conditions as possible, with specimens made as accurately and measured as care-
fully as possible, and from a material whose composition can be conveniently
controlled and measured, and with boundary conditions that can be determined
as accurately as possible and simulated adequately in the theory. If the physical
behavior is to be explored, and primarily in the postbuckling range, specimens
made of materials that behave elastically much beyond the buckling loads (for
example, polyesters like Mylar, Melinex or Diafoil), may be preferable though
their behavior significantly differs from that of structural materials used in prac-
tice. On the other hand, if the data obtained in the experiment is to be employed
for design guidelines, the specimens should simulate the real structure, as well
as boundary conditions and environment. The effects of scaling have to be well
understood in all the cases to ensure correct interpretation of the results.
Hence, how does one plan a simple buckling experiment, say, one for simple
columns, to be used for verification of the inelastic column theory? This is exactly
the experimental task faced by von Kármán in his classical 1907 10 tests, discussed
Demonstration Experiments 187

earlier in this chapter and can be used as an example. Von Kármán indicated how he
planned his experiments. He indeed carried out his experiments under as “perfect”
conditions as possible. The steel specimens were machined and measured as accu-
rately as the then prevalent techniques permitted, the chemical composition of
the steel was ascertained and the measurements of mechanical properties were
extended to include compression tests on very short specimens cut from the same
bars as the buckling specimens. The specimens varied from long ones, serving as
reference tests for elastic buckling, to shorter ones in the range relevant to the
inelastic theory to be verified. Boundary conditions were controlled by the special
end fixtures (Figure 4.2) permitting centering under load, and the relative rigidity
of the test setup was assured by designing the specimens to require always less
than 20 percent of the load capacity of the testing machine. Measurement instru-
ments were 1910 state of the art, but their accuracy was invoked only where the
experimenter felt the measurement was essential (though he checked his predicted
“unimportant” displacements). The experiments were designed for verification of
a theory, though as was mentioned earlier, von Kármán intentionally extended
his aim beyond that, to better understanding of the influence of eccentricity and
the postbuckling behavior of his columns. The effects of residual stresses, were,
however, not considered; they dominate modern column analysis and experiments,
and will be discussed in Chapter 6. There also the column curves which guide the
designer are discussed and evaluated.

4.4 Demonstration Experiments


There are other types of experiments which can be called demonstration experi-
ments. These are experiments specially designed to bring out certain phenomena
or effects, this being achieved by exaggeration of certain properties or geometries,
much beyond their magnitude in real structures, or by replacing the structure by
a mechanical model which simulates the essential behavior of the structure under
load. Such experiments have been employed extensively for the study of equi-
librium paths and initial postbuckling behavior of imperfect structures, primarily
frames, trusses and arches, as well as for teaching demonstrations and for study of
postbuckling behavior of shells, which will be discussed in Chapter 9, Volume 2.

4.4.1 University College London Initial Postbuckling Experiments

In the sixties a series of experiments were performed, by Roorda, Brivtec and


Chilver at the Department of Civil and Municipal Engineering, University College
London, to verify theories developed there for the initial postbuckling behavior
of imperfect structures, which are essentially of the demonstration type ([3.21],
[4.13] [4.15]). As an example, one of these experiments from Roorda’s thesis a
simple strut loaded eccentrically at one end ([4.14] or [3.21]) is described in
detail. Figure 4.4a shows the loading arrangement schematically. The strut is a
188 Elements of A Simple Buckling Test A Column under Axial Compression

(a) (b)

Figure 4.4 Roorda’s experiment on a simple strut loaded eccentrically at one end (from [4.14]):
(a) loading setup schematic, (b) screw arrangement for offset loading

high strength steel strip of a nominal rectangular 1 inch ð 1/16 inch cross section
and length L D 23 inches. The point of application of the load at the top of the strut
has a small variable eccentricity d, introduced with a screw arrangement shown
in Figure 4.4b (taken from a second similar experiment in [4.14], on a strut with
offset loading at both ends). The lower end here remains in a fixed position and is
simply a knife-edge filed on the end of the strut resting in a V-groove. The load
W on the strut, consists of lead shot 2W applied through a rigid load beam. Due
to the eccentricity d, there is a small external moment Wd applied at the upper
end of the strut, in addition to the axial force W. In the limit, when d vanishes,
the strut is a centrally loaded column. The rotation  at the top end represents the
displacement parameter and is measured optically, as the load is gradually varied.
It was found that in order to obtain a centrally loaded column, a small negative
eccentricity had to be introduced to overcome the initial curvature of the strut and
come near a point of bifurcation. The natural experimental equilibrium path for
the centrally loaded column, I in Figure 4.5, was traced simply by adding lead
shot to the loading pan, the path being stable throughout. To obtain the comple-
mentary path, II in Figure 4.5, the direction of the buckling wave was changed
manually, while the load was kept constant at a point slightly above the “ideal”
critical load, W/Wcr  D 1 in Figure 4.5. The system remained in a stable position
on the complementary path. To obtain further points on this path the load was grad-
ually decreased, until the point of minimum load B was reached, whereupon the
strut jumped back to the stable position in the natural buckling shape direction I.
Interpolation of the two branches of the load W-rotation  plot yielded the ideal
experimental critical load Wcrexp at A (which was then used to non-dimensionalize
the plot in Figure 4.5).
The experimental critical load (WŁ ) versus eccentricity (d) behavior was obtained
in another experiment in which d was varied (see Figure 4.6). At each value of d,
the point of minimum load on the complementary equilibrium path (B in Figure 4.5)
was found by gradually decreasing the load along this path (as before in the case
of central loading), until snap-through to the opposite direction of buckle occurred.
Demonstration Experiments 189

Figure 4.5 Roorda’s strut experiment equilibrium paths (from [4.14] and [3.21])

The resulting plot is shown in Figure 4.6. The offset d0 /L of the experimental
curve from the load axis, is the amount of eccentricity necessary to just balance the
effect of other unknown imperfections, such as the initial curvature of the column,
as mentioned earlier for the case of the centrally loaded column. The photographs
in Figure 4.7, taken from another demonstration model, show how the column may
buckle.

4.4.2 Mechanical Models

As an example of a different type of demonstration experiment, one on the mechan-


ical models of Walker, Croll and Wilson [4.16] is discussed. These mechanical
models consist “essentially of two rigid links connected at a pin joint B (see
Figure 4.8a). The other ends of the links are also pinned, one being fixed spatially
at A and the other to C free to move in the longitudinal direction. Lateral loading
W at joint B may be applied independently of the longitudinal loading P at joint
C. Rigidity of the models is achieved by attaching various combinations of linear
coil springs to joints B and C and a torsion spring at joint B.” The frame which
190 Elements of A Simple Buckling Test A Column under Axial Compression

Figure 4.6 Roorda’s strut experiment experimental critical load versus eccentricity d (from
[4.14] and [3.21])

houses the models is shown in Figure 4.9, with a typical two-link model (3) in
position. It consists of a base (1) with two long outrigger arms (2) used for lateral
loading and restraint, their length ensuring practically normal lateral loading and
restraint over the expected deformation range of the model. The moving end of the
model (3) is mounted on the frame by means of a cross-member (4), connected
through linear bearings in housings (5) to run on rigidly aligned parallel hardened
circular surfaces (6). This assures a precise axial movement. Sprung buffers (7)
are clamped on these runners (6) as a safety measure to prevent damage to the
bearings of the model when violent dynamic action occurs. The pinned joints of
the model are, of course, provided with ball bearings (8); the central joint has two
ball bearings and a trunnion (9), to which the loading wires may be connected.
Longitudinal loading P is provided by wires (10) connected to the linear bearing
housings at one end, passing around ball bearing pulleys (11) and pinned to a distri-
bution bar (12) at the other end. A loading pan is hooked on to this distribution bar.
Longitudinal restraint when required is provided by springs (13), or wires, which
Demonstration Experiments 191

Figure 4.7 Unbuckled and stable, buckled forms of a strut in another of Roorda’s demonstra-
tion experiments (from [4.14])

Figure 4.8 Mechanical demonstration model for bifurcation behavior (from [4.16]): (a) sche-
matic presentation of model, (b) the model for stable symmetric bifurcation

are hooked to a ring on the underside of the cross-member (4). A screw arrange-
ment (14) is provided to accommodate varying lengths of springs, or to control
the displacement of the cross-member. The rods on the adjustment screws (15) all
have flats milled on one side to eliminate rotation when an adjustment is made,
and thrust ball races (16) are provided to further ease these adjustments.
Lateral loading W is applied by connecting a wire to the central trunnion (9),
passing it over a ball bearing pulley and hooking on a loading pan. Lateral restraint,
192 Elements of A Simple Buckling Test A Column under Axial Compression

Figure 4.9 Mechanical model demonstration experiment for buckling behavior test setup
(from [4.16])

when needed, is provided by springs (18) attached to the other side of the trunnion
by means of a short wire. A second adjustment screw (19) is also provided to
accommodate varying lengths of spring and, in addition, to impose prescribed
geometrical imperfections into the model.
An optical method is used to measure the angular displacement of the links (3).
A narrow beam of light from a projector is passed through a collimating lens
and reflected from a mirror on to a large circular screen. The mirror (28), shown
in Figure 4.10 is arranged so that the reflecting surface is over the center of the
pivot (8).
The torsional spring, when required, as in the case of symmetric bifurcation to be
discussed (Figure 4.8b), is mounted on the model coaxially with the central bearing
as shown in Figure 4.10. This consists of a coil spring (31) clamped between two
arms (32), which when rotated relative to one another set up a torsional restraint
to the deformation.
Many details of the test setup have been described to indicate the design consid-
erations and various capabilities of such a demonstration test rig. Tests on one of the
models, Model III (of [4.16]) for stable symmetric bifurcation, are now discussed.
The model, outlined in Figure 4.8b, demonstrates bifurcation at a critical load, Pcr ,
from a primary path to a stable secondary path, a behavior which characterizes the
buckling of perfect slender columns and perfect thin plates subjected to in-plane
Demonstration Experiments 193

Figure 4.10 Mechanical model demonstration experiment for buckling behavior detail of
test setup for stable symmetric bifurcation (from [4.16])

Figure 4.11 Mechanical model for stable symmetric bifurcation the influence of small initial
imperfections 0 on the load versus angular displacement behavior (from [4.16])

edge loading. The load paths of the model, with increasing small initial deflec-
tion 0 (at zero load), are shown in Figure 4.11, and illustrate the influence of
small imperfections on its non-linear behavior. The initial imperfections 0 may be
obtained by attaching a small weight W which applies a constant lateral load, as
P increases. Alternatively, the torsion spring B can be adjusted to give the model
194 Elements of A Simple Buckling Test A Column under Axial Compression

a small geometric imperfection at zero load with no lateral load. The two types
of imperfections are analogous to small lateral loadings which may be present in
real columns or plates, and to out-of-straightness or out-of-flatness occurring in
practical columns and plates. Comparisons of the effects of these two types of
imperfections on the model (for example, Figure 16 of [4.16]) showed them to be
very similar, as is the case in columns and plates. Hence the model demonstrates
the behavior of imperfect columns and plates very well. Many other types of non-
linear stability behavior are clearly demonstrated by other models in [4.16], which
makes the test rig a very useful demonstration tool.

4.5 Southwell’s Method

4.5.1 Derivation of Southwell Plot for a Column

Southwell [4.12] searched for a method that would enable one to obtain the theo-
retical buckling stress of a perfect column from experiments on real imperfect
columns. He pointed out that the load-deflection curves (P versus w) may be
approximated by “rectangular hyperbolas, having as asymptotes the axis of zero
deflection and the horizontal line P D PE ”. This hyperbolic relationship had been
known many decades earlier, as for example in Ayrton and Perry’s papers [4.17],
but Southwell recognized that “by a suitable change of coordinates, any such hyper-
bola may be transformed into a straight line, of which the slope is the measure
of PE ”.
Let us briefly rederive the Southwell plot for a simply supported column,
following essentially Southwell’s derivation, except that we use the more
appropriate 4th order equilibrium equation, as used in the derivations for imperfect
columns in textbooks (like [2.4], pp. 64 68, [4.18], pp. 230 242 or [4.19],
pp. 12 15) or in recent discussions of the application of the method (for example
[4.20]). The equilibrium equation of an imperfect (initially crooked) column is

wiv C ˛2 w00 D ˛2 w000 (4.1)


where wx D the additional lateral deflection measured in tests
w0 x D the initial deflection (imperfection)
˛2 D P/EI (4.2)

and the boundary conditions for simple supports are

w0 D w00 0 D wL D w00 L D 0. 4.3

Representing the additional deflection by a Fourier series


1
  nx 
wD Wn sin 4.4
nD1
L
Southwell’s Method 195

and the initial deflection by a Fourier series


1
  nx 
w0 x D W0n sin 4.5
nD1
L

and substitution in Eq. (4.1) leads to


  1   1
n2 2 EI n2 PE
Wn D W0n 1 D W0n 1 4.6
PL 2 P

and hence to
   1 
1 
  nx 
n2 PE
wx D W 1 sin 4.7
 0n P L 
nD1

where PE D 2 EI/L 2 , the Euler Load. The maximum deflection W D wL/2 is
therefore
W D W1  W3 C W5  Ð Ð Ð 4.8
where Wn is given by Eq. (4.6). When the buckling load is approached, or as
Southwell noted “if P is a fairly considerable fraction of PE ”
  1
¾ W1 D W01 PE
WD 1 4.9
P
and the fundamental mode predominates. Hence as P ! PE , the imperfection
component that represents the buckling mode is the one that is primarily magnified.
We can therefore write
W0 D¾ W01 4.10
and the expression for W can be rearranged as
W
W D PE  W0 . 4.11
P
The inverse slope of the plot of W/P versus W, the Southwell plot (Figure 4.14a),
yields the buckling load of the corresponding perfect column.

4.5.2 Application to von Kármán’s Columns

Southwell applied his method to the columns tested by von Kármán [4.4] and
Figure 4.12 (from [4.12]) shows the Southwell plots for the eight slender columns
of the von Kármán tests, whose slenderness ratio (L/) (where  is the minimum
radius of gyration of the cross section) is greater than 90. Note that only at the
higher values of W (or υ in the figure) the relation is linear and the procedure
justified. As a matter of fact, Southwell rejected all points for P < 0.8Pc , on
grounds that when both load and deflections are small their ratio υ/ will not be
196 Elements of A Simple Buckling Test A Column under Axial Compression

Figure 4.12 Von Kármán’s data on compressed columns plotted in the linear form by South-
well (from [4.12])

determinable with any accuracy, and only then applied the method of least squares
to the remaining points. The results of Figure 4.12 were excellent in no case did
the critical load derived from the Southwell plot differ by more than 2.2 percent
from the classic Euler load.
For the “medium” L/ D 45  90 and “thick” L/ < 45 groups of von
Kármán’s struts the method failed (to predict the Euler load), since practically
all measured deflections were already in the inelastic range.
Southwell concluded that it appeared “that the method has given good results
in every case where these could be expected, but that only trial can show whether
in any instance sufficient observations can be taken of deflections which on the
one hand are large enough to give reasonable certainty of υ/P, and on the other
hand are not so large that the material has ceased to be elastic”. This is a valid
assessment of the method even today, if one expects the method to yield the elastic
buckling load of the corresponding perfect structure, the Euler load in the case of
the column. The Southwell method can, however, be extended to plastic buckling as
will be shown in Chapter 16, Volume 2, but then it predicts the plastic bifurcation
load of the corresponding perfect structure, the reduced modulus load according to
the Kármán Engesser theory in the case of a column, instead of the elastic one.
In passing, it is of interest that among the many papers reporting tests on columns
up to 1932, Southwell could find only two papers that recorded related data of load
and central deflection for centrally loaded struts, von Kármán’s 1910 paper [4.4]
and Robertson’s 1925 paper [4.21]. He recommended that future experimenters
should publish such complete tables, and indeed the usefulness of the Southwell
plot has since motivated more complete data recording. With modern data acqui-
sition systems, large amounts of simultaneous data of loads and displacements
Application of the Southwell Method to Columns, Beam Columns & Frames 197

are usually recorded in buckling experiments, but are not always reported. When
publishing his data, the experimenter should give some thought to the future inves-
tigators, who may want to use his data for comparison with new theories and
experiments. Some recent proposals for standardization of presentation of imper-
fection measurements are discussed in Chapter 10, Volume 2.

4.6 Application of the Southwell Method to Columns,


Beam Columns and Frames
Southwell’s method has been widely used. Already in 1932 it was applied success-
fully to experiments on the stability under shearing forces of a flat elastic strip
[4.22]. Then in the second half of the thirties the method was extended and applied
to other structures and theoretically justified in some cases, [4.23], [4.24] and
[4.26]. Southwell [4.12] stated that “the main interest of the method lies in its
generality”, a challenge that was taken up by stability researchers in the decades
that followed.
Fisher [4.23] extended Southwell’s method to the case of a spar under combined
axial and transverse loading, a typical loading for an aeroplane spar in test or
flight. The theory was broadened to the general beam column and verified by
good agreement with experiments on eccentrically and transversely loaded solid
rectangular spars. Ramberg, McPherson and Levy [4.24] applied the method to
experiments on axially loaded sheet-stringer panels, and obtained good results for
the stringers attached to sheets, irrespective of failure being in a twisting or bending
mode, or a twisting-bending mode. They could not apply the Southwell method to
the sheet between stringers, due to lack of the bending strain below buckling (and
not because the method was not applicable to plates, as will be discussed in detail
in Chapter 8).

4.6.1 Lundquist Plot

One of the difficulties in the application of the Southwell method is that it requires
the initial deflection reading to be taken at zero load, where deflection measure-
ments are often questionable. A zero-point “correction” may therefore be necessary.
One way to apply a zero-point correction was suggested by Southwell: replacing
of υ versus υ/P by υ  υ0  versus υ  υ0 /P, where υ0 represents the zero-point
correction, which is chosen from some trial values as the one giving the straightest
line in the upper portion of the plot.
Another method to reduce low-load irregularities was given by Lundquist’s
generalization of the Southwell plot [4.25], for the incremental deflections of a
simply supported column due to incremental loads above initial values Pi and υi .
Equation (4.11) then becomes
W  υi
W  υi   PE  Pi   a1 4.12
P  Pi
198 Elements of A Simple Buckling Test A Column under Axial Compression

where
W0
a1 D Ð 4.12A
1  Pi /PE 
Note that a1 approximates υi , and when Pi is very small a1 D¾ W0 . The “Lundquist
Plot” (Figure 4.14c) is therefore similar to the Southwell plot and consists of a
plot of W  υi /P  Pi , versus W  υi , whose inverse slope yields PE  Pi ,
and whose horizontal intercept is a1 .

4.6.2 Donnell’s Applications of the Southwell Plot

In 1938 Donnell [4.26] reviewed the applicability of the Southwell method and
extended it considerably, discussing the justification for various cases. First he
considered a hinged strut with continuous elastic support. From the vanishing of
the total energy change due to a virtual displacement dWn , he obtained
   
Wn n2 2 EI L2 ˇ Wn
PD C 2 2 D Pn 4.13
W0n C Wn L2 n  W0n C Wn

where Pn is the critical load for the nth type of displacement for the strut without
any initial curvature, and ˇ is the spring constant per unit length (with dimension
force divided by length squared), usually called the modulus of foundation. When
W0n D 0, Eq. (4.13) reduces to the classical formula for a simply supported column
on an elastic foundation originally derived by Engesser in 1884 ([4.27], or see for
example [2.1], p. 98), and in turn to Euler’s formula when ˇ D 0. As for the simple
column, Eq. (4.13) can be rewritten as
 
Wn
Wn D Pn  W0n 4.14
P
which is similar to Eq. (4.11) and represents the Southwell method for a column
on an elastic foundation.
This application of Southwell’s method has not been verified experimentally,
but the case of a column with a single mid-point elastic support has been analyzed
and experimentally verified by Hayashi and Kihira [4.28] for a range of spring
constants (see Figure 4.13).
Donnell suggested an alternative manner of plotting the results. If Eq. (4.13) is
solved for Pn , one writes
 
P
Pn D P C W0n . 4.15
Wn
Differentiating Eq. (4.15), remembering that W0n and Pn are constant with respect
to Wn and Pn , one obtains
 
P
0 D dP C W0n d
Wn
Application of the Southwell Method to Columns, Beam Columns & Frames 199

Figure 4.13 Southwell plot for a column with a single mid-point elastic support of different
spring constants k (from [4.28])

and hence
dP
W0n D  . 4.16
dP/Wn 
If one plots, therefore, experimental values of P and P/Wn  against each other
(Figure 4.14d), a straight line is obtained whose intercept on the P axis is Pn and
whose negative slope is W0n . This may be a little more convenient for determina-
tion of the critical load Pn , since the slope has not to be measured. Furthermore
since measurements are usually taken at equal increments of load, the points are
more evenly spaced. Donnell’s proposal has not been widely accepted, except
as a basis for the force/stiffness method [4.29], which is essentially the Donnell
plot with abscissa and ordinate interchanged, and which will be further discussed
in Chapter 8 in this Volume and Chapter 15, Volume 2. Most investigators have
continued to use the original Southwell plot.
Before discussing the extension of Southwell’s method to plates (which will
be considered in Chapter 8), Donnell singled out one plate problem, which is
in a class by itself, the flat panel hinged on three sides and free on the fourth
(Figure 4.15). This is so, because a good approximation can be obtained by
assuming the deformed shape to be a developable surface, so that the extensional
stresses can be neglected in the strain energy. The initial deflection and additional
200 Elements of A Simple Buckling Test A Column under Axial Compression

Figure 4.14 Different forms of Southwell plots: (a) the original Southwell plot [4.12],
(b) Southwell plot for nth critical load, (c) the Southwell Lundquist plot
[4.25], (d) Donnell’s alternative manner of presenting the Southwell plot [4.26],
(e) Southwell plot expressed in strains, bending strain εb and average axial
strain εc

deflection can be assumed as


s nx 
w0 D 

n W0n sin 
b L 

and 4.17


s nx 

wD n Wn sin

b L
Application of the Southwell Method to Columns, Beam Columns & Frames 201

Figure 4.15 Flat panel hinged on three sides and free on


the fourth, subjected to unidirectional in-plane
compression

where L is the length and b the width of the flat panel, x the axial and s the lateral
coordinate. Again from energy considerations he obtained
     
Wn Et3 b n2 2 61  v Wn
PD C D Pn
W0n C Wn 121  v2  L2 b2 W0n C Wn
4.18
where t is the thickness of the panel and Pn is the critical axial load for the nth
type of displacement, with no initial curvature. Equation (4.18) is essentially the
same as that for a column and hence application of the Southwell plot is justified.
One may note that, following Donnell, the Southwell method was formulated here,
and in Figures 4.14b and 4.14d, for the nth critical load, instead of the lowest one
as usual, to show that it can be employed also for higher critical loads. This has
been discussed by Donnell [4.26], who suggested that harmonic analysis should be
used in conjunction with Southwell’s method, and by Tuckerman [4.30], who also,
with the aid of McPherson and Levy, demonstrated experimentally the application
to the second and third critical load of a column. Figure 4.16 shows a typical
Lundquist Southwell plot for the first and second modes of an eccentrically loaded
column formulated in strains instead of displacements.
Strains are often employed instead of displacements in the Southwell method
(see also Figure 4.14e) since strains can be conveniently measured with strain
202 Elements of A Simple Buckling Test A Column under Axial Compression

Figure 4.16 Southwell Lundquist plots for the first and second modes of an eccentrically
loaded column (from [4.30])

gages. One can readily show that the use of bending strains εb instead of the
lateral deflection W, is permissible. For from elementary bending theory, the axial
strain
ε D zw,xx 4.19
and the bending strain εb is the difference between εx at the extreme fibers, z D
h/2 and z D h/2, and is proportional to εx z D h/2 and thus to w,xx . Hence if
Eq. (4.11) is differentiated twice with respect to x it becomes
w,xx x D L/2
w,xx x D L/2 ³ PE  w0,xx x D L/2 4.20
P
as W D wx D L/2, etc. One can therefore write
εx x D L/2
εx x D L/2 ³ PE  w0,xx x D L/2 4.21
P
or
εb
εb ³ PE  W0,xx . 4.22.
P
Sometimes the load P is replaced by the average axial strain εc , which is propor-
tional to it, and then the Southwell plot is εb /εc  versus εb (see Figure 4.14e).
The flat panel of Figure 4.15 is also of considerable practical significance, as it
represents the equal legged angle, a widely used structural element. Bridget, Jerome
and Vosseller [4.31] carried out at Caltech in 1933 a series of tests on compressed
duralminum angle columns, under the guidance of Donnell, and showed the appli-
cability of the Southwell method irrespective of failure in the column or plate
mode. They used the Southwell method for better presentation of their results,
since it eliminated most of the imperfection effects (geometric imperfections, load
eccentricities and minor variations in stress distributions) related to specific spec-
imens. This “smoothing”, very useful in parametric studies, is demonstrated very
clearly in Figures 4.17 and 4.18 (reproduced from [4.31]). Figure 4.17 shows the
critical loads, for the cases in which plate buckling occurred, as recorded from tests
without the use of the Southwell method, whereas Figure 4.18 shows the buckling
load of the specimens computed by Southwell method from the same set of data,
resulting in a significant reduction of scatter.
Application of the Southwell Method to Columns, Beam Columns & Frames 203

Figure 4.17 Experimental buckling loads P versus width of sides W for angle columns, as
measured (for the case in which plate buckling occurred, from [4.31])

In passing, it may be mentioned that Bridget et al. designed their experiments


with great care, (see also [2.1], p. 403) in particular with respect to the boundary
conditions. They employed the idea of varying end conditions, by enabling
movement of the support ball position under load, to eliminate initial load
eccentricities and to counteract initial curvature or “crookedness”, originated by
von Kármán [4.4].
Though some applications of the Southwell plot to other structures have been
considered, the discussion so far has dealt primarily with simply supported columns
for which Southwell derived his method. As has been shown, the theoretical basis
has been broadened in 1938 by Donnell [4.26] and then in 1939 by Tuckerman
[4.30], who showed that by applying Westergaard’s general theory for buckling
of elastic structures [4.32], Southwell’s method or Lundquist’s modification can
be generalized and applied to a broader class of structures. His elegant derivation,
however, did not yield detailed justifications for practical structures beyond those
considered earlier.

4.6.3 Applications to Frames and Lateral Buckling of Beams

Ariaratnam [4.33] demonstrated analytically that the method also applies to


columns with different combinations of end conditions and with varying flexural
rigidity, as well as to plane frameworks. His derivation considered combinations
of pin end, fixed end and free end boundary conditions for the column. For
frameworks, justification of the validity of Southwell’s method was presented both
for buckling in the plane of the structure and for lateral buckling out of its plane.
Ariaratnam’s analysis justified earlier intuitive extensions of Southwell’s method
204 Elements of A Simple Buckling Test A Column under Axial Compression

Figure 4.18 “Smoothing” data with the Southwell method: buckling loads P versus width
W, computed by the Southwell method from the same data as in Figure 4.17
(from [4.31])

to experiments on frameworks like that of Merchant [4.34], who only proposed


the extended method for in-plane buckling; that of Murray [4.35] who applied the
method to lateral buckling of the members of a truss and used it to determine the
experimental buckling load, which was 8 percent below his approximate theoretical
prediction for lateral buckling; and those of Gregory [4.36] [4.39] who applied
the method to tests of in-plane and out-of-plane (torsional-flexural) buckling of
triangular frames and lattice girders. Gregory used strains and not displacements
and his experimental results yielded very consistent linear Southwell plots. The
buckling loads obtained from those plots were very close to the theoretical
predictions for the corresponding perfect frames: between 3.2 percent above and
4 percent below the theoretical values. In one case [4.38] the maximum load
reached in the test was 9 percent below that obtained from the Southwell plot,
which was about 1 percent below the theoretical value.
Southwell’s method has also been extended by Horton, Cundary, and Johnson,
in their 1967 review of the application of the method to elastic column and plate
structures [4.40], to lateral buckling of beams. By the Rayleigh Ritz method
approximate relationships between the load and both the lateral and torsional
deformations were derived for the case of a deep beam subjected to a concen-
trated force applied in the plane of the beam. Assuming single term expressions
for initial and total deflections and rotations, an approximate relationship between
load and rotation, for small rotations, was arrived at
Application of the Southwell Method to Columns, Beam Columns & Frames 205
 
 0
 D Pcr  4.23
P 2
which is of the Southwell type.
Massey [4.41] had shown earlier experimentally that the Southwell method could
be adapted to lateral instability of rectangular section aluminum alloy beams, loaded
with a concentrated vertical load at midspan. The predictions, using a modified
Southwell plot, were 0.72 3.7 percent above the experimental collapse loads, for
the three beams tested. Way [4.42] re-examined the problem experimentally, first
repeating Massey’s work and then extended it to large deflections and beams elas-
tically restrained at midspan. He showed that the Southwell plot of υ/P versus
υ, for example for the typical plot shown in Figure 4.19 (from [4.42]), where υ is
the lateral midspan deflection (proportional to  of Eq. 4.23), yields a critical load
prediction 3.3 percent above the test value. Way’s other experiments also yielded
similar agreement between the Southwell plot predictions and test values.
Leicester [4.43] extended the validity of the Southwell method by presenting
a theoretical justification for beam-columns. He also considered the special case
of a beam loaded through its shear center which requires a modification to the
Southwell plot. Then for experimental verification, tests on two beams, one loaded
through the shear center and one loaded off the shear center, were carried out.
The test specimens were made of hardwood (Leicester worked at the Division of
Forest Products of CSIRO, Australia), and had significant imperfections purposely
imparted by an interesting process. Each beam was soaked in water and allowed to
dry for half a day while subjected to 90 percent of its Euler load. This procedure
caused the beams to develop a permanent set with a shape close to that of the
buckling mode, and hence made them very appropriate specimens for verification of

Figure 4.19 Southwell plot for lateral instability of a deep beam subjected to a concentrated
load at its midspan (from [4.42])
206 Elements of A Simple Buckling Test A Column under Axial Compression

the Southwell method. One can nearly classify them as demonstration experiments
of the kind discussed in Section 4.4.

4.6.4 Southwell’s Method as a Nondestructive Test Method

Another aspect of Southwell’s method was already pointed out by Donnell in 1938
[4.26] its usefulness as a nondestructive test method. It permits the stability limit
to be determined without destroying the structure. This advantage of the Southwell
method was exploited by Wilson, Holloway and Biggers [4.44] in their experiments
on expensive tapered column models, by carrying their tests, and the data recording,
only up to loads which gave a definite straight line on the Southwell plot. Thus they
prevented damage due to plastic bending which would have occurred at the high-
load nonlinearities of the plot. The use of the Southwell plot as a nondestructive
test method has been extensively studied for shells with as yet not completely
conclusive results and will be discussed in Chapters 9, 13, and 16 of Volume 2.
For columns, the Southwell method is universally accepted as a nondestructive
test technique for determination of elastic stiffness properties of actual structural
components. When the struts have very small initial curvatures, as is common
with modern manufacturing quality control, deflection measurement errors are
introduced because the deflections remain relatively small with load. Hence an
alternative method of testing has been proposed by Tsai [4.45] to load the strut with
an intentional eccentricity to compensate for the low initial geometric imperfection.
Essentially this proposal uses and emphasizes the well known concept of approx-
imate equivalence of the effects of initial curvature and load eccentricity, which
has been employed sometimes in column tests to counteract the initial curvature
of the specimens (see [2.1], pp. 28 36, 190 192).
Rederiving the Southwell type equation in the presence of a predetermined load
eccentricity e at both ends of the strut, it becomes instead of Eq. (4.11)
 
1  
W    1 PE
W D PE  W0  e   P  1 . 4.24
P cos P
2 PE
Simplification of Eq. (4.24), by expanding the cosine, series conversion and multi-
plication, yields a first order approximation
   
W 2 W
W D PE  W0  e D PE  W0  1.234e. 4.25
P 8 P
For W0 D 0 this equation is identical to one derived by Sechler in 1952 [4.46].
Equation (4.25) is very close to a similar approximate formula
   
W 4 W
W D PE  W0  e D PE  W0  1.273e 4.25A
P  P
derived in the early thirties ([2.1], p. 191). With a reasonable load eccentricity e,
the deflections would be larger at low axial forces and hence the accuracy of the
measurements for the Southwell plot would improve. It has to be remembered,
Remarks on the Applicability of the Southwell Plot 207

however, that this method is useful only when the column is very straight, and
in introducing the load eccentricity care has to be exercised that the test points
remain in the elastic regime.

4.7 Remarks on the Applicability of the Southwell Plot


As has been shown, Southwell’s method has been applied successfully to many
types of columns, beam-columns and frameworks, and has been modified and
extended for more convenient use. The 1939 warning of Ramberg et al. [4.24]:
“that it must not be concluded from the success of Southwell’s method in all those
cases in which the existence of a straight-line relation . . . was established over a
large range of deformations, that Southwell’s method is applicable to the whole
range of primary instabilities that may be encountered in monocoque construction,”
has been heeded and theoretical justifications have been derived for various types
of structures, as pointed out in this chapter. Extensions to plates and shells and
plastic buckling have been widely studied and are discussed in Chapters 8, (this
volume) and 9, 13 and 16 of Volume 2. The general conclusion is that, with certain
significant limitations, the Southwell plot, and its extensions, have a general validity
for practically all linear instability problems and sometimes even beyond that.
The limitations of the applicability and validity have been pointed out by many
investigators (e.g., [4.20] and [4.26]) and have been elucidated by Roorda [4.47].
Recalling first the differences between various types of buckling behavior
(Figure 4.20) he explains:
“In the linear theory of elastic stability, any perfect structural system that yields
a well-behaved eigenvalue problem gives rise to the load-deflection characteristic
depicted in Figure 4.20(a). The trivial (unbuckled) solution υ D 0 is crossed by the
horizontal line P D PCR at the point of bifurcation. This type of characteristic might
be described as a neutral characteristic and it arises purely from the linearization
of the problem. Small imperfections give rise to equilibrium paths in the form of
rectangular hyperbolas as shown. In essence, Southwell’s method is based on the
neutral characteristic.
In the nonlinear theory of elastic stability the post-buckling behavior is generally
not of the neutral type but takes one of three forms, depending on the nature of
the nonlinearities in the system (in the following, the deflections are considered to
be finite but relatively small so that only the initial post-buckling behavior need
be considered). The possible load-deflection curves are depicted in Figure 4.20(b),
4.20(c) and 4.20(d), and may be described as the asymmetric, stable-symmetric, and
unstable-symmetric characteristics, respectively. The corresponding load-deflection
curves for an imperfect system are also indicated on the diagram. These are now
not rectangular hyperbolas but have the perfect equilibrium curves as asymptotes.
. . . It now becomes evident that a post-buckling behavior other than the neutral
type gives rise to a curved Southwell line.
. . . Typical Southwell lines corresponding to the four buckling types, namely
neutral, asymmetric, stable-symmetric, and unstable-symmetric, are drawn in
208 Elements of A Simple Buckling Test A Column under Axial Compression

Figure 4.20 Different types of buckling behavior (perfect structures are presented by solid
lines and imperfect ones by broken lines): (a) neutral load-deflection characteristic
(arising from linearization of the problem), (b) asymmetric load-deflection
characteristic, (c) stable-symmetric load-deflection characteristic, (d) unstable-
symmetric load-deflection characteristic

Figure 4.21(a), 4.21(b), 4.21(c) and 4.21(d), respectively. The initial slope in each
case is 1/PCR . For the neutral case, this slope is maintained for all values of υ. For
the asymmetric case, the slope decreases as υ increases in the positive direction
and increases as υ increases in the negative direction. For the stable symmetric
and unstable symmetric cases, the slope decreases and increases, respectively, as
jυj increases.
On the basis of these diagrams, it is now possible to draw certain conclusions
regarding the validity of the Southwell procedure as the measured deflections
become large.
For the neutral buckling characteristic there is no problem. The asymmetric buck-
ling characteristic is the most interesting. If in an experimental structure the imper-
fections are such that they generate a load-deflection curve with monotonically
increasing load (positive deflections in Figure 4.20(b)), then the best straight line
fitted to the experimental points in a Southwell plot will have a slope which is
less than the true slope at zero deflection. Hence, the Southwell procedure would
overestimate the critical load. If, on the other hand, the imperfections generate a load
deflection curve in which the load reaches a local maximum (i.e. negative deflections
in Figure 4.20(b)), then the Southwell procedure underestimates the critical load.. . .
A steep ideal post-buckling curve may give rise to a considerable discrepancy.
Similarly, in the case of symmetric buckling characteristics (Figures 4.20(c,d)),
the following conclusion is drawn: The Southwell procedure overestimates the crit-
ical load for a stable symmetric characteristic and underestimates it for an unstable
symmetric characteristic, regardless of the sense of the initial imperfections.. . .
The sharper the initial curvature in the ideal post-buckling path, the greater the
difference between PCR and its estimated value.”
Remarks on the Applicability of the Southwell Plot 209

Figure 4.21 Southwell lines corresponding to the four types of buckling behavior shown
in Figure 4.20 (from [4.47]): (a) neutral, (b) asymmetric, (c) stable symmetric,
(d) unstable symmetric

Roorda summarizes this exposition with advice to the experimenter. “If the
Southwell procedure is applied to structures other than the column, great care
must be taken in the interpretation of the results. Correct interpretation of the
Southwell plot for such structures depends upon a knowledge of the post-buckling
behavior of the idealized structure.”
In other words, for column type structures, as discussed here and in Chapter 6, for
which the postbuckling behavior is neutrally stable and the bifurcation modes are
well separated, the Southwell method is a convenient and reliable tool for prediction
of the critical load of the perfect structure. For other types of structures, like plates,
which have stable symmetric postbuckling behavior, or shells, which have unstable
symmetric postbuckling behavior, the Southwell method has limitations, which are
discussed in more detail in Chapters 8 (this volume) and 9, Volume 2.
In Roorda’s paper [4.47] two experiments of the demonstration type are
presented to confirm the arguments and to emphasize the proper interpretation
procedure, a Warren truss and a shallow frame. Here the Warren truss (Figure 4.22)
will be discussed. It consisted of triangulated frames made of high tensile steel
strips of 1 in. by 1/16 in. cross section. The truss had a span of 36 in. and the
members were rigidly jointed by fixing the ends in slotted joint blocks, made
from Duraluminum rods of 1.5 in. diameter, as shown in Figure 4.23(a). The truss
was supported on knife edges with a roller support (Figure 4.23(b)) at one end to
prevent horizontal reactions. The load was applied vertically at one of the joints in
the top chord and transmitted to the truss through a double knife edge and movable
knife seat arrangement (Figure 4.23(c)). The two knife seats could be moved to
210 Elements of A Simple Buckling Test A Column under Axial Compression

Figure 4.22 Buckling modes of a Warren truss (from [4.47]): (a) unbuckled form, (b) stable
buckling, (c) unstable buckling

positions near the joint by turning the screws, thus allowing slight variations of
the eccentricity of loading. The rotation of the loaded joint served as a deflection
parameter and was measured optically during the loading process. A semi-rigid
loading device, consisting of a spring-balance and screw jack combination, was
used so that unstable branches of the load-deflection curves could be traced. Two
such load-deflection curves were obtained, one for a value of load eccentricity
which induced counter-clockwise joint rotation (unstable mode of buckling) and
the other for a slightly different eccentricity which induced clockwise joint rotation
(stable mode of buckling). These curves appear in Figure 4.24. The dotted portion
in the unstable load-deflection curve could not be obtained due to insufficient
rigidity of the loading device.
Although in real structures the perfect post-buckling curve can never be attained,
it is clear from these experimental results that this curve would have a finite slope
Remarks on the Applicability of the Southwell Plot 211

Figure 4.23 Roorda’s Warren truss experiment details (from [4.14]): (a) the truss tested,
(b) the roller support, (c) frame joint detail with loading arrangement
212 Elements of A Simple Buckling Test A Column under Axial Compression

Figure 4.24 Experimental load joint rotation curves for Roorda’s Warren truss (from [4.47])

at the point of bifurcation, i.e. the experimental system gives rise to an asymmetric
buckling characteristic.
Plotting the experimental points in accordance with the Southwell procedure
yields the curve shown in Figure 4.25 which is of the type shown in Figure 4.21(b)
(i.e., there is no reversal of curvature as the curve passes through the origin). The
correct buckling load is given by the inverse of the slope of the tangent to the
curve at the origin, and is approximately equal to 65.0 lb. One should note that
due to the curvature of the Southwell plot, it is important to use the tangent at the
origin for a correct interpretation. Photographs of the two buckling modes for this
experimental system are shown in Figure 4.22.
Before closing this section it should be noted that, as pointed out by Bush-
nell [4.48], the Southwell method can also be used for a “non-experimental”
task “in conjunction with computer programs to ascertain certain unknown or
doubtful characteristics of a complicated structure, such as effective stiffness or
boundary conditions. These characteristics would be changed in repeated runs of
the computer program until the critical bifurcation load predicted by the program
agrees with that from the Southwell plot.” Jones, Costello and Reynolds [4.49]
applied the Southwell plot in this “non-experimental” mode to evaluate their
References 213

Figure 4.25 Southwell plot for Roorda’s Warren truss (from [4.47])

numerical calculations for buckling of a ring, as well as a ring-stiffened cylinder


under external pressure.

References

4.1 Euler, L., Sur le force des colonnes, Mem. de l’Acad. de Berlin, Berlin Annee 1759,
XIII, 252 257. English Translation by Van der Brock, Am. J. Physics, 15, July 1947,
309 318.
4.2 van Musschenbroek, P., Introductio ad cohaerentiam firmorum in Physicae
experimentales et geometricae Dissertationes, Lugduni (Leiden) 1729, 652 660 and
Table 27.
4.3 Salmon, E.H., Columns A Treatise on the Strength and Design of Compression
Members, Oxford Technical Publications, London, 1921.
4.4 von Kármán, Th., Untersuchungen über Knickfestigkeit, Mitteilungen über
Forschungsarbeiten auf dem Gebiet des Ingenieurwesens, Verein Deutscher
Ingenieure, Heft 81, Berlin, 1910.
4.5 Prandtl, L., Kipp-Erscheinungen, Inaugural Dissertation, Ludwig-Maximilians-
Universität Munich, 1899, Buchdruckerei Robert Stich, Nurenburg.
4.6 Beedle, L.S. and Tall, L., Basic Column Strength, Proc. American Soc. of Civil Engi-
neers, 86, (ST-7), July 1961, Proc. Paper 2555, 139, 173.
4.7 Tall, L., Centrally Compressed Members, in Axially Compressed Structures, Stability
and Strength, R. Narayanan, ed., Elsevier Applied Science Publishers, London, 1982,
1 40.
214 Elements of A Simple Buckling Test A Column under Axial Compression

4.8 von Tetmajer, L., Die Gesetze der Knickungs-und der zusammengesetzten Druck-
festikeit der technisch wichtigsten Baustoffe. 3rd edn., Franz Deuticke, Leipzig and
Vienna, 1903.
4.9 Kirsch, B., Ergebnisse von Versuchen über die Knickfestikeit von Säulen mit fest
eingespannten Enden, Zeitschrift d. Verein deutscher Ingenieure, Berlin, June 3, 1905,
907 915.
4.10 Palmer, A.C., Structural Mechanics, Clarendon Press, Oxford, 1976, 202 206,.
4.11 Kirby, P.A. and Nethercot, D.A., Design for Structural Stability, Granada Publishing,
London, 1979.
4.12 Southwell, R.V., On the Analysis of Experimental Observations in Problems of
Elastic Stability, Proc. Royal Society, (London), Series A, 135, 1932, 601 616.
4.13 Brivtec, S.J. and Chilver, A.H., Elastic Buckling of Rigidly-Jointed Braced Frames,
Journal of the Engineering Mechanics Division, ASCE, 89, (EM6), Proc. Paper 3736,
1963, 217 255.
4.14 Roorda, J., Instability of Imperfect Elastic Structures, Ph.D. Thesis, University of
London, 1965.
4.15 Roorda, J., Experiments in Post-Buckling in Buckling of Elastic Structures, Solid
Mechanics Division, University of Waterloo, Waterloo, Ontario, Canada, 1980, 73 93.
4.16 Walker, A.C., Croll, J.G.A. and Wilson, E., Experimental Models to Illustrate the
Non-Linear Behavior of Elastic Structures, Bulletin Mech. Engineering Education,
10, 1971, 247 259.
4.17 Ayrton, W.E. and Perry, J., On Struts, The Engineer, Dec. 10, 1886, 464 465, Dec.
24, 1886, 513 515.
4.18 Hoff, N.J., The Analysis of Structures, John Wiley & Sons, New York, 1956, 230 242.
4.19 Murray, N.W., Introduction to the Theory of Thin-Walled Structures, Oxford Univer-
sity Press, Oxford, 1984, 12 15.
4.20 Spencer, H.H. and Walker, A.C., Critique of Southwell Plots with Proposals for
Alternative Methods, Experimental Mechanics, 15, (8), 1975, 303 310.
4.21 Robertson, A., The Strength of Struts, Selected Paper No 28, Institution of Civil
Engineers (UK), 1925.
4.22 Gough, H.J. and Cox, H.L., Some Tests on the Stability of Thin Strip Material Under
Shearing Forces in the Plane of the Strip, Proc. Royal Society (London), Series A,
137, 1932, 145 157.
4.23 Fisher, H.R., An Extension of Southwell’s Method of Analyzing Experimental Obser-
vations in Problems of Elastic Stability, Proc. Royal Society (London), Series A, 144,
1934, 609 630.
4.24 Ramberg, W., McPherson, A.E. and Levy, S., Experimental Study of Deformation
and of Effective Width in Axially Loaded Sheet-Stringer Panels, NACA TN 684,
1939.
4.25 Lundquist, E.E., Generalized Analysis of Experimental Observations in Problems of
Elastic Stability, NACA TN 658, 1938.
4.26 Donnell, L.H., On the Application of Southwell’s Method for the Analysis of Buck-
ling Tests, Stephan Timoshenko 60th Anniversary Volume, McGraw-Hill, New York,
1938, 27 38.
4.27 Engesser, F., Die Sicherung offener Brücken gegen Ausknicken, Zentralblatt der
Bauverwaltung, 1884, (40), 415 417 and 1885, (7), 71 72.
4.28 Hayashi, T. and Kihira, M., On a Method of Experimental Determination of the
Buckling Load of an Elastically Supported Column, Proc. 10th Japan Congress on
Testing Materials, 1967, 163 165.
References 215

4.29 Jones, R.E. and Green, B.E., Force/Stiffness Technique for Nondestructive Buckling
Tests, Journal of Aircraft, 13, April 1976, 262 269.
4.30 Tuckerman, L.B., Heterostatic Loading and Critical Astatic Loads, Research Paper
RP1163, Jour. Res. National Bureau of Standards, 22, 1939, 1 18.
4.31 Bridget, F.J., Jerome, C.C. and Vosseler, A.B., Some New Experiments on the Buck-
ling of Thin Wall Construction, Transactions of the American Society of Mechanical
Engineers, 56, 1934, 569 578.
4.32 Westergaard, H.M., Buckling of Elastic Structures, Transactions of the American
Society of Civil Engineers, 85, Paper 1490, 1922, 566 676.
4.33 Ariaratnam, S.T., The Southwell Method for Predicting Critical Loads of Elastic
Structures, Quarterly Journal of Mechanics and Applied Mathematics, 14, Pt. 2,
1961, 137 153.
4.34 Merchant, W., The Failure Load of Rigid Jointed Frameworks as Influenced by
Stability, The Structural Engineer (UK), 32, 1954, 185 190.
4.35 Murrey, N.W., A Method of Determining an Approximate Value of the Critical
Loads at which Lateral Buckling Occurs in Rigidly Jointed Trusses, Proc. Institution
of Civil Engineers (UK), 7, 1957, Paper 6209, 387 403.
4.36 Gregory, M., The Use of the Southwell Plot on Strains to Determine the Failure Load
of a Lattice Girder when Lateral Buckling Occurs, Australian Journal of Applied
Science, 10, (4), 1959, 371 376.
4.37 Gregory, M., The Buckling of an Equilateral Triangular Frame in its Plane, Australian
Journal of Applied Science, 10, (4), 1959, 377 387.
4.38 Gregory, M., The Application of the Southwell Plot on Strains to Problems of Elastic
Instability of Framed Structures, where Buckling of Members in Torsion and Flexure
Occurs, Australian Journal of Applied Science, 11, (1), 1960, 49 64.
4.39 Gregory, M., The Use of Measured Strains to Obtain Critical Loads, Civil Engi-
neering and Public Works Review (UK), 55, (642), Jan. 1960, 80 82.
4.40 Horton, W.H., Cundary, F.L. and Johnson, R., The Analysis of Experimental Data
Obtained from Stability Studies on Elastic Column and Plate Structures, Israel
Journal of Technology, 5, (1 2), 1967, 104 113.
4.41 Massey, C., Southwell Plot Applied to Lateral Instability of Beams, Engineer, 218,
(5666), Aug. 1964, 320.
4.42 Way, E.R., The Lateral Instability of a Simply Supported Deep Beam Subjected to a
Concentrated Load at its Centroid, Engineer Thesis, Stanford University, California,
1967.
4.43 Leicester, R.H., Southwell’s Plot for Beam Columns, Journal of the Engineering
Mechanics Division, ASCE, 96, (EM6), Proc. Paper 7750, 1970, 945 965.
4.44 Wilson, J.F., Holloway, D.M. and Biggers, S.B., Stability Experiments on the Stron-
gest Columns and Circular Arches, Experimental Mechanics, 11, (7), 1971, 303 308.
4.45 Tsai, W.T., Note on Southwell’s Method for Buckling Tests of Struts, Journal of
Applied Mechanics, ASME, 53, 1986, 953 954.
4.46 Sechler, E.E., Elasticity in Engineering, Dover Publications, New York, 1968, 364.
4.47 Roorda, J., Some Thoughts on the Southwell Plot, Journal of the Engineering
Mechanics Division, ASCE, 93, (EM6), Proc. Paper 5634, 1967, 37 47.
4.48 Bushnell, D., Computerized Buckling Analysis of Shells, Martinus Nijhoff,
Dordrecht/Boston, 1985.
4.49 Jones, R.F., Costello, M.G. and Reynolds, T.E., Buckling of Pressure Loaded Rings
and Shells by the Finite Element Method, Computers and Structures, 7, 1977,
267 274.
5
Modeling Theory and
Practice

5.1 Mathematical and Physical Modeling

Before continuing the discussion of buckling and postbuckling experiments on


columns, it may be useful to pause and consider the general problem of modeling.
First we have to differentiate between physical and mathematical modeling. A
physical model represents the real world, the real physical engineering problem, by
identifying the primary factors that affect the behavior of a structure (in our case)
and describes a simpler structure, or one more amenable to testing and measure-
ments, that still demonstrates the main response behavior of the real structure.
A mathematical model on the other hand is an abstraction of the real problem
in the conceptual world, it is the physical problem transformed into an idealized
image that has built-in assumptions and approximations, but constitutes the basis
for analysis and predictions. Often the physical model leads to the mathematical
one, but it must be remembered that the mathematical model is an idealized abstract
representation of the physical one.
Mathematical modeling is an essential step in the process of analysis, compu-
tation and prediction, (see for example [5.1]), whereas physical modeling leads to
experimental investigations, to model analysis of structures.
In civil engineering primarily, structural modeling has been extensively used for
more than half a century as an experimental method for solution of strength and
deformation problems of structures. It has been considered as a parallel design tool
to analysis (see for example [5.2]) and has been the subject of special textbooks (for
example [5.2] or [5.3]) and numerous papers. Since most experimental studies on
buckling and postbuckling behavior of structural elements, also in other engineering
disciplines, are carried out on models of the actual structures, appropriate modeling
is indeed an essential part of experimental investigation.
The theory of models, and in particular dimensional analysis, can be an important
guide to the experimenter in the choice of his models. Hence, though the principles
of dimensional analysis are usually well known to scientists and engineers and are

Buckling Experiments: Experimental Methods in Buckling of Thin-Walled Structures: Basic Concepts, Columns, Beams and
Plates – Volume 1. J. Singer, J. Arbocz and T. Weller Copyright © 1998 John Wiley & Sons, Inc.
218 Modeling Theory and Practice

given in many textbooks (for example [5.3] [5.11]), the main concepts are briefly
reviewed in the following section.

5.2 Dimensional Analysis


Dimensional analysis is a method by which information about a phenomenon is
deduced from one postulate that the phenomenon can be described by a dimension-
ally correct equation among certain variables. Dimensional analysis does not yield
a complete solution nor does it reveal the complete character of the phenomenon,
but it reduces the number of variables in a problem. This makes it an important
mathematical tool for experimentalists.
In its simplest form, dimensional analysis can be used to check the dimensional
correctness of equations and to classify them into homogeneous and nonhomo-
geneous ones. An equation is dimensionally homogeneous, if its form does not
depend on the fundamental units of measurement, or in other words if the equation
is valid in all consistent systems of units. A more interesting application of dimen-
sional analysis is the prediction of the qualitative form of unknown mathematical
relationships among physical quantities, whose quantitative relationship can then
often be determined by experiments.

5.2.1 The Procedure in Dimensional Analysis

The first step in the dimensional analysis of a problem is to decide what variables
enter the problem, a decision which, to be meaningful, requires some understanding
of the physical nature of the problem. When the dimensions of these variables have
been determined they are grouped into dimensionless products.
Consider for example, (as in [5.4]) a classical problem of fluid dynamics, a
smooth spherical body of diameter D immersed in a stream of incompressible
fluid moving at a velocity V. The drag force F on the body is represented by an
equation of the form
F D fV, D, ,  5.1
in which  is the mass density of the fluid,  is the coefficient of viscosity of the
fluid (the ratio of the viscous shear stress to the normal velocity gradient) and f
is an unspecified function. Equation (5.1) only says that the drag F depends on the
variables V, D,  and , without indicating the nature of this relationship. As will
be shown shortly, in order to be dimensionally homogeneous Eq. (5.1) must have
the following form:
F/V2 D2  D f1 VD/. 5.2
The function f1 is still unknown, but it depends now on only one variable (VD/)
instead of the four separate variables V, D,  and . Note that the expressions
(F/V2 D2 ) and (VD/) are dimensionless. Such groupings are called dimension-
less products. In general, if L denotes a length, the dimensionless product (VD/)
Dimensional Analysis 219

is called Reynold’s number (usually denoted by R), and the dimensionless product
(F/V2 D2 ) is called a pressure coefficient (often denoted by P), since (F/L 2 ) can
be interpreted as a pressure.
The projected area of a sphere A D /4D2 and hence Eq. (5.2) may be written,
F/V2 A D 1/28/f1 R 5.3
where the term 8/f1 R D CD is called the drag coefficient. The equation for
the drag of a sphere may therefore be written, in the well known form,
F D 1/2CD V2 A 5.4
CD being a function of R. Typical experimental results are usually plotted as
CD versus R, and the plot provides information about the drag forces on smooth
spherical bodies of all sizes in incompressible fluids with any density and viscosity,
flowing at any velocity.
Obviously many other dimensionless products can be formed from the variables
F, V, D, ,  of the example considered, but they will all be found to be of the
form
R a Pb 5.5
where a and b are constant exponents. On the other hand the products R and P are
independent of each other, in the sense of each not being a product of the other,
as can be immediately seen here from the fact that  occurs only in R and F only
in P. Other dimensionless products formed will not be new ones, since they are
all expressible in terms of R and P, as per Eq. (5.5).

5.2.2 The Buckingham Pi Theorem

Generalizing the discussion, one arrives at the following definition: “A set of


dimensionless products of given variables is complete, if each product in the set is
independent of the others, and every other dimensionless product of the variables
is a product of powers of dimensionless products in the set.”
This leads to Buckingham’s theorem: “If an equation is dimensionally homoge-
neous it can be reduced to a relationship among a complete set of dimensionless
products.” The theorem was stated by Buckingham in 1914 [5.12], without a
rigorous proof, was discussed in detail by Bridgman in 1922 [5.13], restated and
proved by van Driest in 1946 [5.14], and presented in 1951 with an algebraic
proof in Langhaar’s book [5.4]. Buckingham denoted the independent dimension-
less products or groups Pi terms, and they are usually designated as i . Hence
the theorem is known as the Buckingham Pi Theorem. A corollary to the theorem,
which may be considered the second part of the theorem, states in general terms
that: “the number of independent dimensionless products required to express a
relationship between the variables in any phenomenon is equal to the number of
derived variables involved n, minus the number of dimensions m, or primary vari-
ables, in which these quantities may be measured”. The number of  terms is
therefore (n  m).
220 Modeling Theory and Practice

In the example considered, the sphere immersed in a stream of incompressible


fluid, F, V, D,  and  are the nD 5 derived variables. After choosing mD 3 of
these such that they contain among them all the primary dimensions (say, V, D and
), one forms dimensionless products with the remaining m  nD 2 variables as
follows: 
1 D Va Db c 
. 5.6
2 D Vd De f F
In terms of the primary dimensions M, L, T these products are:
 a  c   
L M M 
1 D Lb 

T L3 LT
 d  f   . 5.7
L M ML  
2 D Le 
T L3 T2
For 1 , 2 to be dimensionless, the exponents for each of the primary dimensions
must vanish. Hence 
L : a C b  3c  1 D 0
T : a  1 D 0 5.8a
M: cC1D0
yielding a D b D c D 1
and 
L : d C e  3f C 1 D 0
T : d  2 D 0 5.8b
M: fC1D0
yielding d D e D 2 and f D 1. The two dimensionless products are therefore:

1 D /VD D 1/R
. 5.9
and 2 D F/V2 D2  D P
Buckingham’s theorem asserts here that if we assume that the five variables are
related by a dimensionally homogeneous equation, similar to Eq. (5.1),
fF, V, D, ,  D 0 5.10
this equation reduces to
f1 , 2  D 0 or fP, R D 0 5.11
which can also be written as P D f1 R, which is Eq. (5.2).

5.3 Similarity

5.3.1 The Concept of Similarity

Dimensional analysis yields a general theory of model design by establishing the


proper relations between the model and its prototype. Usually geometric simi-
larity is maintained, which means that the parts of the model have the same shape
Similarity 221

as the corresponding parts of the prototype. In general there is a point-to-point


correspondence between a model and its prototype, called homology in geomet-
rical terminology. Two points that correspond to each other are homologous, and
parts of the model and its prototype are homologous if there is a point-to-point
correspondence between them.
Similarity extends to other characteristics besides geometry, like mass distri-
bution or stiffness. For example, in an aircraft wing-flutter model similarity of
mass distribution is required, if not for all the details of the structure at least in
a restricted sense, such that the ratio of masses of segments of the wing and its
model, which are included between homologous cross sections, has to be constant,
i.e. similarity of the spanwise distribution of mass. Furthermore, chordwise distri-
butions of mass, in the same restricted sense, and spanwise distributions of mass
moments of inertia are to be similar, and the ratio of stiffnesses of homologous
cross sections of the wing and its model has to be constant.
Applying dimensional analysis, the general equation for the prototype may be
written as
p D f1p , 2p , . . . np  5.12

where the ’s are a complete set of dimensionless products. For the model, a
similar equation holds
m D f1m , 2m , . . . nm . 5.13

If the model is designed and tested so that


1m D 1p 

2m D 2p 

.. .. 5.14
. . 


nm D np
one obtains 
f1p , 2p , . . . 2n  D f1m , 2m , . . . 2m 
5.15
or p D m
and model and prototype are completely similar. The general Eqs. (5.12) and
(5.13) and their equality, Eq. (5.15), emanate from Buckingham’s theorem.
Equation (5.14) are sometimes called the design conditions and Eq. (5.15) the
prediction equation, and if all the design conditions are satisfied the model is called
a “true” model, as complete similarity is assured, provided that all the pertinent
quantities were included in the dimensional analysis that yielded the Pi terms.

5.3.2 Model Laws

The concept of similarity can also be expressed in a different form, as model laws.
For example, the dimensionless equation defining the drag on a body in a stream
of incompressible fluid (Eq. (5.2)), which was derived earlier, can be considered
as follows: If a geometrically similar model is to be tested, the Reynold’s numbers
222 Modeling Theory and Practice

R of model and prototype should be equal, or


VL/m D VL/p . 5.16
This equation may be written as
KV KL K D K 5.17
where
Kv D Vm /Vp
KL D Lm /Lp
K D m /p
K D m /p etc. 5.18
These K’s are called scale factors. It should be noted that Eq. (5.17) expresses the
requirements of similarity in terms of scale factors, in this case. Such expressions
of these requirements are known as similarity conditions.
Equations (5.2) and (5.17) yield the scale factor for the drag force
KF D K K2V K2L D K2 /K . 5.19
Equations (5.17) and (5.19) represent the model law for the problem of a body
immersed in a stream of compressible fluid.
To express the concept of similarity in terms of scale factors more generally,
two homologous (corresponding) rectangular Cartesian space reference frames are
selected (xp , yp , zp ) and (xm , ym , zm ), with which points in the prototype and the
model are designated respectively. The model and prototype systems are related
by equations defining homologous (corresponding) points and times.
xm D Kx xp 

ym D K y yp 
. 5.20
zm D Kz zp 

tm D K T tp
The constants Kx , Ky , Kz are the scale factors for lengths in the x, y and z
directions. For geometric similarity,
Kx D Ky D Kz D KL . 5.21
For a distorted model, two length scale factors are usually equal, and one
unequal, for example Kx D Ky 6D Kz . In such a case the ratio (Kz /Kx ) is known
as the distortion factor. The constant Kt is called the time scale factor. It should
be remembered that in transient phenomena, states that occur at homologous times
are considered and not simultaneous states.
Kinematic similarity denotes similarity of motions and can be defined as
follows (see for example [5.4]): “The motions of two systems are similar, if
homologous particles lie at homologous points at homologous times”. Homologous
Application to Statically Loaded Elastic Structures 223

(corresponding) points and times are defined by Eq. (5.20). With kinematic
similarity, corresponding components of velocity or acceleration are similar. For
geometric similar systems, satisfying Eq. (5.21), the velocity scale factor is:
KV D KL Kt 5.22
and the acceleration scale factor:
Ka D am /ap D KL /K2t D K2V KL . 5.23
Dynamic similarity occurs if the homologous parts of two systems experience
similar net forces. For geometric similarity and for similar mass distributions,
indicated by the existence of a scale factor for mass
Km D mm /mp , 5.24
the scale factor of the total force is
KF D Km KL /K2t . 5.25
Note that models which satisfy similarity conditions for mass and elasticity (as
mentioned for the example of an aircraft wing-flutter model), are usually known
as dynamically similar models.
It may be noted that complete similarity is sometimes called replica modeling,
especially in cases of structural response to transient loads, which are discussed
in a later section. A replica model is defined as “a model which is geometrically
similar in all respects to the prototype and employs identically the same materials
at similar locations” [5.33]). For geometric similarity with different materials, the
term dissimilar material modeling is often employed, which however assumes
constitutive similarity for the materials. Constitutive similarity means that model
and prototype materials have homologous constitutive properties and homologous
stress-strain curves. Materials possessing constitutive similarity will have identical
scaled strength and stiffness properties.

5.4 Application to Statically Loaded Elastic Structures


5.4.1 Prescribed Loads

Consider now the application of dimensional analysis to statically loaded elastic


structures. The structure itself may be specified by the modulus of elasticity E and
Poisson’s ratio , by one length  and the ratios r1 , r2 , etc. of all other lengths
to . The loads may be concentrated loads, distributed ones and moments, but for
simplicity only concentrated loads will be considered here. These may be specified
by one of them P and the ratios r11 , r21 , etc. of the other loads to P. The directions
of the loads may be specified by 1 , 2 , etc. A particular stress component at
some point x, y, z, will be given by a relation
D f1 x, y, z, E, , P, , r1 , r2 . . . , r11 , r21 . . . , 1 , 2 . . . 5.26
224 Modeling Theory and Practice

assuming isotropy, homogeneity and Hooke’s law. Since only two primary
variables, or fundamental units, are required here for the measurement of all the
n quantities in Eq. (5.26), force and length, m D 2, and hence n  2 independent
dimensionless products form a complete set of dimensionless products. In the
present case these dimensionless products can be determined intuitively, but will
then also be rederived by a formal systematic approach. Writing the dimensionless
products intuitively, the complete set of dimensionless products consists here of

 , E, x/, y/, z/, P/E2 , , r1 , r2 . . . , r11 , r21 . . . , 1 , 2 . . . 5.27

and by Buckingham’s theorem Eq. (5.26) reduces to

 /E D f[x/, y/, z/, P/E2 , , r1 , r2 . . . r11 , r21 . . . , 1 , 2 . . .]. 5.28

Note that Poisson’s ratio  and the other ratios, r1 , r2 etc., are dimensionless
quantities. For the formal determination of the dimensionless products, only one
load P and one dimension x will be considered, for simplicity. Hence there are
n D 6 variables and Eq. (5.26) is replaced by

D f2 x, E, , P, . 5.29

Since only two fundamental units are required here, force and length, one has to
choose m D 2 variables from Eq. (5.29) such that they contain these two dimen-
sions, say E and . One of the variables in Eq. (5.29), Poisson’s ratio , is already
a nondimensional quantity. Therefore one forms dimensionless products with the
remaining n  m  1D 3 variables as follows:

1 D Ea b P 
2 D Ec d . 5.30

3 D E  x
e f

In terms of the two primary dimensions F, L these products are:


 a 
F 

1 D L b
F 

L2 
 c    
F F 
2 D L d . 5.31
L2 L2 

 e 
F 

3 D L f
L 

L 2

For 1 , 2 , 3 to be dimensionless, the exponents for each of the primary dimen-


sions must vanish. Hence
L : 2a C b D 0
5.32
F: aC1D0
yielding a D 1 and b D 2, and

L : 2c C d  2 D 0
5.33
F: cC1D0
Application to Statically Loaded Elastic Structures 225

yielding c D 1 and d D 0, and



L : 2e C f C 1 D 0
5.34
F: eD0
yielding f D 1. The three dimensionless products are therefore:

1 D P/E2  sometimes called the strain number 
2 D  /E 5.35

3 D x/
and hence
f3 1 , 2 , 3 ,  5.36
which can also be written as
 /E D f[x/, P/E2 , ]. 5.37
If P and  were chosen as the m D 2 variables from Eq. (5.29), instead of E, ,
the same process would have yielded an alternative grouping of dimensionless
products instead of Eq. (5.35).

or more conveniently the strain numberP/E2 , 


1 D E2 /P since any inversion of a dimensionless product 
only changes the unknown function of the products . 5.35A


2 D  2 /E 

3 D x/
This would give an alternative equation to Eq. (5.37),
N
 2 /P D f[x/, P/E2 , ]. 5.37A
With the additional ratios r1 , r2 etc. and the dimensionless products (y/) and
(z/), obtained in the same manner as (x/), Eq. (5.37) would again become
Eq. (5.28). With the grouping of Eq. (5.35A) an alternative equation to Eq. (5.28)
would result,
N
 2 /P D f[x/, y/, z/, P/E2 , , r1 , r2 . . . , r11 , r21 . . . , 1 , 2 . . .]
5.28A
It may be noted, that a more precise formulation of the corollary to Buckingham’s
Pi Theorem (or its second part) would be that “the number of dimensionless prod-
ucts in a complete set is equal to the total number of variables minus the rank of
their dimensional matrix” as stated by Langhaar [5.4]. To apply this formulation
here one would write the dimensional matrix of the n variables of Eq. (5.26) and
find that the rank1 of this matrix is 2 (see for example [5.10] p. 284). The number of
dimensionless products necessary for a complete set is therefore n  2, as obtained

1 In algebra, the rank of a matrix is said to be r, if the matrix contains a nonzero determinant of order r,
and all determinants of order greater than r that the matrix contains have the value zero.
226 Modeling Theory and Practice

earlier from consideration of fundamental units required for the measurement of


all the n variables.
Equation (5.28) is now applied to both model and prototype. For geometric
similarity r1 , r2 , . . . have to be the same for both model and prototype. Making r11 ,
r21 , . . . and 1 , 2 . . ., the same for both means similarity of load distribution, and
making (x/), (y/), (z/) the same means that the stress is measured at the point
in the model corresponding to the equivalent point in the prototype (homologous
points). Complete similarity is obtained when all the independent dimensionless
products of Eq. (5.28) are the same for model and prototype. Then
 m /Em  D  p /Ep 
or p D Ep /Em  m . 5.38
The scaling rules for the loads are evident from
Pm /Em 2m  D Pp /Ep 2p 
or Pm /Pp  D Em /Ep m /p 2 5.39
and corresponding expressions for other types of loads.

5.4.2 Displacements and Strains

Similar dimensional analyses for a displacement u, or for a strain ε yield

u/ D f4 [x/, y/, z/, P/E2 , , . . . , ] 5.40


and
ε D f5 [x/, y/, z/l, P/E2 , , . . . , ]. 5.41
Hence
um /m  D up /p  and εm D εp . 5.42
It is important to point out here that the displacements have not been assumed to
be small. The similarity rules discussed apply, therefore, also to large deformations
of flexible structures (made of a material obeying Hooke’s law), where the stresses
and displacements are not in general proportional to the loads. For example, very
flexible steel springs or thin plates compressed beyond buckling are in this class,
as long as the proportional limit is not exceeded.
The dimensional analysis also shows how curves of data obtained in model
tests should be plotted to apply directly to the prototype. For example, if a stress
is plotted versus a load P, the curve showing ( 2 /P) versus the strain number
(P/E2 ) is a graphical representation of Eq. (5.28A), or the curve of ( /E) versus
(P/E2 ) represents Eq. (5.28). With the similarity conditions being satisfied, these
curves are identical to the corresponding one obtained from data measured on the
prototype. In elastic structures, the strain number (P/E2 ) has therefore a similar
role to that of the Reynold’s number in fluid dynamics.
Application to Statically Loaded Elastic Structures 227

If the structure is not all of the same material, one has to include in the list of
variables the moduli E, E1 , E2 , etc., and Poisson’s ratios , 1 , 2 , etc. Additional
dimensionless groups (E1 /E), (E2 /E), etc., and 1 , 2 , etc., will then appear, which
have to be the same for model and prototype to satisfy the similarity conditions.
For stiff structures, where the deformations do not significantly affect the action
of the loads, the stresses, strains and displacements are always linear functions
of the load (as follows from the linearity of the basic differential equations and
boundary conditions of the classical theory of elasticity). For such linear struc-
tures the dimensional analysis becomes simpler. Since must be proportional to
P,  /E in Eq. (5.28) cannot be an unknown function of (P/E2 ), and therefore the
right-hand side of Eq. (5.28) f[. . .] should be independent of the group (P/E2 ).
Hence Eq. (5.28) can be written
 /E D P/E2 g1 [x/, y/, z/, , r1 , r2 . . . , r11 , r21 . . . , 1 , 2 . . .] 5.43
and for geometrical and load distribution similarity g1 [. . .] becomes a constant k1 ,
identical for model and prototype, and
 /E D k1 P/E2 . 5.44
It should be remembered that when other types of loads are also acting, their portion
of stress can be simply added to Eq. (5.44), since the principle of superposition
holds here.
In the same manner Eqs. (5.40) and (5.41) simplify to

u/ D k2 P/E2 
5.45
ε D k3 P/E2 
where also k2 and k3 are the same constants for model and prototype. Note that
since E cancels in Eq. (5.44) it confirms the well known result that stresses (or
forces, like redundant reactions) are independent of Young’s modulus E.
The preceding discussion considered problems of prescribed loads. Alternatively,
certain displacements, defined by a displacement U and the ratios of the others to
it, can be prescribed. The problem then becomes that of determining the stress ,
displacement u and strain e in terms of U, the relevant ratios (assumed to be the
same for model and prototype) and the other variables specifying the structure. In
the same manner as before, dimensional analysis yields expressions for prescribed
displacements
 /E D fN 1 [U/, x/, y/, z/, , r1 , r2 . . . , ] 

u/ D fN 2 [U/, x/, y/, z/, , r1 , r2 . . . , ] 5.46

N
ε D f3 [U/, x/, y/, z/, , r1 , r2 . . . , ]
which correspond to Eqs. (5.28), (5.40) and (5.41) for prescribed loads. When
there is no restriction to small deformations, geometrical similarity and identity of
Poisson’s ratios and (U/) values must be assured. Identity of (U/) values means
that the model must reach a geometrically similar deformed state as the prototype.
Then the stress, displacement and strain of Eqs. (5.46) are identical for model and
228 Modeling Theory and Practice

prototype. For linear structures, Eqs. (5.46) simplify, in the same manner as for
prescribed loads, to 
 /E D K1 U/ 
u/ D K2 U/ 5.47

ε D K3 U/
where K1 , K2 , K3 are the same constants for model and prototype.

5.5 Loading Beyond Proportional and Elastic Limits


For loading beyond the proportional and elastic limit, Goodier [5.9] indicated that
the one-dimensional curved stress-strain relations may be expressed in the form
 /E D fε 5.48
in a selected unit system, where, E is an appropriately chosen value for Young’s
modulus, as for example the initial slope of the curve, and fε is a function,
involving suitable numbers, that describes or approximates the measured stress-
strain curve. The function fε may be for example
 /E D ε  ˛εm 5.49
which is a three parameter representation of the experimental nonlinear stress-strain
curve, where E, ˛ and m have to be obtained by fitting the experimental curve. In
Figure 5.1 it is shown that such a representation is essentially similar to another
commonly used representation, the Ramberg Osgood method [2.77]:
ε D  /E C ˇ /En 5.50

Figure 5.1 Representation of curved (nonlinear) experimental stress-strain curves: A point A


on the experimental curve can be obtained by adding BA horizontally to the linear
stress-strain relation, i.e. ε D  /E C ˇ /En , Ramberg-Osgood [2.77], or by
subtracting CA vertically from the linear relation,  /E D ε  ˛εm , Goodier [5.15]
Buckling Experiments 229

One can note in Figure 5.1 that a typical point A on the experimental curve can
be obtained either by adding BA D ˇ /En horizontally (to the linear strain value
at stress A) according to Eq. (5.50), or by subtracting CA D ˛εm vertically (from
the linear stress value at strain A) according to Equation (5.49). Equation (5.49)
will remain valid when the unit of stress is changed, since and E will retain
the same ratio. The nonlinear stress-strain relations between all six components of
stress and the six components of strain can in a similar manner be written in terms
of a chosen E and a function fε, which consists of numbers that do not change
with change of system of units, as discussed in detail in [5.15].
With the new meaning of E, as expressed by Eq. (5.48), the conclusions
of the previous sections hold, except for the change in the definition of
E. With geometrical similarity, using the alternative formulation Eq. (5.28A),
 m 2m /Pm D  p 2p /Pp  at the same strain number (P/E2 ). For the nonlinear
stress-strain relation there is, however, no freedom of choice of a different material
for the model. The numbers appearing in the stress-strain relation Eq. (5.48) must
be the same for model and prototype. As shown in [5.15], the strict condition of
identical material may be relaxed to a requirement of model material stress-strain
relations being obtainable by an affine transformation from those of the prototype.
If parts of the structure are in the plastic regime, their behavior can still be
represented by a curved stress-strain law. Even if the loading causes the stress at a
point not to increase monotonically, but to alternate increases with decreases, the
similarity is still preserved, provided model and prototype go through the same
strain number (P/E2 ) in the same sequence.
Note that the important finding of Goodier is that, with care and some restric-
tions, similarity rules are valid also for nonlinear material behavior and in the
plastic regime. Thus their guidance for experimental studies extends also to large
and plastic deformations, as for example in postbuckling studies.

5.6 Buckling Experiments

5.6.1 Similarity Considerations for Buckling

As was pointed out in Section 5.4, curves like ( 2 /P), (u/) or ε versus the
strain number (P/E2 ) obtained from a model are valid also for the corresponding
prototype, even when the displacements are large. A buckling phenomenon would
appear as a sudden growth of (u/) at a certain value of (P/E2 ). Since the curve
is valid for all geometrically similar elastic structures (made of a material obeying
Hooke’s law), all these similar structures will buckle in the same manner at the
same strain number (P/E2 ). This can be expressed as
Pcr /E2  D C 5.51
where C is the same dimensionless number for all similar structures. Note that
buckling at a definite strain number is analogous to the onset of turbulence at a
definite Reynold’s number in a given type of fluid flow.
230 Modeling Theory and Practice

Following the arguments of Section 5.5, the conclusion of Eq. (5.51) can be
extended to buckling in the plastic regime, except that now the material has to
be the same for model and prototype. The material and hence E being the same,
both model and prototype will buckle at the same value of (P/2 ), or at the same
critical stress cr .
It may be useful to amplify these arguments, originally presented by Goodier in
[5.9] and [5.15], by a more detailed discussion (following a presentation by Chilver
in [1.12]). Considering the general problem of elastic buckling of a structure,
under a well-defined loading system, it is assumed that buckling can be defined
in some appropriate form, such as the development of gross deformations. Where
the loading is due to some external force, such as an external point load, stress or
pressure, and the critical value of this force at which buckling develops is required,
the dimensional analysis of the problem is relatively simple. It is first supposed that
the material of the structure is isotropic and homogeneous with Young’s modulus,
E, and Poisson’s ratio, . Then, for geometrically similar structures, the critical
value of the load, P
Pcr D fE, ,  5.52
where  is a typical linear dimension. Dimensional analysis then gives

Pcr /E2  D f. 5.53


Buckling is usually a structural problem at the extremes of the geometric forms, and
in many such cases, as for example slender columns or thin plates, its dependence
on Poisson’s ratio  is weak. For materials of the same Poisson’s ratio, f can
be replaced by a constant C, which reduces Eq. (5.53) to Eq. (5.51). This will also
hold approximately for the many cases of materials with dissimilar Poisson’s ratios
where buckling is not strongly dependent on .

5.6.2 Choice of Materials for Buckling Experiments

Comparing now the elastic buckling behavior of a given structure and that of a
geometrically similar model structure, not necessarily made of the same material,
but of an elastic material of the same Poisson’s ratio , Eqs. (5.53) or (5.51) yields
the simple scaling law

Pcr.m /Pcr.p  D Em /Ep m /p 2 5.54


which is the same as Eq. (5.39) derived for any static load. Hence a small-scale
model, in a relatively flexible material, can be used effectively to enable useful
experiments to be carried out at relatively low loads.
If the structures are considered in terms of stresses, the scaling rule is
 cr.m / cr.p  D Em /Ep  5.55
the same as Eq. (5.38) for any static load. Thus if model and full-size structure
are of the same material, the critical stresses are equal. Note that for complete
Buckling Experiments 231

geometrical similarity and no dependence on Poisson’s ratio,


 cr /E D constant C. 5.56
One can now write
 cr /E D  cr / y  y /E 5.57
where y is the yield stress of the material. Then for geometrical similarity
 cr / y m D  cr / y p [ y /Ep / y /Em ]. 5.58
Hence a suitable choice of material for the model can usually eliminate plastic
buckling effects in the model. The yield stress, or rather the ratio ( y /E) is therefore
of prime importance in designing a model. For example, a high strength steel model
will give, compared with a mild steel prototype,
 cr / y m D order f1/5 cr / y p g. 5.59
This means that the high strength steel model will buckle at a stress far below
yielding, whereas it might have been at or beyond yielding in the mild steel
full-scale structure. The model would therefore also permit study of postbuckling
behavior while remaining elastic.
The need to increase ( y /E) to study elastic buckling phenomena, and especially
postbuckling behavior, is reflected in the materials used for recent postbuckling
studies. For example, high-strength steel strip and sheet have been used successfully
in the study of the stability of framed structures and flat plates. Polyester films have
been used extensively to study the stability of shells, and many shell buckling
problems have been studied by using electrolytically deposited nickel or copper
shells.
Some typical examples for ( y /E) are:

1. Structural steel  y /E D 0.0013.


2. Medium strength steel (AISI 4130 drawn tubes, used for example in test spec-
imens of stringer-stiffened cylindrical shells [9.206, Volume 2])  y /E D
0.0022.
3. High strength steel (17-7 PH heat treated after hydrospinning, used for
test specimens of ring-stiffened conical shells [9.149, Volume 2])  y /E D
0.0057.
4. Aluminum alloy (7075-T6 drawn tubes used, for example, for test specimens
of stringer-stiffened cylindrical shells [13.52, Volume 2])  y /E D 0.0072.
5. Mylar polyester films (used, for example, for test specimens of cylindrical
shells [9.263, Volume 2])  y /E D 0.0210.
6. Electroformed nickel (used for test specimens of complete spherical shells
[9.90, Volume 2])  y /E D 0.0018.

The values indicate the relative suitability of the materials for models in postbuck-
ling studies.
232 Modeling Theory and Practice

5.6.3 Elasto-Plastic Buckling

Whereas the modeling of structures to simulate elastic buckling is relatively simple,


the situation becomes more complex for collapse involving inelastic effects. Instead
of the curved strain hardening stress-strain relation assumed in Section 5.5, the
material is now assumed to be sharp yielding, as structural steel, and that strain
hardening after yielding may be ignored. For complete geometric similarity, one
can write
max D fE, y ,  5.60

where max is a maximum external stress of the system, such as the average
compressive stress for collapse of a column or plate, and y is the yield stress.
Weak dependence on Poisson’s ratio is assumed and  is therefore omitted from
Eq. (5.60), though one could include it easily if desired.
In the dimensional analysis, the dimensionless products are determined formally
with Buckingham’s Pi theorem. Again only m D 2 fundamental units are required
here, force F and length L, and hence, since the number of variables in Eq. (5.60)
n D 4, n  m D 2. Choosing y and  as the m D 2 variables from Eq. (5.60), one
forms dimensionless products with the remaining n  m D 2 variables as follows:
 a   
F F  
1 D y  E D
a b
2
Lb 2 
 L c  L  . 5.61
F F  
2 D y  max D
c d
L d 
L2 L2
For 1 , 2 to be dimensionless, the exponent for each of the primary dimensions
must vanish, and this yields a D c D 1 and b D d D 0. Hence

 D f1 [E/ y ,  max / y ] 5.62

or in a more convenient form


N y /E.
 max / y  D f 5.63

Hence if full-size structure and model are of the same material, max is the same.
Furthermore, where ( y /E) is the same for the full-size structure and the model,
the values of ( max / y ) are the same. Thus when modeling in the same material is
difficult, such a suitable choice of ( y /E) leads to model results, from which max
for the full size structure can immediately be found from

max .p D max .m  y.p / y.m . 5.64

In cases where plastic collapse follows the initial development of elastic buck-
ling, some progress can be made by assuming that the elastic buckling stress, cr ,
for a perfectly-elastic material plays a role in determining the maximum stress,
max . Suppose, for complete geometric similarity,

max D f y , cr ,  5.65
Buckling Experiments 233

omitting again the dependence on Poisson’s ratio. Then, dimensional analysis


yields, in a similar manner to Eqs. (5.61) to (5.63),
N y / cr .
 max / y  D f 5.66

For columns, for example, this suggests a simple interaction curve between
( max / y ) and ( max / cr ) since Eq. (5.66) may be written as


 max / y  D fN  max / cr  y / max  . 5.67

Figure 5.2 shows the results of some pin-ended column tests by Chilver [1.12]
on a light-alloy material. Here the yield stress is taken as reasonably well-defined
by the 0.2 percent proof stress of the material. A well-defined interaction curve
emerges between  max / y  and  max / cr , where the right side represents the
region where elastic buckling predominates and the left side the shorter columns
that fail primarily by yielding.
In [1.12], Chilver also examined the collapse stresses obtained on axially
compressed plates, square tubes channel sections, I sections and other open sections
made of a range of materials and observed that Eq. (5.66) was indeed applicable
with relatively small scatter. The appropriate functional form for simply supported
plates was found to be


 max / y  D fp  y / cr 1/2 5.68

with fp nearly a linear relation of  y / cr 1/2 . In Figure 5.3 the experimental


collapse stresses of compressed channel sections of different materials (Figure 5

Figure 5.2 Pin-ended column tests on a light-alloy material ( y D 0.2 proof stress), showing
interaction curve between ( max / y ) and ( max / cr ), (from [1.12])
234 Modeling Theory and Practice

Figure 5.3 Experimental collapse stresses of axially compressed channel sections and simply
supported plates made of different materials (from [1.12]), ,  represent channel
sections, and , , žplates

of [1.12]) are superimposed on the collapse stresses of simply supported plates


of different materials (Figure 2 there) to emphasize the applicability even more.
Chilver found that for all the experimental results he examined, the collapse stress
max is indeed a function of the local elastic buckling stress cr , represented approx-
imately by Eq. (5.68).
It should be pointed out that Eqs. (5.66) and (5.68) were derived for a perfectly
elastic-plastic material, which represents structural steels fairly accurately. For
strain-hardening materials, they are only approximate, unless the restrictions of
Section 5.5 apply, i.e. or the material of model and prototype is identical, or the
stress strain curves of the model material can be obtained from those of the proto-
type material by an affine transformation.

5.6.4 Goodier and Thomson’s Experiments on Shear Panels

In order to test the validity of similarity principles for thin-walled structures


representing typical aeronautical structural elements, which buckle elastically and
plastically, Goodier and Thomson performed in 1944 at Cornell University a series
of experiments on square thin sheets in shear, with or without holes [5.15]. Square
panels of thin 2024 aluminum alloy sheet were tested in a hinged frame made of
relatively rigid angle irons (see Figure 5.4). Three sizes of frames, scaled, relative
to the smallest, in the ratios 3.20, 1.99, 1.00 (and made as nearly geometrically
Buckling Experiments 235

Figure 5.4 Goodier and Thomson’s Cornell University experiments on shear panels test
frame (from [5.15]): (a) test setup, (b) details of test frame

similar as possible), and specimens, with two sizes of central lightening holes or
without, and of five thicknesses (from 0.02000 to 0.06400 ) were made. The thin
shear panels represented a problem of large displacements extending also beyond
the elastic limit.
If the bars of the hinged frame are taken to be rigid, the variables are: the
diagonal deflections υ, the stresses , the strains ε, the external shearing load P,
236 Modeling Theory and Practice

the geometric variables of the plate, the side of the square a (see Figure 5.4), the
thickness of the sheet t and the diameter of the central hole D, and the material
constants E and , where below the elastic limit E is Young’s modulus, and if
there is plastic deformation or a curved stress-strain relation, E is a dimensional
constant as defined in Section 5.5. Dimensional analysis of the variables, in the
manner discussed earlier, then yields seven dimensionless products (since n D 9
and m D 2 here), which may be written in the form of the following relations:


υ/a D f1 P/Ea2 , t/a, D/a,  5.69


 /E D f2 P/Ea2 , t/a, D/a,  5.70


ε D f3 P/Ea2 , t/a, D/a,  5.71
Hence, if the model is of the same material as the prototype, the curves of
υ/a versus the strain number (P/Ea2 ), of ( /E) versus (P/Ea2 ) and of ε versus
(P/Ea2 ) are the same for all panels in which (t/a) and (D/a) are the same. Here,
since the strain ε and the diagonal displacement υ can be measured directly, the
test curves are (υ/a) and ε versus the strain number.
The test specimens are arranged in groups with the same (t/a) and (D/a) values,
or very nearly the same, depending on available standard sheet thicknesses. Simi-
larity is confirmed if all members of each group fall on the same dimensionless
curve, as for example one of the curves in Figure 5.5.

Figure 5.5 Goodier and Thomson’s Cornell University experiments on shear panels curves
of tensile and bending strains, as well as nondimensional diagonal and lateral
displacements versus the strain number (P/a2 E), here the ordinates, for plates
without holes of three sizes (from [5.15]). Values for the small frames are desig-
nated , for the medium sized ones ð and for the large ones . Similarity is
clearly confirmed in (a) and (c), whereas in (b) and (d) there is more scatter
Scaling of Dynamically Loaded Structures 237

The experiments were carried out by loading the hinged frame, which was
supported laterally, by means of a hydraulic jack (see Figure 5.4a). The diagonal
deflection υ was measured with a mechanical dial gage. Note that υ is independent
of any rotation of the supporting wall, unless such rotation causes bending of the
frame. The strains were measured with strain gages, which were attached, in plates
without holes, along the center line of the diagonal tension fold, that appears after
buckling. In plates with holes, the strain gages were placed at the edge of the holes
at the positions of maximum bending and maximum tension. The maximum lateral
buckling deflection w was measured at the middle of plates without holes, and
in plates with holes at the edge of the hole along a line parallel to the diagonal
tension lines (here at 42° to the horizontal).
In Goodier and Thomson’s NACA TN 933 [5.15], curves of tensile strain εt ,
bending strain εb , nondimensional diagonal displacement (υ/a) and nondimensional
lateral displacement (w/a) were plotted versus the strain number (P/Ea2 ) for all
the similarity groups. In Figure 5.5 one set of these curves, for plates without
holes D/a D 0 and t/a D ¾ 233 (actually 228, 230, 240) is reproduced from that
NACA TN. In the figures,  represents measured values for the small frames, x
values for the medium sized frames and o values for the large frames. Similarity is
clearly confirmed in Figures 5.5(a) and (c) for tensile strain and diagonal displace-
ment, while for the bending strain and lateral displacement in (b) and (d) there is
more scatter. This scatter is primarily due to errors of measurement. Indeed the
measurement of lateral deflections was pointed out to be unreliable and influenced
by rotation of frame edges. Also the bending strains were very sensitive to errors in
positioning of strain gages and their relative size. However, as a whole, similarity
was demonstrated in [5.15] by all the curves to a reasonable degree, the scatter
diminishing when the group included only two scaled frames.
Since the tests extended well into the plastic region, the results confirm the
validity of the similarity principles, not only for elastic buckling but also for
inelastic postbuckling behavior.
One notes that in buckling experiments simple dimensional analysis can be a
helpful guide in the design of meaningful experiments and that it can be extended
also to deal with yielding and collapse conditions. One of the weaknesses of
the dimensional analysis approach is that geometric imperfections, which have
a significant effect on buckling behavior, are not included in it. If a structure is
strongly imperfection sensitive, even careful experiments will demonstrate strong
scatter. The experiments themselves therefore may present an indication or warning
of imperfection sensitivity, even if theoretical consideration have not brought
it out.

5.7 Scaling of Dynamically Loaded Structures

Since dynamic loads are often the cause of buckling of structures, the similarity
and scaling conditions for time-variant problems have also to be considered.
238 Modeling Theory and Practice

5.7.1 Free Vibrations

Free vibrations of an elastic frictionless system represent a simple example of a


time-variant phenomenon. For geometrically similar systems, the frequency ω of
any specified natural mode of vibration depends on an overall scale length , and
mass density , Poisson’s ratio  and Young’s modulus E. Hence
ω D f, , E, . 5.72
In the dimensional analysis there are n D 5 variables,
p m D 3 fundamental units
and therefore two dimensionless products 1 D ω /E and , which are found
in the usual manner described in Section 5.2. Hence

f1 ω /E,  D 0 5.73
which can also be written as

1 E
ωD f2 . 5.74
 

Since the vibrations are small, the dependence on Poisson’s ratio, f2 , may be
approximated by a proportionality relation k. Then Eq. (5.74) can be replaced by

k E
ωD . 5.75
 

The resulting model law is



1 KE
K D 1, Kω D 5.76
KL K

where the K0 s are the relevant scale factors, KL D m /p , etc. The condition K D
1 may often be disregarded, as Poisson’s ratio has only a negligible effect. If
the model and the prototype are made of the same material, εP anyhow and also
KE D K D 1. Then Eq. (5.76) reduces to
K ω Ð KL D 1 5.77
yielding, for example, a double frequency for a half-size model.

5.7.2 Impact of a Rigid Body on a Structure

The impact of a moving rigid body on a structure is another simple example of


dynamic loading. The damage (like bending or fracture) at some distance from the
point of collision depends on the mass m and velocity V of the incident body, but
has been found to be practically independent of the size of this body (except for
very high speed impact ballistics). Hence one can assume the same length scale
Scaling of Dynamically Loaded Structures 239

factor for the incident body and the structure. With geometrical similarity the size of
the striking body and the impacted structure are then determined by a characteristic
length . The maximum stress at any point depends therefore on m, V, , the
mass density of the structure  and on E and . As discussed in Section 5.5, E
and  can also characterize a material with a nonlinear stress-strain behavior, but
the effects of rate of loading were not considered there. For simplicity these strain
rate effects are neglected here, i.e. E and  are assumed to be insensitive to strain
rate, but this makes the analysis inapplicable to very high-speed impacts. Hence
D f, , E, , m, V 5.78
There are n D 7 variables, m D 3 fundamental units, and therefore four dimen-
sionless products are yielded by the dimensionless analysis:

3 
1 D 
mV2 


3 
E 
2 D . 5.79
mV2 
3 
  
3 D 

m  
4 D 
These products are again obtained in the manner described in Section 5.2. One
may note that, as pointed out in Section 5.4, there are alternative groupings of the
dimensionless products, depending on the choice, in the dimensional analysis, of
the variables that contain the fundamental units. These alternative groupings can,
however, be converted from one to another by multiplication or division of the
’s by each other. Here, one alternative grouping would be (m/3 ), (E/V2 ),
( /V2 ), . Dividing the second and third product by the first yields essentially
the same grouping as in Eq. (5.79).
Equations (5.79) can be written as
 
2 3 E3 m
D mV  f1 , , . 5.80
mV2 3

If the model and the prototype are made of the same material KE D K D K D 1,
Eq. (5.80) yields the following model law:
K D 1, K D 1, Km D K3L . 5.81
This implies that a model and its prototype, geometrically similar and made of the
same material, experience the same stresses when impacted by bodies moving at
the same speed V, provided the masses of the striking bodies are proportional to
the linear dimensions of the structures cubed. Note that the model law of Eq. (5.81)
includes the three conditions given in the equation, as well as the requirements of
the same material for model and prototype KE D K D K D 1.
In this case, if the damage is determined by the stresses, the model and prototype
suffer the same overall damage. One may note that, to be precise, only the onset of
240 Modeling Theory and Practice

damage is entirely determined by the stresses, whereas its propagation depends also
on the inelastic deformation and failure mechanisms which may scale somewhat
differently. Hence the statement, though valid, may involve some approximation.
In the collision of two model vehicles or ships, for example, the models suffer the
same damage as the full scale vehicles or ships that collide at the same speed as
the models, provided models and prototypes are made from the same material, and
are geometrically similar. One should remember that complete geometric similarity
means here that the masses of the models vary as their linear dimensions cubed,
which satisfies the last requirement of Eq. (5.81).
For dissimilar materials, for example if two materials are to be used for the
model and the prototype respectively, for which KE D K D  and K D 1, the
model law would become

K D KE D K D , K D 1, KV D 1 and Km D K K3L 5.81A

Obviously, Eq. (5.81A) reduces to Eq. (5.81) for identical materials,  D 1.


As an example of the model law for dissimilar materials, of the type presented by
Eq. (5.81A), one may consider a half-scale aluminum alloy model of a steel proto-
type structure. Here Young’s modulus of the model is 0.35 that of the prototype,
Em D 0.35Ep , its mass density is also 0.35 that of the steel prototype, m D 0.35p ,
and Poisson’s ratio of both steel and aluminum alloy is approximately 0.3. The
model law Eq. (5.81A) becomes therefore for this example:

K D KE D K D 0.35
Km D K K3L D 0.35 ð 0.53 D 0.0438 . 5.81B
K D 1 and KV D 1

Hence a mass of 0.0438 that striking the prototype, which impacts the aluminum
model with the same velocity, will produce on the model a maximum stress of
0.35 that acting at the corresponding point of the steel prototype.
If the half-scale model and prototype were of the same material, model law
Eq. (5.81) would apply and then a mass of 1/8 that striking the prototype would
cause the same maximum stress on the model as that appearing at the corresponding
point of the prototype.
Strain-rate effects were neglected in the preceding elementary dimensional anal-
ysis for impact, but they are significant for strain-rate sensitive materials, like hot
rolled mild steels (see for example [5.16] [5.19]). Material strain-rate effects do not
scale properly and appear as a “size effect”, which causes laboratory models to be
stronger than the corresponding full-scale structures (see [5.19], [5.20], or [5.21]).
Other phenomena associated with the impact process, like certain non-linear
load-displacement characteristics of structures, crack propagation and dynamic
tearing, also cause deviations from the elementary geometrical scaling. Detailed
experimental studies have therefore been carried out to examine the validity of
impact scaling laws and to explore the deviations observed (for example [5.22]
to [5.24]).
Scaling of Dynamically Loaded Structures 241

5.7.3 Scale Model Testing for Impact Loading

A typical study, aimed at establishing how reliable scale model testing is for
impact loading, is the experimental investigation of Booth, Collier and Miles on
welded steel plate structures [5.23]. They carried out a series of 13 drop tests on
one-quarter scale to full-scale thin plate mild steel and stainless steel structures.
Four types of specimens were tested, two groups of square cellular (eggbox) struc-
tures, one group fabricated from mild steel plates and one from stainless steel; and
two groups of plate girders shown in Figure 5.6, both manufactured from mild
steel plates, but one mounted on a 5° inclined baseplate. Each type of structure
was made in three different scales: 1/4, 1/2 and F/S for the eggbox structures,
and 1/3, 2/3 and F/S for the plate girders. One model was repeated. Geometric
similarity was aimed at and quality assurance checks were carried out to achieve
it as far as possible. The specimens of each group were subjected to the same
impact velocity, as required for scaling, by dropping the scaled test weights (their
masses being proportional to the scale cubed, according to Eq. (5.81)) from the
same drop height. From the model law the stresses, and also the strains, will then
be the same. The deformations are the strains multiplied by the relevant length,
which is scaled. Hence the deformed shapes of the specimens should be geomet-
rically similar and proportional to the scale. However, in the tests considerable
deviations from similarity were observed, as shown for example for one group in
Figure 5.7.
For comparison of post-impact deformations, precise pre- and post-impact
measurements were carried out, the principal ones being the axial (vertical)
deformations at 10 points round the edges of the top plates of all specimens, with
additional points for the eggbox specimens. The deformations were normalized
with respect to the values measured on the full scale. The normalized vertical
deformations U N are plotted on a log scale in Figure 5.8. If linear scalability
applied, these normalized deformations should equal the geometric scale factor
, but in Figure 5.8 it is evident that the measured deformations for the models
are significantly less.
Since the impact velocity V is the same for all the tests here, and the “arresting
length” is proportional to the scale , the strain rate εP is proportional to 1/. For
the small scale model the strain rate is therefore larger, resulting in a higher flow
(yield) stress for the strain rate sensitive steel. As the energy to be absorbed (the
kinetic energy) per unit volume is the same there, the plastic deformation for the
small scale model should be less.
This strain rate effect is significant, which is also corroborated by the fact that
the results for the less strain rate sensitive stainless specimens in Figure 5.8 do not
deviate as much as the mild steel ones from linear scalability.
The overall strain rate effect, however, does not suffice to explain quantitatively
the significant deviations from linear scaling observed in Figure 5.8, as shown by
Calladine (Appendix 6.III of [5.23]) and later by Jones [5.21], who re-evaluated
the test results after an overall compensation for the actual dynamic yield stress.
The corrections were in the right direction, but insufficient.
242 Modeling Theory and Practice

(a)

Figure 5.6 Scale model testing for impact loading drop tests on mild steel plate girders by
Booth et al. at Ove Arup and Partners, London. Full scale specimen that was
tested concurrently with its 1/3 and 2/3 scale geometrically similar models (from
[5.23]): (a) Dimensions (in mm) of full scale plate girder, (b) side view of full
scale specimen
Scaling of Dynamically Loaded Structures 243

(b)

One should point out here, that intuitive plots like Figure 5.8 are of limited value,
since they represent arbitrary selections of incomplete groups of variables, which
may work if there is only one significant dimensionless group. It is preferable
to plot all the proper dimensionless groups (resulting from dimensional analysis)
against one another, then to assess their relative importance, and finally plot only
the principal ones.
In Figure 5.8 apparently other factors, in addition to the overall strain rate factor,
influence the scaling. For example, the strain rate effect may also have changed
244 Modeling Theory and Practice

Figure 5.7 Drop tests by Booth et al. on mild steel plate girders deformed shape of models
and full scale specimen dropped from the same height, producing the same height,
producing the same impact velocity (from [5.23]): (a) 1/3 scale plate girder, (b)
2/3 scale plate girder, (c) full scale plate girder
Scaling of Dynamically Loaded Structures 245

Figure 5.8 Scale model testing for impact loading: Drop tests on steel eggbox specimens and
plate girders by Booth et al. evaluation of scalability of normalized deformations
(from [5.23] with some omissions for clarity). Note that the measured deformations
for the models are significantly below the linear scalability line

the mode of deformation (see [5.25]) and thus the modal distribution of energy, or
inertia effects may be important.
Calladine and his students, [5.23], [5.26] and [5.27] emphasized the importance
of inertia effects in structures of the type tested by Booth et al. They point out
(in [5.26]) that structures can be classified into two types, type I, for which inertia
effects are insignificant, and type II, for which they are important.
246 Modeling Theory and Practice

Figure 5.9 Calladine’s classification of structures into type I and type II in quasi-static
conditions (from [5.26]): (a) Load-deflection curves, and (b) energy-deflection
curves for idealized structures, where F represents load, s deflection and U energy
absorbed

Typical load-deflection curves for the idealized type I and type II structures in
quasi-static conditions are shown in Figure 5.9(a), and the corresponding energy-
deflection curves obtained by their integration are shown in Figure 5.9(b). Type
I has a “flat-topped” load deflection curve, and a corresponding linear energy-
deflection curve, whereas type II has a “steeply falling” load-deflection curve, and
a corresponding nonlinear energy-deflection relation U / s1/2 . Laterally loaded
beams, plates and shell elements usually exhibit behavior of type I. Behavior of
type II occurs in structures that have a high initial load before buckling initi-
ates, which falls off rapidly after buckling and collapse, like columns or in-plane
compressed curved panels and shells.
The eggbox structures and plate girders tested by Booth et al [5.23] can probably
be classified as type II structures, for which transverse inertia effects are significant.
These inertia effects are related to the transverse acceleration of the structural
elements and the rapid rotation of the plastic hinges, which are more pronounced
in type II structures. Furthermore, in such structures a substantial fraction of the
incident kinetic energy will be lost in the initial impact and will not be available for
bending deformations of the structure. In [5.27] Tam, under the guidance of Calla-
dine, carried out a careful experimental study on the dynamic collapse mechanism
of typical type II structures. With the aid of high-speed photography, piezoelectric
transducers and pairs of strain gages at different locations, two primary consecutive
stages of energy dissipation were identified. In stage 1, immediately after impact,
energy was lost in inelastic collision between the falling mass and impacted struc-
ture by means of axial squashing. During stage 2, the remaining energy of the
falling mass was dissipated within the specimen by means of rotation of the plastic
hinges. The results of the experimental study supported the theoretical concept of
energy loss due to inelastic impact, proposed by Zhang and Yu [5.28] as the mech-
anism of energy dissipation in type II structures, and served as the basis for the
improved model of [5.27].
Application of dimensional analysis to the dynamic response showed that it is not
possible to maintain the equality of all the independent dimensionless groups [5.27].
Scaling of Dynamically Loaded Structures 247

In particular the different dimensionless groups due to inertia effects and strain-rate
effects cannot be maintained simultaneously. The more important dimensionless
groups for model testing have therefore to be chosen from physical considerations,
as is customary in fluid dynamics, but for tests of energy absorbing structures there
is as yet insufficient experience for a judicious choice.
The difference in scaling requirements for inertia effects and for elementary
geometric scaling may provide the missing part to the explanation of the observed
deviations from linear scalability in the tests of Booth et al. (Figure 5.8).
Calladine and English [5.26] tested experimentally two simple types of mild
steel specimens: thick walled tubes compressed between parallel plates and pairs
of joined pre-bent plates (see Figure 5.10a), whose measured load deflection
curves approximated type I and II structures respectively. The results presented
in Figure 5.10b show clearly that, whereas the quasi-static deflections (Vo D 0)
of the two types of structures are nearly equal, their dynamic deflections differ
significantly. The dynamic deflections of the type II specimens are much less than
those of the corresponding type I specimens indicating a stiffening inertia effect

Figure 5.10 Cambridge experiments on energy absorbing structures: (a) specimens before
and during deformation and method of loading for type I and type II structures,
(b) experimental results obtained by Calladine and English for the two types of
structures. The kinetic energy delivered was kept constant at 122 J (from [5.26])
248 Modeling Theory and Practice

in the dynamic behavior of type II structures. A similar but more extensive test
program was later carried out by Tam and Calladine [5.27], which also included
scaled specimens. The results of the experimental investigations in [5.27] confirm
the different dynamic behavior of the two types of structures and the scaling
difficulties inherent in type II structures.
In the test program of [5.27], the mild steel specimens were very nearly identical
to those shown in Figure 5.10. 186 specimens were tested, 36 representing type I
structures and 150 representing type II structures, of which 85 were geometrically
scaled to 1.56 of the original or approximately so, and 23 were made of aluminum
alloy. As a result of a limitation on the amount of kinetic energy that could be
delivered from the drop-hammer to the specimens, the width of the up-scaled mild
steel specimens did not scale according to the geometric scale-factor. It was shown,
however, that the mode of collapse was independent of the width of the specimens
and hence the scaling of the width could be relaxed.
To facilitate comparison of the behavior of the two types of structures and of
scale effects in the dynamic test results, a dimensional analysis was carried out in
[5.27]. In order to incorporate the load-deflection characteristics of the structure
into the formulation, the quasi-static energy deflection-curves (like Figure 5.9) were
used to interpret the final shortening of structure in the dynamic tests. The quasi-
static energy, which would be required to give the same final shape as that in the
structure tested dynamically, is defined as a new parameter SE. This parameter
accounts for the load-deflection characteristics and also, as a first approximation,
for the strain-hardening effects. The ratio (KE/SE), the ratio of the kinetic energy
used in the dynamic test to the quasi-static energy required to give the same
final deformed shape of the structure tested dynamically, emerges therefore as an
indicative parameter.
For the S1 specimens, representing type I structures the ratio (KE/SE) was
found to depend more strongly on the impact velocity Vo than on the mass ratio 
(between hammer and specimen). This suggests that indeed the strain-rate effects
must dominate the dynamic behavior of type I structures. For the S2 specimens,
representing type II structures, the ratio (KE/SE) was found in [5.27] to depend
more strongly on  than on Vo , which suggests that inertia effects indeed dominate
the deformation process in type II structures. Furthermore, for type I structures the
ratio (KE/SE) was found to increase with increasing impact velocity Vo . This
points again, in type I structures, to the important role of strain-rate effects, which
are augmented with Vo and reduce the dynamic deflections. For the same deflection,
requiring the same SE, more KE is therefore required with growing Vo , yielding
the observed increase in (KE/SE) with increasing Vo .
On the other hand, for type II structures the ratio (KE/SE) was found to increase
with decreasing mass ratio  and decrease slightly with increasing impact velocity
Vo . This behavior of the type II structures emphasizes the role of the inertia effects,
which are augmented with increase in the mass of the structure, i.e. decrease in
, and thus reduce the dynamic deflections. For the same deflections, requiring
the same SE, more KE is therefore needed for smaller m, yielding the observed
increase in (KE/SE) with decreasing  (see Figure 5.11). Also, for the same KE
Scaling of Dynamically Loaded Structures 249

Figure 5.11 Cambridge experimental study of dynamic collapse mechanism of type II


structures comparison of experimental results with theory (from [5.27]):
(a) mild steel 0.64 scale models S2a, (b) mild steel full scale prototype S2c
specimens. In the tests both mass ratio  and impact velocity Vo were varied. The
data points in the figure are experimental values of the ratio of kinetic energy to
quasi-static energy (KE/SE), while the solid lines represent (KE/T2 ) from theory
(T2 being the energy available for rotation of plastic hinges, which in the absence
of strain rate effects would be equal to SE). For clarity, only values for two typical
mass ratios  have been plotted in the figures

an increase in Vo means a smaller hammer mass G, which for the same  requires
a smaller structure mass m. Thus an increase in Vo reduces the inertia effects
(which depend on m), permitting larger dynamic deflections that would require a
larger SE to reach them. The result would be the observed decrease in (KE/SE)
with Vo for type II structures (which can also be seen in Figure 5.11).
250 Modeling Theory and Practice

The aluminum alloy type II specimens, tested in [5.27], exhibited similar trends,
of increase in (KE/SE) with decreasing  and decrease in (KE/SE) with increasing
Vo , as observed in the mild steel specimens. Since aluminum alloy is strain-rate
insensitive over a wide range of strain rates, this substantiates the role of inertia
effects in the dynamic behavior of type II structures.
For the same amount of kinetic energy delivered to S1 (type I) specimens and to
S2a (type II) specimens, even though the quasi-static deflections are approximately
equal, the dynamic deflections of S2a specimens were found (in [5.27]) to be
considerably less than those of the S1 specimens.
For the type II specimens scalability was also studied. In Figure 5.11 (which
is taken from Figure 5.10a and 5.10b of [5.27]) the ratio (KE/SE) versus Vo is
plotted for the type II S2a and S2c specimens at two typical mass ratios  D 245.3
and  D 409.0. Specimens S2a represent the models here at a scale of 0.64 of the
prototypes S2c, though, as pointed out, the width of the specimens was not scaled
geometrically. One notes that the values of (KE/SE) are larger for the model
than for the prototype, and therefore the deflections are smaller for the model. For
example, at Vo D 4 m/sec and  D 245.3,
KE/SEmodelS2a D 1.70
and
KE/SEprototypeS2c D 1.53
Since here Vo and G are identical (for an identical  and mmodel D mprototype , as
a result of the width limitation in these tests, Gmodel D Gprototype ) the KE energy
imparted is the same. Therefore the ratio of the quasi-static energy SE, required
for the same final shape as that in the dynamic tests, is
SEmodel 1.53
D D 0.90.
SEprototype 1.70
p
The quasi-static energy SE is proportional to u and hence the ratio of deflections
is for these conditions,
 
umodelS2a SEmodel 2
' D 0.81.
uprototypeS2c SEprototype
The model deflection relative to its scale ˇ D 0.64 (height or thickness) is there-
fore 0.81 ð 0.64 D 0.52 that of the prototype. If simple linear scaling applied, the
relative deflection of the model would have been 0.64 that of the prototype.
Comparison of this experimental relative deflection obtained here, in this example
from Tam’s tests [5.27], with the corresponding deflection ratio 0.475 that would
appear from the results of Booth et al [5.23] presented in Figure 5.8, shows a
similar magnitude. This indicates that the structures tested in [5.23] exhibit a similar
behavior to those of [5.27], that of type II structures, with the accompanying scaling
problems.
The experiments of Booth et al and the work by Calladine and his students point
out some of the difficulties of model testing for impact loading. It is evident that, in
Scaling of Dynamically Loaded Structures 251

spite of the recent clarification of the physical process, additional experimental and
theoretical investigations of scaling of dynamically loaded structures are warranted,
in particular of type II, since at present no comprehensive rules for scaling of such
structures can be formulated. This need has also been stressed by other authors
(see for example [5.21]), and further studies in this direction have been and are
being carried out.

5.7.4 Plates Subjected to Impulsive Normal Loading

A recent similarity study carried out on thin plates subjected to impulsive normal
loading, a commonly encountered case of dynamic loading, presents another good
example of the use of dimensionless parameters for providing prediction guidelines.
In their review of experimental investigations on impulsively loaded thin plates
Nurick and Martin [5.29] compared the data collected with the aid of dimensionless
numbers. First they tried Johnson’s dimensionless damage number [5.30]:
V2
˛D 5.82

where , V and are, as before, the material density of the plate, the impact
velocity and the maximum stress respectively ( being denoted here as d the
damage stress). Note that ˛ of Eq. (5.82) can be obtained from the dimensionless
products obtained earlier in Eq. (5.79), since from there

3 /1  D [3 /m/ 3 /mV2 ] D V2 /  D ˛. 5.82A


Comparison of the results of the plate experiments reviewed showed that ˛
predicts an order of magnitude deformation, characterizing the regime of response
behavior. Since ˛ does, however, not consider the method of impact, the target
geometry and dimensions, and the interpretation of the damage stress, considerable
variations appear, as can be seen in Figure 5.12 (reproduced from [5.29]) which
shows a plot of deflection-thickness ratios for circular plates of varying dimensions
and material properties. The abscissa in this figure is the square root of Johnson’s
damage number written in terms of impulse
I2 I20
˛0 D D 5.83
A20 t2  t2 
where I is the total impulse, A0 is the area of the plate over which the impulse is
imparted, I0 D I/A0  and t is the thickness of the plate. One may note that since
the impulse imparted I D VA0 t,
I2 V2 2 A20 t2 V2
˛0 D D D D ˛. 5.83A
A20 t2  A20 t2
In order to provide a suitable means for comparison of experimental results from
different investigations, with less variations than those in Figure 5.12, Nurick and
252 Modeling Theory and Practice

Figure 5.12 Thin circular plates subjected to impulsive loading deflection thickness ratio
versus square root of Johnson’s damage number for different plate geometries
and loading conditions (from [5.29]). The lines represent the least squares fit to
the respective data: (1) Nurick et al. [5.30], R D 50 mm, t D 1.6 mm, mild steel
uniformly loaded; (2) Wierzbicki and Florence [5.31], R D 50 mm, t D 6.3 mm,
mild steel uniformly loaded; (3) Bodner and Symonds [5.32], R D 32 mm, in (a),
(c), (d) t D 1.9 mm, mild steel, in (b) t D 2.3 mm, titanium; (a), (b) uniformly
loaded, (c) loaded R0 /R D 1/2, d loaded R0 /R D 1/3)

Martin formulated an extension to Johnson’s damage number ˛. Their modified


damage number
 D  5.84
incorporates geometry factors and . The first
D [˛0 ˇA0 /A2 ]1/2 5.85
where A is the total area of the plate and ˇ is a number defined by the geometry
of the plate (e.g. in a rectangular plate ˇ is the ratio of length to breadth, or in a
circular plate ˇ D 1).  is the relationship between the distance from the center of
the plate to the nearest boundary and the plate thickness (e.g. for a circular plate
 D R/t). A loading parameter , which accounts for the effect of partial loading
is also included in . For circular plates it is assumed to be
 D 1 C lnR/R0  5.86
where R0 is the radius of the loaded area (e.g. for uniform loading over the full
area of the plate  D 1).
A dimensionless plot of central deflection-thickness ratio versus the modified
damage number  for experimental results on circular plates is shown in Figure 5.13
(reproduced from [5.29]). The data are for steel, titanium and aluminum plates,
some with partial and some with uniform loading, from the tests of [5.31] [5.33]
Scaling of Dynamically Loaded Structures 253

Figure 5.13 Thin circular plates subjected to impulsive loading deflection thickness ratio
versus the modified damage number c , for steel, titanium and aluminum plates,
showing least squares correlation (solid line) bounded each side by a one
deflection-thickness ratio (broken lines) confidence limit (from [5.29]). Most of
the data points lie within these bounds

and some others cited in [5.29]. The experimental results in Figure 5.13 are bounded
on either side of the least square fit by a one deflection-thickness ratio confidence
limit, and most of the data points lie within these bounds. A similar good fit within
such bounds was shown in [5.29] also for rectangular plates.
The modified damage number  proposed by Nurick and Martin appears there-
fore to be a reliable dimensionless parameter for scaling of plates subjected to
impulsive normal loading.
For example, for circular plates, the geometry number ˇ D 1 by definition, and
the scalability requirement is
1/2 1/2
m D ˛0m A0 /Am m m D ˛0p A0 /Ap p p D p . 5.87
254 Modeling Theory and Practice

For geometric similarity


m D Rm /tm  D Rp /tp  D p 5.88
and for loading similarity
m D p . 5.89
Substituting ˛0 from Eq. (5.83A) into Eq. (5.87) yields, with Eqs. (5.88)
and (5.89),
m D Im / m1/2 m
1/2
A0m /A0m tm R02  D Im / m1/2 1/m
1/2
tm R02  D p . 5.90
Hence, if model and prototype are of the same material, m D p ,
 m / p  D Im /Ip 2 tp /tm 2 Rp /Rm 4 . 5.91
For a half scale mode, tp D 2tm and Rp D 2Rm , and hence
 m / p D Im /Ip 2 22 Ð 24 D Im /Ip 2 . 5.91A
An impulse of (1/8) the prototype impulse will therefore result in the same
damage stress in a half scale similar model of the same material as in the prototype
structure.

5.7.5 Response of Structures to Blast Loading

The response of structures to blast loading is another example of dynamic loading


which very often necessitates small scale model tests. Hence extensive efforts have
been devoted to derivation of appropriate model laws and their verification by
experiments (see for example [5.34] [5.45]). Baker, Westine and Dodge summa-
rized in their book [5.34] the experience and state of the art (in 1973) of scale
modeling of structures subjected to blast loading. They also presented detailed
methodologies for the application of dimensional analysis to a broad spectrum of
dynamically loaded structures.
As mentioned earlier, the modeling employed in the response of structures to
blast loading can be either “replica modeling” (geometric similarity with identical
materials) or “dissimilar material modeling” (geometric similarity with different
materials). Before scaling the response of the structure, a blast scaling law has
to be derived. Such a law is Hopkinson’s blast scaling law, first formulated in
1915 [5.36], and derived in detail in [5.34]. It implies that, if the same materials
(explosive and fluid medium through which the blast wave is transmitted) are
employed in model and prototype and geometric similarity is ensured, the pressure,
velocity, density, etc. are identical at homologous times and locations. As shown in
the derivation in Chapter 4 of [5.34], the time scales in Hopkinson’s law directly
as the length scale factor , or
KT D KL . 5.92
Hopkinson’s blast scaling has been confirmed experimentally by many investigators
over a wide range of distances and explosive source energies, as pointed out in
Scaling of Dynamically Loaded Structures 255

Chapter 4 of [5.34]. Other blast scaling laws which have been proposed are also
discussed there, with one, Sachs’ scaling law [5.37], being a generalization of
Hopkinson’s law which accounts for changes in ambient conditions, like altitude.
Sachs’ law has also been verified by many tests.
The scaling of the small-deflection elastic response of structures to blast loading
was first presented by Brown [5.38] for “replica modeling”. Following [5.35], the
modeling can be described by imagining the following experiment: An energy
source of characteristic dimension d is initiated a distance R from an elastic struc-
ture of characteristic dimension L, producing a transient pressure loading on the
structure of amplitude P and duration T, and causing the structure to respond
in its natural modes of vibration with periods 1 , 2 , . . . , n , and corresponding
displacements amplitudes X1 , X2 , . . . , Xn (see Figure 5.14). Strain-time histories

Figure 5.14 Baker’s replica scaling of response of structures to blast loading (from [5.35]):
(a) blast wave scaling, (b) scaling of response to blast wave
256 Modeling Theory and Practice

of response of the structure are characterized by the periods n and the corre-
sponding strain amplitudes en . Let the entire experiment be scaled geometrically
by a scale factor , making the energy source of characteristic dimension d and
locating the structure of characteristic dimension L at a distance R from the
source. Then, replica modeling predicts that the pressure loading on the structure
will be similar in form to that obtained in the first (full-scale) experiment, with
amplitude P and duration T; and that the structural response will also be similar in
character, with the natural periods being 1 2 , . . . , n and displacement ampli-
tudes X1 X2 , . . . , Xn , and strain amplitudes e1 , e2 , . . . , en (see Figure 5.14b).
The blast scaling is Hopkinson’s blast (scaling as shown in Figure 5.14a).
In the analysis summarized in this conjectured experiment, gravity effects in
both fluid and solid media were neglected, as well as heat conduction and viscosity
effects in fluid media and strain-rate effects in solid media.
Baker showed in [5.35] that the replica modeling, which applies to the small-
deflection response of elastic structures, describes the large-deflection response as
well, with similar neglects. He also showed there, by arguments similar to those
presented in Section 5.5, that replica modeling should also apply to elastic-plastic
response of structures to blast loading.
There has been extensive experimental verification of the replica structural
response law summarized in Figure 5.14, for small-deflection and large-deflection
elastic response, as well as for elastic-plastic response.
A typical example for elastic response modeling are the tests of Hanna, Ewing
and Baker on four geometrically scaled model steel containment shells for nuclear
reactors subjected to internal blast [5.39]. The shells were thin-walled cylindrical
shells (R/t D 240) with hemispherical ends of 1/8, 1/4, 1/2 and full scale. Geomet-
rically similar charges were detonated at homologous locations within the shells
and strains were measured with 16 strain gages located at homologous positions
and at the same orientations. The shells were half buried to assess the effect of
earth support. Peak strains for homologous locations on the shells were in general
similar, and the records obtained at scaled times showed similarity between corre-
sponding traces. The results of the study verified that the elastic response of each
shell could be predicted from measurements of the response of any other model
shell, with the aid of the replica model laws for structural response.
Another example of replica modeling of elastic response is the study by Denton
and Flathau of semicircular buried arches, of 1/3, 2/3 and full-scale, subjected to
external over-pressures [5.40]. Peak strains and deflections were measured and the
results scaled quite well.
Baker [5.35] and Ewing and Hanna [5.41] studied larger-deflection response of
slender cantilever beams to blast loading. The beams were made of 6061 aluminum
alloy and were of 1/4, 1/2 and full-scale. Small-deflection and large-deflection
elastic response, as well as permanent plastic deformation, were measured and the
results verified the predictions of the replica model laws also for large elastic and
plastic response to blast loading.
The application of replica response modeling to explosive forming, which
involves large plastic deformations, has been verified experimentally on aluminum
Scaling of Dynamically Loaded Structures 257

specimens by Ezra and his co-workers [5.42] and [5.43]. Its application to punch-
impact loading has been verified for mild- and stainless-steel plates by Duffy
et al. [5.24], by comparison of half- and full-scale tests (with maximum scaled
differences within 10 percent).
In order to allow more freedom in dynamic model testing, replica response
modeling laws have been extended to “dissimilar material modeling”, where the
“dissimilar” materials are restricted here to materials possessing constitutive simi-
larity, i.e. having identical (or nearly identical) stress-strain curves, as for example
the annealed brass and annealed aluminum in Figure 5.15 (reproduced from [5.44]).
Though dissimilar material modeling had been used before, it was first systemati-
cally studied and verified by Baker and Westine in 1969 [5.44]. The general blast
response scaling law for dissimilar material structures differs from replica scaling
in that the requirements on the model and prototype materials are now
/0 m D /0 p 5.93a
 i /Em D  i /Ep 5.93b
 i /Ci m D  i /Ci p 5.93c
E/a0 m D E/a0 p 5.93d
where,  is the density of the material, 0 the density of the ambient atmo-
sphere, i the stresses, E the elastic modulus of the material, Ci its plastic moduli
and a0 is the sound velocity in the ambient atmosphere. From Eq. (5.93a) it is
apparent that testing in identical atmospheres requires identical materials, whereas

Figure 5.15 Baker’s modeling of blast response to structures stress-strain curves for mate-
rials possessing constitutive similarity used in shock-tube tests of cylindrical shells
(from [5.35])
258 Modeling Theory and Practice

testing at a reduced or increased atmosphere requires a different model material.


Equations (5.93b) and (5.93c) are the requirements of constitutive similarity, and
if model and prototype are tested at the same temperature (yielding the same
a0 ), Eq. (5.93d) requires equality of the ratios E/ for model and prototype. This
requirement that KE D K can often be satisfied, as shown in the example of an
aluminum alloy model for a steel prototype earlier in this section (see Eq. (5.81b)).
Baker and Westine [5.44] verified the dissimilar materials blast response scaling
law experimentally for clamped-end cylindrical shells and for cantilever beams. The
cylindrical shell prototypes for air-blast loading under sea-level ambient conditions
were of heat-treated Inconal X (an iron-nickel-cobalt alloy), while the model shells
were of 6061-T6 aluminum alloy (with m D ¾ 1/3p ), and were blast loaded in a
1/3 density atmosphere, as required by Eq. (5.93a). (E/) for model and prototype
materials are practically equal, as required by Eq. (5.93d), and their stress-strain
curves are very similar, satisfying Eqs. (5.93b) and (5.93c) approximately. Another
series of clamped cylindrical shells were subjected to long-duration blast loading
in a shock tube, the test being in the quasi-static regime, where the model law
only dictates geometric similarity and constitutive similarity for the materials. In
these tests the prototypes were made of annealed aluminum alloy and the 1/3
scale models of annealed brass, the two materials having very similar stress-strain
curves indeed (see Figure 5.15). The correlation between model and prototype
structural response was generally good for the cylindrical shells, and even better
for the cantilever and pin-ended beams tested in a similar manner (see [5.44]),
except for the comparison between steel prototype beams and lead-plastic models
(for which constitutive similarity was not preserved, on account of a steep strain-
hardening of the lead-plastic). For the cylindrical shells there is significant scatter
in the results due to the failures being caused by buckling, for which in cylindrical
shells scatter is usual, on account of their sensitivity to initial imperfections. For
the quasi-static regime experiments, the internally trapped air was not properly
scaled and the internal pressure in the aluminum prototype was too high, which
stiffened their postbuckling behavior and reduced their inelastic response. But, in
spite of the scatter, the results of all the experimental investigations carried out
by Baker and Westine, demonstrated clearly that elastic-plastic structural response
modeling with dissimilar materials is feasible and permits a significant broadening
of response modeling techniques.
One may note that scale modeling of structures subjected to impulsive normal
loading or to blast loading appears to yield fairly reliable results, which were
verified by a considerable spectrum of experiments. This seems at first surprising
in view of the difficulties in scaling of dynamically loaded structures experienced by
Booth et al. [5.23] and by Calladine and his students, [5.26] and [5.27]. But a closer
study of the experimental investigations summarized in [5.29] and [5.34], shows
that these tests deal with type I structures, according to Calladine’s classification,
whose behavior agrees well with the scaling laws, whereas the difficulties and
significant discrepancies arise in type II structures.
An additional verification of the scalability of type I mild steel structures
subjected to dynamic lateral loads, was presented by two series of preliminary
Scaling of Composite Structures 259

scaled experiments carried out by Donelan and Dowling [5.45] in preparation for
drop tests on Magnox fuel transport flasks. Models of 1/3, 1/2 and full scale were
employed and the dynamic behavior could indeed be obtained from the models.
One can therefore conclude that scaling under dynamic loading of type I struc-
tures can be employed with confidence, whereas for type II structures no compre-
hensive scaling rules can yet be formulated and further studies are needed to
develop such reliable rules. It is therefore important to study the quasi-static
load-deflection behavior of the structure to be scaled, before one embarks on the
application of geometrical scaling for dynamic loading. Furthermore, as pointed
out by Jones in [5.21] and [5.46], this conclusion applies only to ductile dynamic
response whereas geometrical scaling does not apply when tearing, cutting or
ductile-brittle transitions occur during a structural response.

5.8 Scaling of Composite Structures

5.8.1 Problems in Scaling of Laminated Composites

Laminated fiber composite structures are widely used in modern aerospace, marine
and automotive vehicles, on account of their high strength- and stiffness-to-weight
ratios. Since no design data base, comparable to that available for steel and
aluminum structures, has yet been assembled, the designers have to resort to exten-
sive testing. Full scale prototype testing is, however, very expensive, especially
with advanced composites, which has motivated scale-model testing and develop-
ment of scaling rules, aimed at making the interpretation of the model tests more
reliable.
Advanced fiber reinforced composite materials consist of thin, stiff, strong fibers
(for example, carbon, boron, glass or aramid fibers), with diameters of about
6 20 µm (0.00015 0.0005 in.) embedded in a comparatively low performance
matrix. The usual matrix materials are thermosetting resins (primarily epoxy resins)
or thermoplastic resins (like polyetheretherketone PEEK or polyethersulphone
PES). The material is built up of laminae, each having a thickness of about 0.13 mm
(0.005 in.) and a specified direction of the fibers. This lamina thickness has become
a standard for Graphite Epoxy composites and many others, so that the thickness
of laminates is usually specified only in terms of the number of laminae in the
lay-up.
For similarity, scaling of composite structures should ideally include also scaling
of the microstructure of the material, namely, laminae thickness and fiber diameter
should also be scaled. This is practically impossible, as pointed out by Morton
[5.47] who studied the scaling of laminated fiber composites very extensively.
When damage occurs, the inability to scale the laminae and fibers presents a serious
limitation on the modeling of laminated composites, since there are frequent inter-
actions between microstructural and macrostructural properties. But for overall
structural response, prior to substantial damage, scaling rules that do not scale
the microstructure suffice, even in the case of impact loading, as was shown by
260 Modeling Theory and Practice

experiments on laminated composite beams and on laminated composite plates


([5.47] [5.51] and [5.53], [5.54], respectively).

5.8.2 Scaling Rules for Laminated Beams and Plates

Morton derived such scaling rules for laminated beams subjected to transverse
impact loading using dimensional analysis [5.47], and similar scaling rules were
derived by Qian and Swanson [5.53] from the differential equations governing the
impact response of transversely impacted orthotropic laminated plates.
These rules show that if the geometry of the beams or plates is scaled as 
(assuming that the lay-up can be scaled, which in practice is only rarely feasible),
the maximum strain in the beams or plates is constant with scaling if the impact
velocity V is unchanged. Also the contact force scales as 2 , and if the contact area
scales geometrically the contact pressure (or stress) is unchanged. If the impactor
also scales geometrically, the impact mass scales as 3 . The time of response will
also be scaled, the time to maximum load and strain scaling as . It is assumed that
the same material is being used in model and prototype. Note that the events in a
smaller model ( < 1) will therefore appear to occur faster than in the prototype.
For example, a 1/5 scale model will reach its maximum strain in 1/5 the time it
takes the prototype to do so.
In terms of scale factors, the model law is, as in Eq. (5.81),
K D 1, KV D 1, Km D K3L
with KT D KL and Kε D 1, 5.81A

where KL D  and T represents the time of response and ε the strain, and with the
requirements for the same material now being KEij D Kij D K D 1, where Eij
and ij are the equivalent directional elastic properties of the laminate (assuming
appropriate scaling of the lay-ups).
The scaling laws are essentially the same as those derived in Section 5.7 for the
impact of a rigid body on a structure, summarized in Eq. (5.81), or as the replica
structural response law for blast loading, summarized in Figure 5.14b. Hence, prior
to significant damage, the model laws for the overall structural response of lami-
nated fiber composite structures are unaffected by the nature of their microstructure.
This conclusion for dynamically loaded structures will certainly hold also for stat-
ically loaded ones (assuming again appropriate lay-up scaling).

5.8.3 Scaling for Strength and Large Deflections of Composites

When substantial damage occurs, matters are more complicated, since the failure
mechanisms of laminated composites are not yet well understood. As pointed out
by Morton [5.47], different damage mechanisms may appear, including fiber frac-
ture, delamination and matrix cracking, and there are frequent interactions between
micro- and macrostructural properties. All damage modes start on a microscopic
Scaling of Composite Structures 261

scale and eventually interact with the macroscopic scale on a laminate level. For
example, a matrix microcrack may grow across a lamina until it reaches the inter-
laminar boundary, and then either delamination or fiber fracture occurs, or both.
Furthermore, the notch-sensitivity of laminated composites, depends not only
on the sensitivity of each lamina, but also on the lay-up of the laminae. Also
the rate-sensitivity depends on the fibers (for example, glass and aramid fibers
are highly rate-sensitive, whereas carbon fibers are rate-insensitive), on the matrix
(thermoplastics being more rate-sensitive than the epoxy resins), and on the lay-up
of the laminae.
Hence the scale modeling of laminated composite structures when substantial
damage occurs is complex, the choice of scaling parameters is difficult and exten-
sive testing is required to establish guidelines for this choice. As yet, few tests have
been performed and insufficient information is available for appropriate scaling
guidelines.
Morton [5.47] tested a series of scaled laminated composite beams, supported
on rollers and impacted centrally by a free falling mass. The specimens were
fabricated from unidirectional carbon/epoxy (AS4/3502) prepreg in four types
of lay-ups: A 902C2 ; B C45C1 , 45C1 s ; C 90C1 , 0C1 s ; and D quasi-
isotropic 45 , 90 , 45 , 0 s . The scaled laminates were produced with  D 1, 2
and 3 (8-, 12- and 16-ply), except the quasi-isotropic lay-up, which was only made
with  D 1, 2 (8- and 16-ply). The beam dimensions and impactor shape and mass
were scaled approximately according to the scaling laws (5.81A). The use of the
same prepreg material and forming the lay-ups by scaling as indicated, with suffi-
cient number of plies for each orientation, ensured the similarity of elastic behavior
in all the scale models.
From recorded strain output traces (strain versus time), the duration of impact
and the maximum strain were obtained as output parameters. The duration of
impact should scale with the scale factor , and therefore the measured durations
divided by  should be constant. The results satisfied this requirement within š10
percent as is shown for example in Figures 5.16(a) (d), reproduced from [5.47].
The differences in the elastic parts of Figures 5.16(a) (c) is attributed primarily to
departure from scaling in the thickness of the beams, since the molded thicknesses
of the laminates did not scale exactly, though the number of plies did. For the
quasi-isotropic laminates (lay-up D), on the other hand, the scaling of the molded
thickness was nearly exact and therefore the elastic parts of Figure 5.16(d) practi-
cally coincide, as expected. An elastic analysis in [5.47] predicts that the impact
durations should be independent of the impact velocity, which is confirmed by
the test results in Figure 5.16. Deviations from the behavior indicates the initia-
tion of impact damage, which invalidates the scaling laws and the elastic analysis
employed in predicting the behavior.
When damage occurs significant size effects appear. Figure 5.16 clearly shows
that damage occurs at higher impact velocities in the smaller specimens than in
the corresponding larger ones. In general, small scale models are observed to
be stronger than their respective prototypes and to carry proportionally higher
post-damage loads. This is probably due to macrostructural fracture effects, which
262
Modeling
Theory and Practice

Figure 5.16 Scaled normalized impact durations for composite beams at various normalized impact velocities (from [5.47]): (a) lay-up A, (b) lay-up
B, (c) lay-up C, (d) lay-up D
Scaling of Composite Structures 263

tentatively indicate that the absolute size of matrix cracks rather than their scaled
size are important in laminated composites, and hence are more detrimental in the
large prototype than in its model.
Similar agreement with scaling law predictions was found for the normalized
impact force in the elastic part, and similar deviations after impact damage occurs
were observed, with small scale models being consistently stronger.
In similar studies on the response of Graphite Epoxy beam columns by Jackson
and Fasanella [5.48] [5.51] the emphasis was on large deflection response. First
the scaling effects in static large deflection response was studied and then the
scaling in similar dynamic responses was investigated. The scaled beams were
loaded in a beam-column fashion by an eccentric axial load (see Figure 5.17).
This structural configuration, though simple, possessed such interesting features as
large deflections, combined tensile and compressive loading, and global failures.
The beams were made of a high modulus graphite fiber and an epoxy matrix
system designated as AS4/3502, in four different laminate stacking sequences:
unidirectional, angle ply, cross ply and quasi-isotropic. The full scale beam was
3 in. (7.62 cm) wide, with a 30 in. (76.2 cm) gage length and 48 plies thick, with
an average ply thickness of 0.0054 in. (0.137 mm).
For the static tests, the scale model beams were constructed by applying seven
different geometric scale factors, 1/6, 1/4, 1/3, 1/2, 2/3, 3/4 and 5/6 to the full

Figure 5.17 NASA Langley experiments on the scaling of the response of Graphite-Epoxy
beams test setup (from [5.48]): (a) schematic drawing of the flexural test
configuration, (b) details of the scaled hinge-beam attachment
264 Modeling Theory and Practice

scale beam dimensions (see Figure 5.18). The thickness dimension was scaled
by reducing the number of layers in each angular ply group of the full scale
laminate stacking sequence, which consisted of at least six plies of similar orien-
tation. Using this approach, it was not possible to make a 1/2 or 3/4 scale quasi-
isotropic beam.

Figure 5.18 NASA Langley scaling experiments on Graphite-Epoxy beams statically tested
scaled beams, 1/6 scale to full scale (from [5.50]): (a) failed unidirectional beams,
(b) failed quasi-isotropic beams
Scaling of Composite Structures 265

It should be noted that ideally, in true replica models of the prototypes, the
microstructure should also be scaled. This would involve scaling of the individual
lamina thicknesses and fiber diameters for each scale model, which is not practical.
Scaling by reduction of number of layers was therefore used as an approximation.
The beams were machined from panels which were hand layed-up from prepreg
tape and cured according to manufacturer’s specifications. Slight variations were
observed in the thickness dimensions of the cured specimens, the maximum devi-
ation in normalized thickness being 6 percent. For each laminate type and size of
beam, three replicate tests were carried out.
For the impact tests, 1/2, 2/3, 3/4, 5/6 and full scale beams were fabricated,
and the specimens with the scaled eccentric hinges of Figure 5.19 were placed and
loaded in a drop-tower (described in detail in [5.49] and [5.50]).
For the statically loaded beams, the normalized load versus end displacement
plots (the vertical load was normalized by the corresponding Euler buckling load

Figure 5.19 NASA Langley scaling experiments on Graphite-Epoxy beams dynamically


tested 1/2 scale and full scale beams (from [5.49] or [5.50]): (a) dynamically
failed unidirectional beams, (b) dynamically failed quasi-isotropic beams
266 Modeling Theory and Practice

Figure 5.20 NASA Langley scaling experiments on Graphite-Epoxy beams experimental


normalized load versus end displacement for statically loaded scaled beams, 1/6
scale to full scale (from [5.50]): (a) unidirectional beams, (b) quasi-isotropic beams

and the end displacement by the gage length) were plotted for the four lay-ups. The
curves for the unidirectional and quasi-isotropic beams, shown in Figure 5.20, and
those for cross ply beams and angle ply specimens (shown and discussed in [5.48]
and [5.50]) which are roughly similar, indicate that the response scales well at small
end displacements for all lay-ups. The unidirectional (Figure 5.20a) and cross ply
beam responses scaled as predicted by the model law even at large displacements,
Scaling of Composite Structures 267

whereas for angle ply and quasi-isotropic beams (Figure 5.20b) some deviation
from scaled response was observed, due to damage initiation.
Jackson’s experiments therefore verify and amplify the earlier conclusions of
Morton [5.47] about the scalability of the elastic or small deflection structural
response of composite structural elements. Also in the dynamic tests the initial
response scaled adequately for all the lay-ups. For the unidirectional beams the load
and strain responses also scaled according to the model law at large deflections
(see for example Figure 5.21a), but for angle ply, cross ply and quasi-isotropic
beams scaling of load and strain histories, at large deflections, was found to be
inconsistent (see for example Figure 5.21b). The deviations from scaled response
at large deflections were observed to depend on the laminate stacking sequence.
The models (in [5.48] [5.50]) were tested until failure. In general, failure modes
were consistent between scale models within a laminate family, both for static
loading (see, for example, Figures 5.18(a) and (b)) and for dynamic loading (see
Figures 5.19(a) and (b)). However, a significant scale effect was observed in
strength, as can also be clearly seen in the normalized load versus end displacement
curves (Figures 5.20(a) and (b)), where the smaller scale model beams failed at

Figure 5.21 NASA Langley scaling experiments on Graphite-Epoxy beams midpoint strain
versus scaled time plots for scaled dynamically loaded beams, 1/2 scale and full
scale (from [5.50]): (a) unidirectional beams, (b) quasi-isotropic beams
268 Modeling Theory and Practice

higher normalized loads and much higher normalized end displacements than their
full scale prototypes. Since the usual failure theories for composites cannot predict
this scale effect, there appears to be a scale effect in the failure behavior, as was
also pointed out by Morton [5.47]. A similar scale effect in strength was also
observed in the dynamic tests, [5.49] and [5.50].
This strength scale effect was also studied by Kellas and Morton in an exten-
sive series of tensile tests on replica model bars [5.52]. The specimens were all
of AS4/3502 Graphite-Epoxy, in four different lay-ups and four different scaled
sizes, 1/4, 1/2, 3/4 and full-scale. The full scale specimen was 32 plies thick, and
the thickness was again scaled by reducing the number of layers in each angle
ply group. As in the bending experiments, also in the tensile tests, the stiffness at
small strains was independent of specimen size. There was, however, also here a
significant scale effect in tensile strength, the small scale models failing at signif-
icantly higher normalized loads than their prototype. In Figure 5.22 (from [5.51])
this “strength scale effect” is summarized for both tension and bending, showing
its magnitude and dependence on the lay-up of the composite structure.

5.8.4 Scaling of Composite Plates

Another series of experiments, aimed at evaluating the scaling laws, on scaled


laminated carbon/epoxy (AS4/3501-6) plates impacted laterally by cylindrical
projectiles, were carried out by Qian et al. [5.53] and [5.54]. Five sizes of
geometrically scaled square plates with  D 1, 2, 3, 4 and 5, from 50 mm ð 50 mm
by 1.072 mm thick to 250 mm ð 250 mm by 5.36 mm thick (from 2 by 2 by
0.042 inch to 10 by 10 by 0.211 inch), were tested. The corresponding lay-up was
[š72 , 02 ]s with  the geometric scaling factor and the specimens being 8-, 16-,
24-, 32- and 40 ply.
The plates were clamped on two opposing edges and free on the other two
edges. The clamped edges were normal to the 0 degree fibers, whose direction is
designated as the X direction. The projectiles were shot from a horizontal air gun
at the plates in a vertical position. The impactors were also scaled geometrically,
requiring five different barrels, their mass varying as the cube of the scale factor ,
and at each scale three different tip configurations were tested. All plate sizes were
tested at four impact velocities, V D 4.57, 12.2, 18.3 or 24.4 m/s (15, 40, 60 or
80 ft/s). For each condition, four specimens were tested, one of them instrumented
with strain gauges (the strain gauge sizes were scaled geometrically with the plates).
An extensive series of tests indeed.
The strain response (in the Y direction, parallel to the clamped edges) is shown
for three sizes of specimens in Figure 5.23 (reproduced from [5.53]), all for the
same impact velocity of 4.57 m/sec (15 ft/sec), which is believed to be below
the threshold for damage formation. The time has been divided here by the scale
factor  to show how the strain traces nearly coincide, since according to the
scaling laws the time scales as . The results in Figure 5.23 show indeed that
the dynamic response scale is in close agreement with the scaling rules derived.
Similar agreement is found for other strain traces, and the strain predicted from the
Scaling of Composite Structures 269

Figure 5.22 Jackson and Morton’s summary of strength scale effects for Graphite-Epoxy
beams (from [5.51]): (a) normalized strength versus specimen size for four
laminates loaded in tension, (b) normalized failure load versus scale factor for
unidirectional, angle ply, cross ply and quasi-isotropic beams subjected to flexural
loading

dynamic plate analysis also compares quite well with experiment. The experiments
therefore verify the scaling laws for overall structural response during impact at
moderate damage levels.
The experiments were designed so that significant damage would be developed
in the plates impacted at the higher velocities. The damage then took the form of
contact point indentation, matrix cracking, broken fibers and delamination. With
damage, the scaling laws will become more complicated than those derived for the
linear structural response. This is apparent in Figure 5.24 (reproduced from [5.54])
presenting the delamination areas determined by C-scan for three specimen sizes
at a constant impact velocity of 12.2 m/s (40 ft/s). If the size of the delaminations
270 Modeling Theory and Practice

Figure 5.23 Impact tests of five scaled composite plates comparison of strain response
behind impact point, showing time scaling, for the same impact velocity of
4.57 m/s (from [5.48])

Figure 5.24 Scaling effects in the impact response of composite plates increase in delami-
nation area with specimen size for centrally impacted plates at a constant impact
velocity of 12.2 m/s (from [5.49])

were governed by the simple geometric scaling, the delamination area would scale
as 2 . The measured delaminations are however significantly larger for the larger
specimens, a size effect which is apparently consistent with fracture mechanics.

5.8.5 Scaling of Composite Cylindrical Shells

Recently Swanson, Smith and Qian extended the studies to the response of cylin-
drical filament wound carbon/epoxy (IM7/55A) cylindrical shells impacted by
cylindrical projectiles [5.55]. Two sizes of cylindrical shells were designed so
that all geometric parameters were scaled by a factor of approximately 3.3. The
Scaling of Composite Structures 271

scaling included the thickness of the ply groups, the cylinder diameters and wall
thicknesses, and the sizes of the projectiles. The experimental set-up was similar
to that employed in the earlier tests, [5.53] and [5.54].
Scaling rules, developed by Christoforou [5.56] from the differential equations
governing the impact response of transversely impacted laminated cylindrical shells,
were employed for the scaling of the impact experiments. Again also for cylindrical
shells, if their geometry is scaled as  (assuming that the lay-up too is scaled) the
strain is constant if the impact velocity is unchanged. The contact force scales then
as 2 , the time of impact duration scales as  and the impact mass scales as 3
(the projectiles are scaled geometrically). All as in beams and plates.
Though the detail scaling of the lay-up was not complete (the basic fiber diameter
and number of fibers per windings were not scaled), these scaling rules proved to
be quite accurate for impact velocities below the damage threshold. An example
is shown in Figure 5.25, where typical strain responses of the small and large
cylinders appear to compare rather well. As in the beam and plate test results,
the time scale in this figure is divided by the geometric scale factor , according
to the scaling rules. These studies on cylindrical shells were limited to structural
response below the level of damage formation.
It should be pointed out that to date all the experiments verifying the scala-
bility of dynamically loaded laminated composite structures deal only with type I
structures, according to Calladine’s classification, and the reservations regarding
the scalability of type II structures, discussed in the previous section, apply here
as well.

Figure 5.25 Scaling of composite cylindrical shells typical comparison of strain gauge
response between small and large cylinders, illustrating scaling of strain and time
scale (from [5.55])
272 Modeling Theory and Practice

It can be concluded that the buckling of composite structures, which is essentially


a structural response, can be scaled reliably, whereas for postbuckling behavior,
which may involve significant damage, scaling requires great caution, and addi-
tional experimental studies are warranted to develop appropriate guidelines.
Some recent studies on the impact behavior of quarter-scale of model composite
sailplane fuselage segments [5.57] indicate that there is room for optimism. Since
qualitative comparisons with field observations of actual crash damage in composite
sailplanes, with that in the model impact tests simulating typical nose-down crashes,
showed that the model failed in the same failure mode as the full-scale fuselages
and at appropriately scaled loads.

5.9 Model Analysis in Structural Engineering


5.9.1 Model Analysis as a Design Tool

Structural modeling has been used extensively, primarily in civil engineering, as an


experimental method to supplement and even replace analysis. Model analysis of
structures, as an alternative to theoretical analysis, initiated in the first decades of
the 20th century and reached maturity and widespread use in the second half of the
century (in Germany for example, model-analysis called there Modellstatik is
considered a special discipline, with a special chair and institute in at least one
university). Model analysis employs measurements on a scale model to determine
the stresses, deformations, strength and failure modes of the prototype, whereas
theoretical analysis uses an imaginary mathematical model for the same predic-
tions. The proponents of model analysis claim (see for example [5.2]) that their
model simulates the real structure more realistically (in particular with regard
to material behavior, boundary conditions, loading conditions and possibly also
imperfections) than the idealized mathematical model of the analysts, the idealiza-
tions being imposed by theoretical limitations and the extent of the computational
efforts. With the rapid development of more sophisticated computational methods
and faster computers, the theoretical simulations have recently improved signifi-
cantly, but in many cases the economics are still in favor of model analysis as a
design tool. Hence, though modern computer aided design is slowly conquering the
field, model analysis remains a viable tool, especially for new structural concepts
and materials, whose behavior is not yet well known.
Model analysis used to be divided into two basic methods (as in [5.3]): (a) the
indirect method, which determines the influence lines of frameworks, and (b) the
direct method, that measures the stresses and deformations of the structure. The
indirect method with its simple celluloid models, which was widely used up to the
sixties is now only of historic interest, since influence lines are today determined
much easier and faster with a computer. The direct method, on the other hand,
continues to serve as a valuable design tool.
For buckling problems, model experiments are mostly used to study behavior
of new structural concepts and to verify theoretical and numerical methods on
Model Analysis in Structural Engineering 273

relatively simple structural elements under well defined conditions. Such exper-
iments cannot precisely be classified as model analysis proper. Combinations of
theoretical and numerical methods with series of increasingly more realistic model
studies, which can be considered an “extended model analysis”, are however often
employed in the design process of civil, marine and aerospace structures, where
buckling is the governing parameter. Some examples will demonstrate this.

5.9.2 Model Analysis in Vibration Studies

One illustrative example of the use of model analysis, not for a buckling problem
but for a related one, are the vibration studies on various machine structures and
their supporting elements carried out on plastic models by Wright and Bannister at
the Westinghouse Research Laboratories in 1970, [5.58] and [5.59]. They strongly
advocate the use of plastic models, made from plexiglas (acrylic resins) or Tenite
II (cellulose acetate butyrate) for analysis and design improvement of complicated
structures on economic grounds, stating that “for complex structures, a model is
often the cheapest computer one can buy”. This statement may not be entirely
accepted today, but their next one, “a model study gives good physical under-
standing of the behavior of the entire structure” is certainly still very valid and
appropriate.
Wright and Bannister noted that dissimilar materials models are valuable for
vibration studies on complex structures. They showed that plastic models have
several advantages for static and vibration tests: (1) deflections under given applied
loads are large and easily measured, whereas the required driving forces are small;
(2) model natural frequencies are relatively low, allowing for the use of small-
models measuring equipment with limited upper frequency response; (3) model
cost is low; (4) structural modifications can easily be made; and (5) since the
required impedances of the model are much smaller, high-impedance foundations
are easily provided.
Using dimensional analysis, they generated in [5.59] a replica model law for
elastic vibrations of complex structures for dissimilar materials. For the simplest
case, of a freely suspended structure with solid joints, there are seven indepen-
dent variables: L a characteristic length, c the velocity of sound in the material,
 its mass density,  its Poisson ratio,  the damping loss factor (representing
uniformly distributed material damping of the complex-modulus type), Fi the sinu-
soidal driving force, and ω its circular frequency. Note that the longitudinal speed
of sound c is an alternative
p to Young’s modulus E, for specification of the elastic
properties, since c D E/. Also introduction of the damping loss factor  brings
in material damping, which is an inelastic effect, into elastic vibrations.
It should be remembered that the velocity of sound is the material property of
prime importance in natural frequency tests. Hence it must be measured at the
ambient temperature of the model test at all frequencies of interest. In [5.58] the
velocity of sound c and the damping loss factor  were determined by measuring the
resonant and antiresonant frequencies of free-free columns and beams, machined
from the actual batch of plastic used to construct the model.
274 Modeling Theory and Practice

Since there are three fundamental dimensions, there will be four Buckingham
pi terms: 
1 D ωL/c 

2 D Fi /c2 L 2  
. 5.94
3 D  


4 D 
Satisfaction of the pi terms of Eq. (5.94) (as a matter of fact usually only the
first two of them exactly, as will be discussed below), and some other physical
relationships (like Hooke’s law), yields similarity conditions for the model vibration
tests (only the important ones are listed):

Kω D Kc /KL for frequency 


Kε D 1 for strain 



K D K Kc 2
for stress 
5.95
KV D Kc for velocity  


KF D K K2c K2L for force 



KK D K K2c KL for stiffness
where the K0 s are the scale factors,

Kω D ωm /ωp 
Kc D cm /cp etc. 5.96

K D m /p
For example, for a 1/6 scale model of a steel prototype (KL D 1/6), the scale
factors for a steel model and plexiglas or Tenite II models would be as shown in
Table 5.1.
It was found difficult to satisfy also 3 and 4 . Poisson’s ratio of plexiglas, for
example, is about 0.38 versus  D 0.28 for steel, a 35 percent difference, but resulting
only in an error of about 4 percent in the natural frequencies, which can be corrected
if the structure is not too complex. On the other hand, the damping loss factor
for the plastic material is at least an order of magnitude higher than that of steel,
which makes plastic models inaccurate in scaling response amplitudes near resonance
peaks. Frequencies and mode shapes are, however, modeled very accurately.

Table 5.1: Scale factors for 1/6 scale models of a


steel prototype
Scale Factor Steel Plexiglass Tenite II
Kc 1 (1/2.50) (1/4.21)
Kω 6 2.40 1.42
KV 1 (1/2.50) (1/4.21)
K 1 (1/6.70) (1/6.70)
Kε 1 1 1
K 1 (1/41.9) (1/119)
KF (1/36) (1/1510) (1/4290)
Model Analysis in Structural Engineering 275

Figure 5.26 Comparison of scaled vibration amplitudes in a submarine propulsion unit and its
Plexiglas model (from [5.59])

Wright and Bannister also extended the dimensional analysis to structures with
joint friction, with impedance terminations, suspended in an inviscid fluid, and
located in an incompressible fluid (see [5.59]). For these more complicated cases,
some of the new pi terms may also be difficult to satisfy. For example for joint
friction, the coefficient of friction  is different for steel and plastic contacts. But
sometimes judicious matching may overcome such problems. For example, the
material damping of plexiglas is approximately equal to the joint damping of some
large bolted and welded steel machinery structure at low frequencies.
Hence in spite of inaccuracies in some pi terms (which can be assessed), model-
prototype comparisons show often good agreement, as for example the scaled
vibration amplitudes in the submarine propulsion unit and its plexiglas model in
Figure 5.26 (reproduced from [5.59]).

5.9.3 Buckling Experiments on Models of a Composite Ship


Hull Structure

Another example is the buckling experiments carried out at the U.K. Naval
Construction Research Establishment in the early seventies on four small-scale
plastic models, representing the bottom structure of a prototype glass fiber-
reinforced plastic minesweeper ([5.60]). During the design of the hull structure,
serious problems of elastic instability were encountered, which arose mainly as a
result of the high strength, low stiffness characteristics of the hull material (whose
tensile and compressive strength equalled the yield strength of mild steel, but
276 Modeling Theory and Practice

whose E was only seven percent that of steel). The need for high strength under
explosive loads and fabrication considerations led to the adoption of a transversely
framed hull, in which buckling could be expected to cause catastrophic failure, with
practically no postbuckling reserve of strength. Theoretical studies [5.61] predicted
that, under longitudinal compression, failure of transversely framed bottom and
deck panels would occur by local instability. Four forms of local instability
were indicated for the panels stiffened by transverse top-hat frames, shown in
Figure 5.27. Types-3 and -4 forms of instability had been overlooked by previous

Figure 5.27 Smith’s buckling experiments on models of a composite ship hull


structure forms of local instability in panels with transverse top-hat frames (from
[5.60])
Model Analysis in Structural Engineering 277

design methods, and type-3 buckling was now predicted to occur at significantly
lower stresses than in the other modes. The model experiments were therefore
carried out primarily to demonstrate the existence of the predicted critical type-3
instability, but also to verify the predicted buckling stresses and check the validity
of the assumed boundary conditions.
The structural modeling was carried out in two stages: First, compression tests
on two small-scale flat rectangular panels with transverse top-hat frames (see
Figure 5.28a), and secondly tests on two further models, one representing the

Figure 5.28 Smith’s buckling experiments on models of a composite ship hull structure test
rig for Perspex models 1 and 2 of a transversely stiffened panel under longitudinal
compression (from [5.60]): (a) the transversely stiffened panel (schematic), (b) the
test rig
278 Modeling Theory and Practice

complete vee-bottom structure of the ship and the other representing a full 3-
dimensional ship compartment, for evaluation of the assumed boundary conditions.
The models were all made of Perspex (polymethylmethacrylate), whose high
strength to stiffness ratio1 (about 5 8 times that of aluminum alloy) allows elastic
instability to develop in many practical structural forms long before material failure
occurs.
The 1524 mm ð 622 mm and 1524 mm ð 533 mm Perspex rectangular panels
were supported in the test rig (Figure 5.28b) at their end and sides by steel tie-bars,
having bottle-screws to allow vertical adjustment. These tie-bars were pin-jointed
at one end to a heavy steel reaction frame and at the other end to the edge of
the test panel. Vertical displacement was thus restrained at the edges of the panel,
with negligible restriction of in-plane displacements and edge rotation, closely
approaching classical simple supports. Similar supports were employed recently by
Minguez [8.61] for the unloaded edges of his plate tests discussed in Chapter 8,
the tie-bars being replaced there by longer tensioned steel wires (see Figures 8.37
and 8.38) to ensure an even better approach to simple supports. The loaded ends
of the panels were reinforced here by steel sandwich plates, which distributed the
concentrated jack loads uniformly to the panel, but no doubt restricted the in-
plane displacements and rotation of the loaded edges. Since, however, the critical
buckling form was a many wave local instability, the reinforcements only reduced
the risk of premature failure at the ends and had negligible influence on the local
instability of the panels away from the two edge bays. Strain gage measurements
on the panels verified that the tie-bar supports indeed ensured negligible load loss
to the test rig (less and 2 percent).
The deflection profiles were measured along the centerline of each panel at
selected load steps and well defined buckling patterns could be discerned. The
type-3 instability, which was overlooked by conventional design methods, but
was predicted by the studies of [5.61] to be critical, was indeed demonstrated
experimentally to be critical (very clearly so in one of the flat panels and in the
two larger models of the second stage, but less obviously in the second flat panel).
The results of the two flat panels indicated that panels, stiffened by transverse
top-hat frames and having small imperfections, are likely to fail catastrophically
at a load close to the initial buckling stress, with little if any postbuckling reserve
of strength. The model analysis therefore emphasized important aspects of the
behavior of the structure, which were not obvious from the calculations.
In the second stage of modeling, the Perspex models and their supports in the
test rig were similar to the flat panels, but represented the actual combination of
panels and boundary conditions. The aim of the model of the full width vee-bottom
structure, comprising two panels similar to those tested in stage one, and incorpo-
rating a stiff keel girder, was to verify that under buckling conditions the deadrise
angle and keel girder would impose a plane of antisymmetry at the centerline.
The results verified this and indicated that the torsional stiffness of the keel girder

1 The strength to stiffness ratio can be expressed as that of the ultimate stress to Young’s modulus.
In perspex  u /E D¾ 0.03 for tension and  u /E ¾ D 0.05 for compression or bending, compared to
 u /E ¾
D 0.006 for both loading cases in a typical aluminum alloy 2024-T3.
Model Analysis in Structural Engineering 279

augmented the compressive strength of the panels. The purpose of the model of the
complete ship compartment was to check the design assumption that the curvature
at the ship’s bilges would resist buckling sufficiently to impose longitudinal node
lines, which would limit the effective transverse span of the bottom panels, an
assumption which was indeed confirmed. Both models of the second stage also
clearly demonstrated the dominance of the type-3 local instability.
Though the results of the small scale Perspex model tests were not considered
sufficient by the designers for complete assessment of the collapse behavior, and
therefore additional tests were later carried out on large-scale GRP panels and hull
sections, the small-scale models provided the initial guidelines on the buckling and
collapse behavior and an assessment of the design assumptions. They therefore
represent a good example of extended model analysis.

5.9.4 Design of Thames Barrier Gates

The design of the Thames Barrier gates is an example of the use of comprehensive
structural model testing as a primary design tool in a major civil engineering project
(see [5.62] and [5.63]). Dowling and Owens point out there that previous experience
in the Civil Engineering Laboratories of Imperial College, London, where the
model tests were carried out, “had proved for such complicated structures as ships,
the usefulness of small-scale Araldite models as a design tool to complement, and
indeed sometimes replace, expensive finite element analysis”.
The model analysis for the 61 m long rising sector gates consisted therefore
of a 1:25 scale Araldite model, as well as a large 1:6 scale steel model. The
small-scale model (see Figure 5.29), made of an epoxy casting resin Araldite 219,
was commissioned to check the linear elastic finite element modeling used in the
analysis of the actual gate, in particular in respect to the stress distributions in the
perforated webs and near the gate-to-gate arm connections. The results of the tests
(which included measurements from 700 strain gauges) increased the confidence
in the FEM analysis, but showed that some adjustments were necessary to it,
in particular to account for the arching action of the curved skin and for shear
stress peaks adjacent to openings. They also highlighted an extreme sensitivity of
the reactions to misalignment. All the observations from the small-scale Araldite
models were incorporated into the iterative analysis-test-analysis design procedure.
It may be mentioned here, that Araldite models had been used extensively also
by other investigators for buckling tests. For example, Tulk and Walker [5.64]
at University College, London, also employed small-scale Araldite 219 models
to elucidate the elastic buckling characteristics of stiffened-plate panels subjected
to in-plane compressive loading. They emphasized that because of the very high
elastic strain capacity of Araldite (maximum elongation up to 5 percent), its use in
models permits repeated testing well into the postbuckling regime to explore the
behavior of the structure without any permanent deformation. They pointed out
its convenient molding and that, by use of Araldite also as an adhesive, built up
structures which are practically homogeneous and free of residual stresses can be
obtained.
280 Modeling Theory and Practice

Figure 5.29 The 1:25 scale Araldite model for the Thames Barrier gates (from [5.63]):
(a) construction of the model, (b) the test rig with model under test
Model Analysis in Structural Engineering 281

Figure 5.30 The 1:6 scale steel model for the Thames Barrier gate general view during
reverse head test (from [5.62])

Returning to the model analysis of the Thames Barrier gates, the purpose of
the larger welded structural steel model was an ultimate load test to study the
complete response of the gate (including inelastic buckling) up to collapse, to
quantify the reserves of strength possessed by the structure beyond the initiation
of significant yielding and to establish the post-yield buckling behavior of the
compression flanges and webs. The steel 1:6 model (Figure 5.30) was instrumented
with 1000 strain gauges, their locations being determined by the results of the
Araldite model tests. Collapse was caused by inward buckling of the curved skin
at the change of section at the quarter point of the gate.
One should note that the extended model analysis, employed here as an integral
part of the iterative design process, not only improved the design but also reinforced
the confidence of the designers, at a time when civil engineering confidence had
been shaken by a series of recent tragic bridge disasters.

5.9.5 Photoelastic Models


Before leaving the topic of model analysis, one should also mention photoelastic
models, which are models made of transparent elastic material that when loaded
and examined in a field of polarized light, exhibit interference fringes that represent
the stress distribution in the model. Though photoelasticity is a major branch of
model analysis, it seems more appropriate to postpone its discussion to Chapter 20,
Volume 2, together with other optical methods.
282 Modeling Theory and Practice

5.10 Analogies
If the concept of model analysis is taken one step further, one obtains analo-
gies, where the model has lost any physical similarity with the prototype and
has preserved only a mathematical affinity with it. Analogies are widely used in
dynamic systems, since many time-dependent phenomena are analogous to elec-
trical ones. The next step leads to simulation by computer, which is also often
employed to extend the range of experiments. Though analogies are rarely used
for buckling phenomena, they are briefly discussed here on account of their poten-
tial for unconventional experiments. More detailed discussion of analogies can be
found in many texts (for example [5.65], [5.66] or [5.67]).
In engineering, the usual definition of analogy is: Two or more apparently
different physical systems are said to be analogous, if their characteristics can
be expressed in identical mathematical form. Analog methods were already devel-
oped in the second half of the 19th century, but they reached their prime in the
first half of this century. Among the most notable analogies was the membrane
analogy for the study of the torsional stress distribution in a shaft, developed by
Prandtl in 1903 [5.68].
This analogy is a good example of the concept and is discussed in nearly every
text on the theory of elasticity (as for example [4.46]). It is based on the identity
of the equation of vertical equilibrium of a stretched and inflated membrane
p
r2 z D  5.97
T
where z is the elevation of the membrane, T the uniform tension throughout the
membrane and p the small pressure differential, with an equation derived (by
integration) from the equation of compatibility of the stress function  for Saint
Venant torsion of a bar, whose cross-section is identical with the planform of the
membrane
r2  D 2G 5.98

where G is the shear modulus and  the angular twist per unit length. The boundary
conditions for z and  also have to match, which they do, since along a boundary s

∂z
D0 because the membrance boundary is 

∂s 

in a plane of constant z
and . 5.99
∂ 

D0 
because the boundary of the bar is 

∂s free of normal stress
The resultant shear stress at any point,  D d/dn, is represented by the slope
of the membrane dz/dn, taken normal to the contour line through that point.
Furthermore, the torsional moment Mt is represented by twice the volume under
the inflated membrane, since
References 283
 
Mt D 2  dx dy is analogous to 




 . 5.100


2 ð Volume D 2 z dx dy 

Note that from Eqs. (5.97) and (5.98), p/T D 2G when  D z.
The membrane analogy was applied extensively in the first decades of this
century to measure the slopes of soap films or rubber membranes of complicated
cross sections. A very famous and widely used apparatus is that of Griffith and
Taylor developed in 1917 [5.69], but many others and more sophisticated ones were
developed in the following decades. This demonstrates the direct use of an analogy
for measurement of the behavior of the analogous system in order to calculate that
of the original system.
However, the analogy serves also for better understanding of the essential
features of the original system. For example, in the case of the membrane analogy
it is easier to visualize the shape of the soap film and its slope and volume, and
evaluate from them the behavior of the twisted bar, than to assess it directly. This
is probably today the most important function of the membrane analogy, and as a
matter of fact of most analogies.
The direct application by measurement of the behavior of the analogous system
is today usually superseded by numerical solutions (like Finite Element Methods),
solved conveniently by available computer programs. Hence the many other inge-
nious hydrodynamic and electrical analogies, developed in the thirties, forties,
fifties and sixties, will not be discussed (and the reader be referred to the texts
mentioned earlier, where many references are also given), except one important
example, the electrical circuit analogies for structures developed by MacNeal in
the early sixties [5.70] which will be briefly mentioned. MacNeal proposed analog
computation for solution of many practical problems of structural analysis using
direct analog computers. He derived detailed electrical circuit analogies for these
problems and obtained with them and the appropriate analog computers very effi-
cient solutions for complicated problems, which represented a significant advance
in structural analysis. Modern digital computation has superseded these analog
methods, but their ingenuity should be noted as one non-conventional approach to
combined experimental-computational analysis.
Finally, it may be worth noting that in the last decade electrical analog tech-
niques have been revived in fracture mechanics for experimental study of crack
propagation and evaluation of stress intensity factors, recently also for composites
(see for example [5.71].

References

5.1 Dym, C.L. and Ivey, E.S., Principles of Mathematical Modeling, Academic Press,
New York, 1980.
5.2 Müller, R.K., Handbuch der Modellstatik, Springer-Verlag, Berlin, 1971.
284 Modeling Theory and Practice

5.3 Charlton, T.M., Model Analysis of Structures, John Wiley & Sons, New York, 1954.
5.4 Langhaar, H.L., Dimensionless Analysis and Theory of Models, John Wiley & Sons,
New York, 1951.
5.5 Murphy, G., Similitude in Engineering, Ronald Press, New York, 1950.
5.6 Pankhust, R.C., Dimensional Analysis and Scale Factors, Chapman & Hall, London,
Reinhold, New York, 1964.
5.7 Gukhman, A.A., Introduction to the Theory of Similarity, Academic Press, New York,
1965.
5.8 Taylor, E.S., Dimensional Analysis for Engineers, Oxford (Clarendon Press), London
and New York, 1974.
5.9 Goodier, J.N., Dimensionless Analysis, Appendix II in Handbook of Experimental
Stress Analysis, M. Hetényi, ed., 1st edn., John Wiley & Sons, New York, 1950.
5.10 Durelli, A.J., Phillips, E.A. and Tsao, C.H., Introduction to the Theoretical and
Experimental Analysis of Stress and Strain, McGraw-Hill, New York, 1958.
5.11 Ipsen, D.C., Units, Dimension, and Dimensionless Numbers, McGraw-Hill, New
York, 1960.
5.12 Buckingham, E., On Physically Similar Systems; Illustrations of the Use of Dimen-
sional Equations, Phys. Rev. Series 2, 4,(4), 1914, 345 76.
5.13 Bridgman, P.W., Dimensionless Analysis, Yale University Press, New Haven, 1922,
1931 (revised edition).
5.14 Van Driest, E.R., On Dimensional Analysis and the Presentation of Data in Fluid-
Flow Problems, Journal of Applied Mechanics, 13, 1946, A-34 A-40.
5.15 Goodier, J.N. and Thomson, W.T., Applicability of Similarity Principles to Struc-
tural Models, NACA TN 933, 1944.
5.16 Manjoine, M.J., Influence of Rate of Strain and Temperature on Yield Stresses of
Mild Steel, Journal of Applied Mechanics, 11, 1944, 211 218.
5.17 Marsh, K.J. and Campbell, J.D., The Effect of Strain Rate on the Post-Yield Flow
of Mild Steel, Journal of the Mechanics and Physics of Solids, 11, 1963, 49 63.
5.18 Bodner, S.R., Strain Rate Effects in Dynamic Loading of Structures, in Behavior of
Materials Under Dynamic Loading, N.J. Huffington, ed., ASME, New York, 1965,
93 105.
5.19 Symonds, P.S., Viscoplastic Behavior in Response of Structures to Dynamic Loading,
in Behavior of Materials Under Dynamic Loading, N.J. Huffington, ed., ASME, New
York, 1965, 106 124.
5.20 Jones, N., Structural Aspects of Ship Collisions, in Structural Crashworthiness,
N. Jones, and T. Wierzbicki, eds., Butterworths, London and Boston, 1983,
308 337.
5.21 Jones, N., Scaling of Inelastic Structures Loaded Dynamically, in Structural
Impact and Crashworthiness Vol 1, G.A.O. Davies, ed., Elsevier Applied Science
Publishers, London, 1984, 45 74.
5.22 Duffy, T.A., Scaling Laws for Fuel Capsules Subjected to Blast, Impact and Thermal
Loading, in Proceedings Intersociety Energy Conversion Engineering Conference,
SAE Paper No. 719107, 1971, 775 786.
5.23 Booth, E., Collier, D. and Miles, J., Impact Scalability of Plated Steel Structures, in
Structural Crashworthiness, N. Jones, and T. Wierzbicki, eds., Butterworths, London
and Boston, 1983, 136 174.
5.24 Duffy, T.A., Cheresh, M.C. and Sutherland, S.H., Experimental Verification of
Scaling Laws for Punch-Impact Loaded Structures, International Journal of Impact
Engineering, 2, 1984, 103 117.
References 285

5.25 Bodner, S.R. and Symonds, P.S., Experimental and Theoretical Investigation of the
Plastic Deformation of Cantilever Beams Subjected to Impulsive Loading, Journal
of Applied Mechanics, 29, 1962, 719 728.
5.26 Calladine, C.R. and English, R.W., Strain-Rate and Inertia Effects in the Collapse
of Two Types of Energy-Absorbing Structure, International Journal of Mechanical
Sciences, 26, 1986, 689 701.
5.27 Tam, L.L., Strain-Rate and Inertial Effects in the Collapse of Energy-Absorbing
Structures, Ph.D. Thesis, University of Cambridge, England, February 1990.
5.28 Zhang, T.G. and Yu, T.X., A Note on a “Velocity Sensitive” Energy Absorbing
Structure, International Journal of Impact Engineering, 8, 1989, 43 51.
5.29 Nurick, G.N. and Martin, J.B., Deformation of Thin Plates Subjected to Impulsive
Loading A Review, Part II: Experimental Studies, International Journal of Impact
Engineering, 8, 1989, 171 186.
5.30 Johnson, W., Impact Strength of Materials, Edward Arnold, London, 1972.
5.31 Nurick, G.N., Pearce, H.T. and Martin, J.B., The Deformation of Thin Plates Subjec-
ted to Impulsive Loading, in Inelastic Behaviour of Plates and Shells, L. Bevilacqua,
ed., Springer-Verlag, Berlin, 1986.
5.32 Wierzbicki, T. and Florence, A.L., A Theoretical and Experimental Investigation of
Impulsively Loaded Clamped Circular Viscoplastic Plates, International Journal of
Solids and Structures, 6, 1970, 555 568.
5.33 Bodner, S.R. and Symonds, P.S., Experiments on Viscoplastic Response of Circular
Plates to Impulsive Loading, Journal of the Mechanics and Physics of Solids, 27,
1979, 91 113.
5.34 Baker, W.E., Westine, P.S. and Dodge, F.T., Similarity Methods in Engineering
Dynamics, Spartan Books, Hayden Book Co., Rochelle Park, N.J., 1973.
5.35 Baker, W.E., Modeling of Large Transient Elastic and Plastic Deformations of Struc-
tures Subjected to Blast Loading, Journal of Applied Mechanics, 27, 1960, 521 527.
5.36 Hopkinson, B., British Ordnance Board Minutes 13565, 1915.
5.37 Sachs, R.G., The Dependence of Blast on Ambient Pressure and Temperature, Ballis-
tics Research Lab. (BRL), Report No. 466, Aberdeen Proving Ground, Maryland,
1944.
5.38 Brown, H.N., Effects of Scaling on the Interaction Between Shock Waves and Struc-
tures, Ballistics Research Lab. (BRL), Report No. 1011, Aberdeen Proving Ground,
Maryland, 1957, Appendix I.
5.39 Hanna, J.W., Ewing, W.O. and Baker, W.E., The Elastic Response to Internal Blast
Loading of Models of Outer Containment Structures for Nuclear Reactors, Nuclear
Science and Engineering, 6, 1959, 214 221.
5.40 Denton, D.R. and Flathau, W.J., Model Study of Dynamically Loaded Arch Struc-
tures, Journal of the Engineering Mechanics Division, Proc. of ASCE, 92, (EM3),
1966, 17 32.
5.41 Ewing, W.O. and Hanna, J.W., A Cantilever for Measuring Air Blast, Ballistics
Research Lab. (BRL), Technical Note 1139, Aberdeen Proving Ground, Maryland,
1957.
5.42 Ezra, A.A. and Penning, F.A., Development of Scaling Laws for Explosive Forming,
Experimental Mechanics, 2, 1962, 234 239.
5.43 Ezra, A.A. and Adams, J.E., The Explosive Forming of 10 feet Diameter Aluminum
Domes, Proc. of the First International Conference of the Center for High Energy
Forming, Estes Park, Colorado, June 19 23, 1967.
286 Modeling Theory and Practice

5.44 Baker, W.E. and Westine, P.S., Modeling the Blast Response of Structures Using
Dissimilar Materials, AIAA Journal, 7, 1969, 951 959.
5.45 Donelan, P.J. and Dowling, A.R., The Use of Scale Models in Impact Testing, in
The Resistance to Impact of Spent Magnox Fuel Transport Flasks, The Institution of
Mechanical Engineers, London, 1985.
5.46 Jones, N. and Jouri, W.S., A Study of Plate Tearing for Ship Collision and Grounding
Damage, Journal of Ship Research, 31, 1987, 253 268.
5.47 Morton, J., Scaling of Impact-Loaded Carbon-Fiber Composites, AIAA Journal, 26,
1988, 989 994.
5.48 Jackson, K.E. and Fasanella, E.L., Scaling Effects in the Static Large Deflection
Response of Graphite-Epoxy Beam-Columns, NASA Technical Memorandum (TM)
101619, June 1989, also Proceedings of the American Helicopter Society National
Technical Specialists’ Meeting on Advanced Rotorcraft Structures, Williamsburg, VA,
Oct. 25 27, 1988.
5.49 Jackson, K.E. and Fasanella, E.L., Scaling Effects in the Impact Response of Gra-
phite-Epoxy Composite Beams, SAE Technical Paper 891014, General Aviation
Aircraft Meeting and Exposition, Wichita, KS, April 11 13, 1989.
5.50 Jackson, K.E., Scaling Effects in the Static and Dynamic Response of Graphite-
Epoxy Beam-Columns, NASA TM 102697, July 1990.
5.51 Jackson, K.E. and Morton, J., Evaluation of Some Scale Effects in the Response
and Failure of Composite Beams, Presented at First NASA Advanced Composite
Technology (ACT) Program Conference, Seattle, WA, Oct. 21 Nov. 1, 1990.
5.52 Kellas, S. and Morton, J., Strength Scaling of Fiber Composites, NASA Contractor
Report 4335, November 1990.
5.53 Qian, Y. and Swanson, S.R., Experimental Measurement of Impact Response in
Carbon/Epoxy Plates, AIAA Journal, 28, 1990, 1069 1074.
5.54 Qian, Y., Swanson, S.R., Nuismer, R.J. and Bucinell, R.B., An Experimental Study
of Scaling Rules for Impact Damage in Fiber Composites, Journal of Composite
Materials, 24, (5), May 1990, 559 570.
5.55 Swanson, S.R., Smith, N.L. and Qian, Y., Analytical and Experimental Strain
Response in Impact of Composite Cylinders, Composite Structures, 18, (2), 1991,
95 108.
5.56 Christoforou, A.P., Investigation of Impact in Advanced Composites, Ph.D. Disser-
tation, University of Utah, Department of Mechanical Engineering, 1988.
5.57 Kampf, K.-P., Crawley, E.F. and Hausman, R.J., Experimental Investigation of the
Crashworthiness of Scaled Composite Sailplane Fuselages, Journal of Aircraft, 26,
1989, 675 681.
5.58 Wright, D.V. and Bannister, R.C., Plastic Models for Structural Analysis, Part I:
Testing Types, The Shock and Vibration Digest, 2, (11), 1970, 2 10.
5.59 Wright, D.V. and Bannister, R.C., Plastic Models for Structural Analysis, Part II:
Experimental Design, The Shock and Vibration Digest, 2, (12), 1970, 3 10.
5.60 Smith, C.S., Investigation of Ship Buckling Problems Using Small-Scale Plastic
Models, Proceedings 5th International Conference on Experimental Stress Analysis,
Udine, 1974, 4.127 4.135.
5.61 Smith, C.S., Buckling Problems in the Design of Fiberglass-Reinforced Plastic Ships,
Journal of Ship Research, 16, (3), 1972, 174 190.
5.62 Tappin, R.G.R., Dowling, P.J. and Clark, P.J., Design and Model Testing of the
Thames Barrier Gates, The Structural Engineer, 62A, (4), 1984, 115 124.
References 287

5.63 Dowling, P.J. and Owens, G.W., Structural Model Testing of a Rising Sector Flood
Gate, in Thames Barrier Design, Institution of Civil Engineers, London, 1978,
117 124.
5.64 Tulk, J.D. and Walker, A.C., Model Studies of the Elastic Buckling of a Stiffened
Plate, Journal of Strain Analysis, 11, (3), 1976, 137 143.
5.65 Mindlin, R.D. and Salvadori, M.G., Analogies, Ch. 16 in Handbook of Experimental
Stress Analysis, M. Hetényi, ed., John Wiley & Sons, New York, 1950, 700 827.
5.66 Sutherland, R.L., Engineering Systems Analysis, Addison-Wesley Publishing Co.,
Reading, MA., 1958.
5.67 Lee, G.H., An Introduction to Experimental Stress Analysis, John Wiley & Sons,
New York, 1950, 225 244.
5.68 Prandtl, L., Zur Torsion von prismatischen Stäben, Physikalische Zeitschrift, 4, 1903,
758 759.
5.69 Griffith, A.A. and Taylor, G.I., The Use of Soap Films in Solving Torsion Problems,
Proceedings of the Institution of Mechanical Engineers, London, 1917, 755 809.
5.70 MacNeal, R.H., Electric Circuit Analogies for Elastic Structures, John Wiley & Sons,
New York, 1962.
5.71 Srinivasan, G.V. and Virkar, A.V., Application of the Electrical Analog Technique
in Fiber-Reinforced Composites, Engineering Fracture Mechanics, 32, (3), 1989,
479 492.
6
Columns, Beams and
Frameworks

6.1 Buckling and Postbuckling of Columns

6.1.1 Column Curves and “Secondary” Effects in Column


Experiments

In the three-quarters of a century since von Kármán’s thesis, the buckling and
postbuckling of columns has been studied extensively, with a significant portion
of the efforts devoted to experimental investigations. Many of these deal with the
interaction between material properties, residual stresses, shape of cross-section and
the postbuckling behavior, and many are design oriented. As a matter of fact, while
considerable progress has been made towards better prediction of the buckling load
in the inelastic region, and the maximum load a column can carry its strength,
the designers have usually been using empirical column curves. A column curve
is a plot of load, or stress, versus slenderness ratio, and is the line of best fit
through the scatter band of column test results (see for example Figure 6.1). One
should note that for over a century the universal practice was to lump together test
results for different materials and different cross-sections which therefore appeared
as a galaxy of points (see for example Figure 6.2, or one of the many figures in
Chapter 4 of [4.3]). A number of empirical and semi-empirical design formulae
have been developed, some dating back as far as the 18th century (see [4.3]), which
are usually called column curves. Until the fifties, the most significant ones were:
the Rankine Gordon formula (see also [4.3]), the Tetmajer straight line [4.8] and
the Johnson Parabola
 2
L
cr D y  C 6.1

where C is a constant depending on the proportional limit and Young’s modulus
of the material of the column. The 1893 Johnson Parabola [6.2] is the basis of the
modern column curves, which have been developed in recent decades, by special
national and international bodies and committees of experts. The leader among

Buckling Experiments: Experimental Methods in Buckling of Thin-Walled Structures: Basic Concepts, Columns, Beams and
Plates – Volume 1. J. Singer, J. Arbocz and T. Weller Copyright © 1998 John Wiley & Sons, Inc.
290 Columns, Beams and Frameworks

Figure 6.1 Column curves and test results for rolled H-shapes (from [6.1])

these is the Structural Stability Research Council (formerly the Column Research
Council) of the Engineering Foundation, a US (but essentially international) body,
that has for nearly 50 years fostered research and developed design and test proce-
dures for column stability. The SSRC “Guide to Stability Design Criteria for Metal
Structures” [6.3] is not only an internationally recognized design guide (though
aimed primarily at civil, mechanical and marine engineers, it is also an author-
itative source for work carried out in other fields), but also a guide to modern
column testing, to which we will refer later. Other similar, well-known bodies are
the Column Research Committee of Japan [6.4] or the European Convention for
Constructional Steelwork (ECCS) [6.5].
The original aim has been to develop a single design curve (the CRC curve in
Figure 6.2), but since the wide scatter is not a test phenomenon, the alternative
of multiple column curves has been extensively studied. The resulting wealth of
empirical information, augmented by many theoretical studies eventually brought
about the adoption of the concept of multiple column curves for the design of
steel columns both by the US Structural Stability Research Council (SSRC), [6.3],
and the European Convention for Constructional Steelwork (ECCS), [6.5]. The
SSRC multiple column curves are shown in Figure 6.3a and the ECCS ones in
Figure 6.3b. Note that, as pointed out by Tall in [4.7], the American and European
multiple column curves correlate very well, though they were obtained by different
approaches, the US studies using actual measured values while the European studies
used theoretical data as a basis for computations that were then compared with
experimental data. In Europe, the ECCS multiple column curves were adopted for
design practice, but in the US designers still prefer a single design curve (see [4.7]
and [1.13]).
Now, as mentioned already, the bulk of the theoretical and experimental studies
on buckling of columns in recent decades has dealt with inelastic behavior of
steel columns and the influence of yield strength, of geometry, of residual stresses
Buckling and Postbuckling of Columns 291

Figure 6.2 Test results for columns of different shapes, yield strength and fabrication methods
(from [4.7])

(resulting from the different manufacturing processes), and of out-of-straightness


(as the geometric imperfections are called in columns). Tall [4.7] summarized
these experiments and the resulting design curves from the point of view of civil
engineers, emphasizing the effect of residual stresses, which have been the major
factor in the design of welded steel columns (see also [4.6], [6.1], [6.3] or [6.6]).
Residual stresses occur in a structural member as a result of plastic deformations
during manufacture. They may be due to differential cooling after hot-rolling,
to fabrication processes like flame-cutting or cold-bending, or due to localized
292 Columns, Beams and Frameworks

Figure 6.3 Multiple column curves (from [4.7]): (a) proposed US SSRC multiple column
curves (where L D light, H D heavy), (b) European ECCS multiple column curves

heat input in welding operations. Typical magnitudes and distributions of residual


stresses in rolled and welded steel shapes are shown in Figure 6.4 (reproduced
from [4.7]). As seen in the figure, welded columns usually have higher residual
stresses than rolled columns and their magnitude depends on the geometry of the
cross-section. They also tend to have a greater out-of-straightness. Hence welded
columns have lower strengths than corresponding rolled columns (see Figure 6.5
Buckling and Postbuckling of Columns 293

Figure 6.4 Residual stresses due to welding in cross-sections of small columns (from [4.7])

Figure 6.5 Column tests results for small to medium rolled and welded shapes (from [4.7])

reproduced from [4.7], and this strength has to be assessed by a more complicated
analysis of the behavior in the inelastic range.
Both size and yield strength influence the strength of a steel column. The process
of cooling in heavy shapes (large size cross sections) yields larger residual stresses
than in small size shapes, whether rolled or welded, and hence heavy columns have
reduced strengths. Since the residual stresses are mainly a function of geometry,
they are of the same order of magnitude in high strength steels as in mild steels.
Thus the effect of residual stresses is smaller in columns made of steels with higher
yield strength.
294 Columns, Beams and Frameworks

This summary of the main characteristics of the buckling strength of steel


columns, derived from decades of extensive steel column tests, emphasizes the
important interaction of material properties, geometry and fabrication processes on
the buckling behavior and strength of columns. It furthermore indicates that one
has to be very careful to include also such “secondary” effects in the design and
evaluation of buckling and postbuckling tests of all structural elements.

6.1.2 Column Testing

Having briefly discussed the main “secondary” effects in column experiments,


one can proceed to the test procedures. Here one can turn to the SSRC “Guide
to Stability Design Criteria for Metal Structures” [6.3] for guidance on modern
column testing, which indeed appears there in the form of Technical Memoranda.
In Appendix B there, after a preface which points out that some of the proposed
methods are not always used, the recommended test procedures for compression
testing of metals are presented (pp. 703 708) and then those for stub-column tests
(pp. 708 717). The object of a compressive stub-column test is to determine the
average stress-strain relationship over the complete cross-section, which can then
be employed as the actual material properties for the column test. Technical Memo-
randum No. B4 there “Procedure for Testing Centrally Loaded Columns” ([6.3],
pp. 717 732) is based on a 1970 Lehigh University Fritz Engineering Labora-
tory Report [6.7], which summarized the extensive test experience accumulated at
that university. It discusses the reasons for experimental scatter in column tests
and then presents a suggested test procedure. Because of its importance, the main
points of this memorandum are briefly recapitulated, some paragraphs being quoted
verbatim.
The reasons for the wide scatter band of the experimentally determined values
of column strength when plotted versus the effective slenderness ratio KL/r, in
which KL denotes the effective column length and r the appropriate radius of
gyration of the cross section, are enumerated as:

1. geometrical imperfections (out-of-straightness)


2. eccentric application of load
3. nonhomogeneity of material
4. residual stresses
5. variation in the action of the loading machines
6. imperfections in end fixtures.

These effects have already been discussed, except the last two which are directly
related to the tests themselves. The buckling and postbuckling behavior of a
column, or of any other structural element, is influenced by the action of the
loading devices. These may be categorized as gravity, deformation (screw-type) and
pressure (hydraulic) testing machines, each differing in its force-deflection charac-
teristic. The gravity type has the simplest characteristic, which can be represented
by straight lines parallel to the deflection axis. The screw-type load deflection
Buckling and Postbuckling of Columns 295

characteristic is also well defined, and its shape depends on the elastic response
of the loading system. The hydraulic testing machine is the most common today,
but its load-deflection characteristic is not as easily defined and testing is always
conducted under some finite loading rate, which influences the results. However,
modern testing machines (both hydraulic and screw-type) have continuous feed-
back computer control that assures precise predetermined displacement or loading
rates, which can be kept very low. For example, both MTS and Instron testing
machines can apply displacement rates as low as 1 micron/hour, with a resolution
of a few percent of that rate, as well as similar loading rates. These machines can
also apply a constant force for considerable test times (about 20 hours or so).
End conditions can vary from full restraint (fixed) to zero restraints (pinned,
simple supports) with respect to end rotation and warping. The pinned-end condi-
tions are recommended for column tests, since then the critical cross section is
located near the mid-height of the column and is therefore little influenced by end
effects. With pinned-end conditions it is, however, necessary to provide end fixtures
with minimum restraint to column end rotation. Under fixed-end conditions, on the
other hand, there are often problems of variation of the end restraints, and hence
the effective length, with load, which make the tests less reliable.
Figure 6.6 shows several practical pinned ends (from [6.8]), some are “position-
fixed” like (a), (d) and (h), and the others are “direction-fixed”, having cylindrical
end fixtures, with which the column is essentially pin-ended about one axis (usually
chosen to be the minor principal axis of the column cross-section) and essen-
tially fixed end, or clamped, about the other. The cylindrical (and hemispherical)
fixtures, (e), (f), (g) and (d), are designed to have their center coincide with the
centroidal axis of the cross-section at the column end. Thus the actual column
length remains the effective one when the column starts to bend. Knife edges,
Figure 6.6(b), conical points (a), or free warping ends like (h), are suitable only
for small columns. Complete roller bearings, as shown in Figure 6.6(g), were used
for the well known 1938 tests of Karner and Kollbrunner at the ETH, Zürich [6.9]
on centrally and eccentrically loaded small aluminum alloy (Avional M) and struc-
tural steel columns. For large columns, requiring the application of large axial loads,
roller bearing blocks, like Figure 6.6(e), are sometimes used, or a relatively large
hardened cylindrical surface bearing on a hard flat surface is employed, for example
the end fixtures used at the Lehigh University Fritz Engineering Laboratory for
loads between 400 1000 tons, shown in Figure 6.7 (from [6.10]). Hemispherical
fixtures, approximately similar to those of Figure 6.6(d), are also used sometimes in
tests of large columns. For example, in the experiments on plastic column behavior
at high axial loads, carried out at Imperial College London in the seventies [6.11],
the large horizontally “floating” test columns were held in position by spherical
PTFE (polytetrafluoroethene) bridge bearings. This was an economical solution,
that permitted flexing about both axes. There were, however, significant friction
losses in the PTFE bearings, which were measured by H-section load cells attached
at the end of the test columns. Roller bearings similar to those of Figure 6.6(g) were
employed for pin ends in the “classical” 1939 Aluminum Company of America-
NACA tests of extruded aluminum H-sections [1.29]. Another method for pin ends
296
Columns, Beams and Frameworks

Figure 6.6 End fixtures for pin-ended columns (from [6.3])


Buckling and Postbuckling of Columns 297

Figure 6.7 Standard large column end fixture at Lehigh University Fritz Engineering Labora-
tory (from [6.10])

in those tests, a type of “position fixed” ends, was a mixture of the hemisphere in
Figure 6.6(d) and the rollers in Figure 6.6(e) consisting of bearing plates provided
with a spherical seat resting in a nest of 25 hardened steel balls, whose center
of rotation coincided with the ends of the flat-ended specimen that rested on the
bearing plates. The corresponding US National Bureau of Standards NACA tests
used knife edges for pinned ends. Flat-ended ends were obtained by centering the
mutually parallel machined flat ends of the specimen on the fixed heads of the
testing machine.
Hemispherical pin-end supports of the type shown in Figure 6.6(d) were used in
the eighties in a typical experimental study of local and overall buckling of welded
steel box columns [6.12] and are shown in Figure 6.8(a). A second example of
hemispherical pin-end supports, of the type shown in Figure 6.6(d), are those used
in the late eighties in the large beam column tests at the University of Toronto,
discussed in Section 6.6.2 (see Figure 6.58 and [6.118]). Another roller type of
pin-end supports, a combination of those shown in Figures 6.6(e) and 6.6(g), that
was employed in the eighties in a series of tests on heavy I-section columns at
Karlsruhe University in Germany [6.13] is shown in Figure 6.8(b). The special
SKF rollers in these supports ensured a low friction coefficient  < 0.07.

6.1.3 Test Procedures


In the actual test procedure, some important points (from Technical Memorandum
No. 4 of [6.3]) should be remembered. These are obviously only general guidelines,
most appropriate to columns used in civil engineering.
298 Columns, Beams and Frameworks

Figure 6.8 Typical hemispherical and roller bearings employed for heavy columns:
(a) hemispherical pin-end fixture used in a buckling study of welded steel box
columns at Nagoya University, Japan (courtesy of Professor T. Usami), note the
roller bearing in the center of the picture, which provided the pin-end support;
(b) roller pin-ends used at Karlsruhe University, Germany, for heavy I-section
columns (from [6.13])
Buckling and Postbuckling of Columns 299

a. Preparation of Specimens
Both ends of the specimen should be milled. “Columns may be tested with the
ends bearing directly on the loading fixtures, provided the material of which the
loading fixtures are made is sufficiently harder than that of the column to avoid
damaging the fixtures. Otherwise, base plates should be welded to the specimen
ends, matching the geometric center of the specimen to the center of the base
plate. The welding procedure should be such that compressive residual stresses at
the flange tips caused by the welding are minimized. For columns initially curved,
the milled surfaces may not be parallel to each other, but will be perpendicular to
the centerline at the ends because milling is usually performed with reference to the
end portions of the columns. For relatively small column specimens, it is possible
to machine the ends flat and parallel to each other by mounting the specimens on
an arbor in a lathe. For small deviations in parallelism, the leveling plates at the
sensitive crosshead of the testing machine may be adjusted to improve alignment.”

b. Initial Dimensions
The variation in cross-sectional area and shape, and the initial curvature (“camber”
referred to major axis and “sweep” referred to minor axis), and twist, will affect
the column strength. Therefore, detailed initial measurement of these parameters
of the specimen is important.

c. Aligning the Column Specimen


“Aligning the specimen within the testing machine is the most important step in
the column testing procedure, prior to loading. Two approaches have been used to
align centrally loaded columns. In the first approach the column is aligned under
load such that the axial stresses are essentially uniform over the mid-height and
the quarter-point cross sections. (In the test one actually measures strains.) The
objective in this alignment method is to maximize the column load by minimizing
the bending stresses caused by geometrical imperfections of the specimen. In the
second alignment method, the column is carefully aligned geometrically, but no
special effort is made to secure a uniform stress distribution over the critical cross
section. Geometric alignment is performed with respect to a specific reference
point on the cross section. The method of geometric alignment is recommended
for columns as it is, generally, simpler and quicker.” As a matter of fact, in recent
years the first method has practically disappeared, and geometric alignment with
exact measurements of initial out-of-straightness, coupled with analytical strength
predictions, is usually used.
In other structural elements, however, the “uniform stress” approach is usually
preferable, as will be discussed later.

d. Instrumentation
“. . . It is usually desirable to measure the more important deflections and twists
to compare the behavior of the column specimen under load, with theoretical
300 Columns, Beams and Frameworks

predictions of behavior. The instrumentation for column tests has changed markedly
in recent years due to progress made in measuring techniques and data acquisition
systems, and it is now possible to obtain automatic recordings and plotting of the
measurements (online real-time presentation). Such recordings are more convenient
and more precise than manual readings.
The most important records needed in column testing are the applied load and
the corresponding lateral displacements, twist, and overall column shortening. A
typical column set-up and instrumentation are shown in Figure 6.9 (from [6.3]).
Lateral deflections normal to both principal cross-sectional axes may be auto-
matically recorded by means of potentiometers attached at quarter points of the
column (more points may be used for longer columns). Lateral deflections may
also be measured from strip scales attached to the column and read with the aid
of a theodolite.
Strains are measured using electric-resistance strain gages. For ordinary pinned-
end column tests, it is sufficient to mount eight strain gages at each end and at the
mid-height level. . . . As shown in Section A-A of Figure 6.9, the gages should be
mounted in pairs ‘back-to-back’ to enable the local flange bending effects to be
cancelled by averaging the readings of each pair of ‘back-to-back’ gages.
In the fixed-end test condition more strain gages are mounted below and above
the quarter- and three-quarter levels. This is done to determine the actual effec-
tive length of the column by locating the inflection points using the strain gage
measurements.”
With modern multi-channel data loggers more strain gages can be readily used
to obtain additional check data and thus improve the reliability of the test.
“End rotations are measured by mechanical or electrical rotation gages
(see [6.3]). . . . The angles of twist are determined at mid-height and at the two
ends by measuring at each level the differences in lateral deflections of the two
flanges. For better accuracy, the measurements may be taken at points located at
the ends of two rods attached transversely on the adjacent sides of the column, as
shown in Section B-B of Figure 6.9.
The overall shortening is determined by measuring the movement of the sensitive
crosshead relative to the fixed crosshead using a dial gage or potentiometer,” or
preferably a few gages, potentiometers or LVDTs.
Large “steel column specimens are usually whitewashed with hydrated lime.
During testing, the whitewash cracking pattern indicates the progression of yielding
in the column (the cracking reflects the flaking of the mill-scale at yielded zones).”

e. Testing
“After the specimen is aligned in the testing machine, the test is usually started
with an initial load of 1/20 to 1/15 of the estimated ultimate load capacity of the
column. This is done to preserve the alignment established at the beginning of the
test. At this load all measuring devices are adjusted for initial readings.
Further load is applied slowly, typically at a rate of 1 ksi/min (6.9 MPa/min),
and the corresponding deflections are recorded instantly. This stress rate, (or corre-
sponding strain rate) is established when the column is still elastic. The dynamic
Buckling and Postbuckling of Columns 301

Figure 6.9 Typical column test setup and instrumentation (from [6.3])

curve is plotted until the ultimate load is reached, immediately after which the
‘maximum static’ load is recorded. . . . A static condition, as is needed to obtain
the ‘maximum static’ load, is when the column shape is unchanged under a constant
load for a period of time. This means that the chord length of the column must
remain constant, or practically, the distance between the crossheads must remain
constant during the period.”
302 Columns, Beams and Frameworks

This condition can easily be maintained with screw type machines, but was diffi-
cult to maintain in hydraulic machines a decade ago. In modern testing machines
with feedback computer control these difficulties have been eliminated.
“After the maximum static load is recorded, compression of the specimen is
resumed at the strain rate which was utilized for the elastic range. . . . The specimen
is compressed in the ‘unloading range’ until the desired load-displacement curve
is attained. [An example of such a curve is shown in Figure 6.10 (from [6.3])].”
One should note that, as pointed out in Chapter 2, the dynamic load is larger
than the static one. This means that a column can sustain a considerably higher
buckling load if the load is applied rapidly, i.e. under impact, as will be further
discussed in Chapter 18, Volume 2.

f. Presentation of Test Data

“The behavior of the test specimen under load well into the post-buckling region
is determined with the assistance of measurements of lateral deflections at various
levels along the two principal directions, rotations at the ends, strains at selected
cross sections, angles of twist, and the column shortening. These measurements
are compared to theoretical predictions. The results of the test are most clearly
presented in diagrammatic form.”
For example, Figure 6.11(a) (from [6.3]) shows the mid-height load-deflection
curve of a typical structural steel H-section column, along the minor axis, where
the primary bending occurs. Figure 6.11(b) shows the corresponding similar curve
along the major axis. These curves present the most significant data of the column
test. Similar curves are usually presented for strains, end rotations, angles of twist
and overall shortening versus load.

Figure 6.10 Typical load-deflection curve of a column (from [6.3])


Buckling and Postbuckling of Columns 303

Figure 6.11 Load-deflection curves for a typical structural steel H-section column (from [6.3]):
(a) midheight deflection along minor axis, (b) midheight deflection along major
axis

g. Evaluation of Test Results

“The test results may be evaluated by comparing the experimental load-deflection


behavior, or axial strain-bending strain behavior, and the theoretical prediction.
A preliminary theoretical prediction can be made on simplified assumptions of
material properties, residual stresses, and measured initial out-of-straightness. The
prediction may be improved if the actual residual stresses and the variations in
material properties are used in the analysis. These properties should be deter-
mined from preliminary stub-column tests of specimens obtained from the original
source stock.”

6.1.4 Columns in Offshore Structures

The widespread design and development of large offshore structures in the last
decade has motivated considerable research efforts related to the buckling of these
structures. A recent state-of-the-art review [6.14] considers only tubular columns in
the discussion of columns, since both the main and the bracing members of a typical
offshore structure are usually circular cylinders. It is of interest to note that the
emphasis in this review and design guide is on information based on experimental
investigations. The discussion of column buckling experiments in [6.14] is essen-
tially similar to that in the SSRC Guide [6.3], except that the comparisons are with
the special codes and design recommendation developed for offshore structures
and that interaction with local buckling (shell buckling) is considered in detail.
Since offshore platforms are usually designed as highly redundant space frames,
where buckling of an individual member will not necessarily lead to failure
304 Columns, Beams and Frameworks

of the structure as a whole, and since they are subjected on rare occasions to
extreme loads, post-collapse characteristics are very important for assessment of the
survivability of offshore structures. The post-collapse behavior of tubular columns
strongly depends on whether collapse is initiated by local instability.
If local stability predominates, the post-buckling behavior is that of a cylindrical
shell, which is highly unstable, as has been discussed in Chapter 2. If local buckling
is avoided, as in tubes with low (D/t) ratio, the post-collapse behavior is controlled
by the ratio of the Euler critical stress xkE to the yield stress y , and by the magni-
tude of initial imperfections (out-of-straightness). Figure 6.12 (from [6.14]) shows
typical average axial stress-strain curves for tubular columns. (These curves are
typical presentations corresponding to experimental results.) For the two extreme
cases (a) xkE × y and (c) xkE − y , collapse is gradual and a significant load
carrying capacity is retained. However, in case (b), when 0.7  xkE /y  1.3,
post-buckling can be highly unstable for tubes with small imperfections, with
significant reduction in load-carrying capacity in the post-collapse range. The pres-
ence of large imperfections considerably reduces the pre-buckling stiffness and the
buckling load in all cases, but collapse then occurs very gradually and with little
reduction in load carrying capacity.

6.1.5 End-Fitting Effects in Column Tests


As mentioned in Chapter 4, von Kármán in his classical 1910 paper already recog-
nized two possible sources of errors connected with the end fittings of simply
supported (pin-ended) columns: their rigidity and their restraint to column end rota-
tion. The calculations showed, however, that for his experiments and other typical
columns these errors were insignificant. Von Kármán’s analysis of the elastic buck-
ling of columns with rigid end connections was presented again in the “classic”
1940 text of von Kármán and Biot [6.15] in a slightly different form, that however
yielded similar results. Later investigators reexamined this effect for short columns,
suspecting there a more significant influence. For example, Chilver in 1956 [6.16]
considered the same type of column, but extended the von Kármán study beyond
the elastic range, to be suitable for short columns. Figure 6.13 shows schematically
a typical column with rigid knife ends. The fraction (2a/L) is the rigid portion of
the total length of the column. The “effective” length a of the end fittings is the
length that can be regarded as completely rigid from the bending point of view.
The analysis of the schematic column of Figure 6.13 obviously applies also to
other types of pin-ended fittings (like spherical ends or roller and ball bearings).
The approximate solution for the elastic buckling load with rigid ends is
2 EI
Pcr D [1 C 1/32a/L3 /22 ]2 . 6.2
L2
Hence, the “adjusted length” of the column Lad , that includes the effect of the rigid
ends, is
Lad D L/[1 C 2 /122a/L3 ]. 6.3
Some typical values are given in Table 6.1.
Buckling and Postbuckling of Columns 305

Figure 6.12 Typical average axial stress-strain curves for tubular columns with small and large
imperfections (from [6.14])

One may note that even for large values of 2a/L, say 0.4, the elastic buckling
load would increase only by about 11 percent. For any practical proportions, say
2a/L < 0.2 the error would be less than 1.3 percent.
For short columns, beyond the elastic range, the stress-strain curve has to be
considered, from which the value of the tangent modulus Et can be obtained for
306 Columns, Beams and Frameworks

Figure 6.13 Simply-supported column with rigid knife end


fittings schematic (from [6.16])

Table 6.1 Effect of rigid ends in columns


Increase in Pcr Decrease in ‘‘adjusted length’’
(2a/L) (percent) L  Lad /L (percent)
0.1 0.16 0.08
0.2 1.32 0.66
0.3 4.49 2.2
0.4 10.8 5.0
0.5 21.7 9.3

any value of the compression stress. From Eq. (6.2) the buckling stress is

 2 Et
cr D [1 C 1/32a/L3 /22 ] 6.4
L/2
where  is the radius of gyration of the cross-section of the column and the elastic
modulus E has been replaced by the tangent modulus Et , to generalize the solution
for both elastic and plastic buckling. Chilver re-arranged Eq. (6.4) to

 Et /cr D L/f1 C 2 /12[2a/3 /L/3 ]g1 6.5
p
and then plotted  Et /cr versus L/ for three values of 2a/r D 5, 10 and
15 (see Figure 6.14). The last value of 2a/r D 15, which for a slenderness ratio
L/ D 20 corresponds to 2a/L D 0.75 or an unsupported length of L/4, is
probably an extreme practical condition. Usually 2a/ will not exceed 5, corre-
sponding to 2a/L D 0.25 for L/ D 20 or to 2a/L D 0.125 for a slenderness
ratio L/ D 40, and Chilver’s curves in Figure 6.14 show that then the effect of
Buckling and Postbuckling of Columns 307

p
Figure 6.14 The function  Et /cr for pin-ended columns with rigid end fittings versus the
slenderness ratio L/ (from [6.16])

the end fitting is negligible. His calculations for two typical wrought aluminum
alloys indicated that for short columns, where buckling is elasto-plastic, rigid end-
fittings increased the buckling load of a pin-ended column by not more than
5 percent for 2a/ D 15 and much less for the usual small values of 2a/,
being almost negligible for 2a/ < 10.
Hence a suitable design rule for column end fittings was suggested:

2a/ < 10/L/ 6.6

for negligible effect of end fitting rigidity. For example, with a slenderness ratio
of 40 it would be permissible to support rigidly a quarter of the total length of the
column, or L/8 at each end.
The second source of error identified by von Kármán, the possible rotational
restraint, was not amenable to similar simple analyses. The simple knife edges,
rollers or conical points, shown in Figures 6.6(a) 6.6(c), were, however over the
years, found to be rather material and load dependent and not very consistent.
Hence also for small columns, roller or ball bearings are nowadays preferred.
A simple undergraduate student column experiment, carried out routinely at the
Technion in Haifa, can serve as an example of this trend. A couple of decades ago
the simply supported column was represented by a simple knife edge end fitting,
as shown in Figure 6.15. There, two bolts A pressed on two small rectangular steel
plates B, which slid on pins C, to fix the column D (of rectangular cross-section)
in the steel end fitting. If the test column had not been clamped centrally, which
308 Columns, Beams and Frameworks

Figure 6.15 Early knife-edge end fittings for student column experiments at Technion Israel
Institute of Technology

became evident in the strain gage readings as the load was applied, its position
could be adjusted by just loosening and tightening the bolts A, even under load.
The knife edge of this simple end fitting, however, lost its edge after some use and
a small and undefined rotational constraint appeared.
In order to make the tests more consistent, especially as the study of the correla-
tion between vibrations and buckling was added to the column experiment, a new
end fitting with roller bearings has been introduced, Figure 6.16. Here the column

Figure 6.16 Modern roller-bearing end fittings replacing the earlier knife-edge fittings of
Figure 6.15 for student column experiments at the Technion
Crippling Strength 309

D rests between two wedge-shape jaws B, which fix it in the rotating body C of the
end fitting. The compressive load tightens the grip of the jaws, due to the wedge
shape of the cavity in which they fit. The body C rotates in two ball bearings E
and therefore the end fitting represents a good pin-end. The holding bracket F of
the end fitting rests on the platens of the testing machine, or is attached to it with
suitable bolts.
Though slightly more complicated than its knife edge predecessor, the ball-
bearing end fitting has proven itself as a consistent, inexpensive and convenient
end fixture, and has presented no problems to the students.

6.2 Crippling Strength

6.2.1 Crippling Failure

When a short column with a solid cross-section is compressed, it will usually


fail by compressive yielding of the material, followed by squashing or a shear
failure, if the column is very short. The design stress, which limits the buckling as
the length of the column diminishes, is therefore the compressive yield stress. If,
however, the cross section of the column is thin-walled, the yielding is replaced by
local buckling of the thin-walled, flange-plate elements of the column, which can
also occur elastically. Tests show that often, after such local buckling (sometimes
called wrinkling) has occurred, the column still has the ability to carry a greater load
before it fails. Local buckling and local failure loads are therefore not the same.
Figure 2.11, in Chapter 2 shows the stress distribution for a typical channel
section after local buckling has occurred, but prior to failure. As the load is
increased, the local buckles on the flat sections grow, but most of the increasing
load is transferred to the much stiffer corner regions, until the stress intensity
reaches a high enough value to cause excessive deformation and failure called
“crippling” (see for example Chapter C7 of [2.78]). When local buckling occurs
at relatively low stress levels, the crippling stresses will be significantly higher.
But, as already pointed out in Section 2.1.3, when local buckling takes place at
high mean stress levels (say 0.7 0.8cy ), the buckling and crippling stresses are
practically the same. In both cases, however, the crippling stress (also sometimes
referred to as crushing stress or “maximum average stress”) replaces the yield stress
as the limiting design stress for short columns with thin-walled cross-sections.
It was also pointed out in Chapter 2, that in the absence of satisfactory analytical
solutions (partly because the manner in which stresses build up in the corner regions
is not well understood), the crippling stress has to be calculated by semi-empirical
methods. Since the crippling strength is one of the most basic data for air-frame
design, very extensive tests were carried out in the forties and fifties to establish the
required data base (for example [6.17] [6.21]). Based on these tests (usually carried
out in standard testing machines, with flat-ended specimens bearing directly on the
guided platens, assumed to simulate clamped ends) empirical and semi-empirical
methods of crippling stress prediction were derived. The methods of Crockett,
310 Columns, Beams and Frameworks

summing the crippling loads of elements including the curved junctions [6.17], and
of Needham, the angle method [6.21], were widely adopted by industry; but the
method proposed in 1958 by Gerard [6.22], based on a careful review of previous
work and a comprehensive semi-empirical investigation, was more general and
was therefore widely accepted.

6.2.2 Gerard’s Method for Calculation of Crippling Stresses


Gerard proposed the following formula for the crippling stress cc :
    m
cc gt2 E 1/2
Dˇ 6.7
cy A cy

where cy is the compressive yield stress, A is the cross-sectional area, t the
thickness and g is the number of imaginary cuts needed to divide the cross-section
into a series of flanges plus the number of flanges that would exist after the cuts are
made. The parameters ˇ and m are empirical constants determined from test data.
Gerard based the derivation of Eq. (6.7) on the fact that the failure stress of
plates after buckling depends strongly on the stresses along the supported unloaded
edges. Thus in curve fitting Eq. (6.7) to the available experimental data, one must
differentiate between cross-sectional shapes where the unloaded edges are free
to warp in the plane (such as angles, plates supported in V-grooves and square
tubes) and cross-sectional shapes with straight unloaded edges (such as T-sections,
cruciforms and H-sections). Ways to determine g for typical cross-sections and
values for the corresponding parameters ˇ and m are summarized in Figures 6.17a
and 6.17b, respectively. Up to now all the material effects are included in the
parameter E/cy . Ways to account for the strain hardening effects in the corners
of formed sections, or for the use of a cladding correction factor for sections made
out of clad aluminum-alloy sheet, are discussed in Gerard’s paper [6.22] and on
pp. 477 479 of Rivello’s textbook [2.10].
Finally, it was recommended that the cut-off or maximum crippling stress for
thin-walled cross-sections should be limited to the values summarized in Table 6.2,
unless the use of higher crippling stresses could be supported by appropriate results.
The use of Gerard’s method to calculate the crippling stress of columns with
thin-walled composite cross-sections is illustrated in Bruhn’s textbook [2.78] by
numerous worked out examples.

Table 6.2 Cut-off or maximum crippling stresses for


different cross-sections (from [2.78])
Type of sections Max. cr
Angles 0.7 cy
V-groove plates 1.0 cy
Multi-corner sections, including tubes 0.8 cy
Tee, Cruciform and H-sections 0.8 cy
2-corner sections, Z-, J- sections, channels 0.9 cy
Crippling Strength 311

Figure 6.17 Method of cutting simple elements for determining the empirical constant g in
Gerard’s formula for the crippling stress (from [6.22])

6.2.3 Crippling Strength Tests


Though Gerard’s method and the other semi-empirical methods for crippling stress
prediction were derived in the fifties, they are still in use today. Because of
the semi-empirical nature of these methods and their material dependence, many
additional crippling strength tests have since been performed, in particular when-
ever new structural configurations or materials were introduced (see for example
[6.23] [6.31]).
For example, when a trapezoidal corrugated plate was considered as a compres-
sion element at the Technical University of Münich, Germany, in the seventies,
a series of crippling strength experiments were carried out (see [6.23]). Short
stub aluminum alloy columns (or rather short corrugated plates), with cast and
subsequently machined top and bottom epoxy end beams (see Figure 6.18a) were
carefully tested for local buckling (see Figure 6.18b) and then for crippling failure
(see Figure 6.18c).
Or, with the introduction of composite laminates in the US aerospace industry,
a series of crippling tests on graphite/epoxy columns and plates were performed
at the Convair Division of General Dynamics in San Diego, California in the late
312
Columns, Beams and Frameworks

(a)

Figure 6.18 TU Münich crippling tests on short corrugated aluminum alloy plates (from [6.23]): (a) dimensions of specimens, (b) local buckling
just prior to failure specimen 3B, (c) crippling failure specimen 5C
Crippling Strength 313

seventies. A crippling fixture and experimental procedure presented in [6.25] was


used and flat plate specimens (different laminates), as well as square tubes and
I-sections, were tested. The no-edge-free flat plate compression specimens were
supported along the unloaded edges by V-grooves in steel blocks and the one-
edge-free tests were carried out with just one V-block support. The loaded edges
in both cases were potted with epoxy in aluminum alloy blocks. The square-tube
and I-section specimens were potted with cerrobend in aluminum end blocks and
then placed between the platens of a universal testing machine. Load-displacement
plots like Figure 6.19, usually showed that the crippling load Pcc considerably
exceeded that of incipient buckling Pcri , which again was usually slightly above
the theoretical elastic buckling load Pcrth . Sometimes the crippling failure was
accompanied by severe delamination (common in composite laminate failures, as
discussed in Chapter 14, Volume 2). Empirical crippling curves, based on a nondi-
mensional crippling equation similar to that proposed by Gerard for thin-walled
metal columns Eq. (6.7), were obtained. It was also found that the no-free-edge
empirical crippling curves could be used to predict the crippling strength of square
tubes.
The wide-spread use of thin-walled cold-formed and welded columns in the
eighties (discussed in Sub-section 6.2.5) also motivated many crippling tests in
civil engineering studies, though under the label of stub-column tests. For example,
at the Structural Stability Laboratory of the University of Liége, Belgium, extensive
stub-column experiments were carried out on steel columns with thin-walled open
profiles (see [6.26] [6.30]). Large series of U,C and angle sections were tested

Figure 6.19 Convair/GD crippling tests on graphite/epoxy columns and plates typical load-
displacement plot for a one-edge-free specimen (from [6.24])
314 Columns, Beams and Frameworks

Figure 6.20 TU Aachen crippling tests on thin-walled A Li extended columns crippling


failure of a short specimen (of 5 cm length), crippling of angle section near the
semi-circle portion of the cross section (from [6.31])

in a universal testing machine, with fixed end fittings. The strain hardening in the
corners, typical of crippling tests, was taken into account as well as the increased
warping rigidity and warping restraints important in angle-section stub columns.
The results of the comprehensive tests on all the open profile stub columns indicated
that caution must be exercised in applying current design rules to columns with
such sections.
Another more recent example are the crippling tests on aluminum-lithium alloy
extruded stringers carried out at the Institut für Leichtbau of the Technical Univer-
sity, Aachen in 1992 [6.31]. The compression experiments on the thin-walled
bulb type (semi-circle angle cross-section) extruded A Li stringers DAN 50-
13 included column local buckling and postbuckling tests as well as short column
crippling tests. One of these is shown in Figure 6.20, where failure was by crip-
pling of the angle section near the semi-circle. At failure, these A Li specimens
exhibited long cracks, which apparently were due to brittleness of the material or
some layered structure caused by the extrusion process.

6.2.4 Crinkly Collapse

The local postbuckling behavior of the corners of thin-walled columns attracted the
attention of many investigators over the years, since it was not fully understood.
In many thin-walled columns failure by a localized buckling mode, involving the
Crippling Strength 315

collapse of the corners, was observed and was called “crinkly collapse”. There was
some similarity between this phenomenon and the crippling failure discussed in
the previous section.
It was found that this “crinkly collapse” type of failure was initiated by a combi-
nation of geometric instability and material plasticity, presenting a difficult problem
of elasto-plastic analysis. Since it appeared that the elastic collapse was similar to
the elasto-plastic one, an analysis of the simpler elastic problem was initiated
at the University of Southampton in the late seventies. Simultaneously an experi-
mental study on silicone rubber model columns was carried out there [6.32], which
demonstrated the localized crinkly collapse mode, that was initiated by geometric
elastic instability at the corners associated with their waviness prior to failure. The
curing silicone rubber (commercially available) was chosen as the model material
to permit the large elastic deformations required for an elastic “crinkly collapse”.
Special silicone rubber sheets were cast, cut and bonded to form square section
tubes. The ends of these specimens were plugged with wood and machined perpen-
dicular to the column axis. The models were tested in an Instron testing machine,
which provided automated plotting of load and compressive strain.
With increasing compression, local buckles appeared at an early stage. Then, as
the amplitude of the buckles grew, a gradual decrease in stiffness occurred and
noticeable curvature of the column corners appeared (see Figure 6.21a). Collapse
was accompanied by a sudden drop in load and by the appearance of two crin-
kles, either in opposite or adjacent corners of the specimens (see Figure 6.21b).
Upon unloading, the specimens jumped back to the original buckled configuration,
at a somewhat lower strain, and demonstrated elastic behavior (see Figure 6.22).
Four nominally identical specimens were tested to investigate experimental repro-
ducibility. Their overall stress-strain curves shown in Figure 6.22 indeed demon-
strated excellent reproducibility. Experiments with columns of different length
showed that the stiffening effect of the end supports (not accounted for in the
analysis), which effectively clamp and locally stiffen the column with the wooden
plugs, caused a significant length dependence of the test results. The silicone
rubber modeling indeed served well to exhibit the large deflection of the “crinkly
collapse”.
Before leaving the topic of crippling collapse of columns, it may be appropriate
to mention a related phenomenon, web crippling, which occurs in thin-walled
beams subjected to concentrated or patch loads and which is discussed in Chapter 8.

6.2.5 Thin-Walled Cold-Formed and Welded Columns

Though not directly related to crippling, the buckling and postbuckling behavior of
thin-walled cold-formed and welded columns has some characteristics that make
their discussion appropriate here.
As already mentioned in Chapter 1 and in Sub-section 6.2.3, there has been a
significant increase in the use of thin-walled cold-formed and welded columns
in the last decade. This has been accompanied by an extensive research effort.
For the light, thin-web welded columns, the focus was on the residual stress
316 Columns, Beams and Frameworks

Figure 6.21 Southampton University silicone rubber columns (square tubes) demonstrating
crinkly collapse (from [6.32]): (a) noticeable curvature at corners, (b) appearance
of corner crinkles, (c) crinkly-cum-overall buckling

patterns (see for example [1.30]), which were found to be different and larger than
those commonly experienced with thick-walled or rolled sections. These significant
residual stresses and the early local plate buckling of flanges and webs resulted in
a nonlinear behavior of the columns, that required comprehensive tests in order to
provide design data. Similar nonlinear behavior also occurred in thin-walled cold-
formed columns due to early plate buckling, and therefore they too warranted
Crippling Strength 317

Figure 6.22 Southampton University silicone rubber columns (square tubes) demonstrating
crinkly collapse overall stress-strain curves of four nominally identical speci-
mens, showing the experimental reproducibility (from [6.29])

extensive experiments (see for example [6.33] [6.35]). It may be pointed out
that in the more recent tests careful geometrical imperfection measurements were
performed on all specimens, both in cold-formed and welded columns, and the
effect of these imperfections was then evaluated (see for example [6.35]).
For short columns, these tests resembled the crippling tests (discussed in Sub-
section 6.2.3) and the observed interactive buckling failures involved only local
buckling and overall bending modes. For longer columns, interaction with torsional
buckling modes is also possible (see for example [6.36]), and the long column
tests are therefore discussed separately in the next section, in connection with
torsional-flexural buckling.
It may also be mentioned that in the last two decades numerical simulations,
which take into account the real geometrical imperfections and material inho-
mogeneities, have often been employed to supplement the results obtained from
column buckling tests, in order to reduce the experimental work needed for deter-
mination of buckling curves and design data. Such numerical simulations have,
for example, been used for rolled high strength steel round tubes and rectangular
tubes built-up from welded cold-formed high strength steel channels in a joint study
carried out at Milan Polytechnic and Liége University in the late seventies [6.37].
This numerical simulation could, however, not be applied to the cold-formed square
tubes of the same program, since their residual stresses were biaxial (and rather
difficult to measure), whereas the numerical simulation took into account only
uniaxial residual stresses. Many more tests were therefore required for the buck-
ling strength data of these cold-formed square tubes. Note that the investigators
318 Columns, Beams and Frameworks

could take advantage of numerical simulation in their column buckling experi-


ments, only when the basic observed phenomena were well defined and amenable
to precise measurement.
As an example of a typical modern laboratory type setup for compression tests
on small thin-walled columns, that was recently used for experiments on cold-
formed sections at the University of Strathclyde in Glasgow, Scotland [6.38], is
now discussed. These tests were carried out under carefully controlled loading and
boundary conditions, in order to accurately assess the failure predictions of the
recent British code for cold-formed steel sections, BS 5950 Part 5, and to examine
other aspects of column behavior, such as growth and shape of local buckling and
postbuckling deformations.
Uniform compression of all the plate elements of the column was considered to
be of particular importance. The two generally used approaches, loading through
flat parallel platens or through platens attached to ball joints, were therefore thought
to be insufficient. With flat parallel platens any out-of-flatness or skew of the spec-
imen ends could result, because of the very small end displacements involved,
in substantial load concentrations on a single element of the section or even at
a single location. With platens attached to ball joints, they could align to elimi-
nate load concentration, but the compression might still be nonuniform due to the
“wandering centroid” phenomenon. Hence the Strathclyde University test setup
was designed to overcome these loading uniformity problems.
The relatively simple test rig, shown in Figure. 6.23, was built to be used on
a Tinius Olsen testing machine. It consisted of two platens, a top platen (2 in.
Figure. 6.23) attached to the crosshead of the testing machine, and a bottom one
(5) mounted on a ball table (7), which in turn was placed on the Tinius Olsen
platform (8). Four leveling jacks (6) were provided for holding the bottom platen
in position during the tests. For adjustment of the test rig prior to testing, the
platens were connected by adjustable screw tie rods (not shown in the figure),
which were then removed for the actual experiment.
The procedure was as follows: First the top platen, the test specimen and the
bottom of the platen were joined and held in place using the adjustable tie rods.
This assembly was then placed on the ball table (7), positioned on the platform of
the testing machine (8) and secured to the machine crosshead (1). A small preload
was applied and the leveling jacks (6) were adjusted to give equal bearing forces at
the support. The jacks were then locked in position, ensuring that the specimen was
properly seated and that the compression would be uniform. Finally the preload
was removed and the tie rods were dismantled.
A deflection measuring device (not shown in the figure), consisting of a stiff light
framework, that supported a system of tubular guides along which an LVDT could
move, was attached to the bottom platen. This transducer measured the out-of-
plane deflection of the plate elements of the column, while another LVDT attached
to the framework measured the position of the deflection LVDT across the section.
The deflection transducer could be moved across a plate element remotely, with
deflections and positions being automatically recorded on an X Y plotter. During
the tests, the displacements were thus measured along the central horizontal line
Crippling Strength 319

Figure 6.23 Strathclyde University compression tests on thin-walled cold-formed sections test
rig assembly (from [6.38]): 1. Tinius Olsen cross head 2. top platen 3. end plate
4. test specimen 5. bottom platen 6. leveling jack 7. ball table 8. Tinius Olsen
platform 9. end plate glued to specimen with Araldite glue

of the main plate element (except in two specimens where this line coincided with
the nodal line). Load-displacement records were taken with the usual test machine
equipment. Additional dial gages measured the relative displacement of top and
bottom platens, to check the uniformity of compression. Centrally located pairs
of strain gages were attached to the main plate of four of the six specimens, for
experimental determination of the critical loads. On one specimen, additional four
pairs of strain gages were added at the section corners to examine the uniformity
of strains over the cross-section. More extensive application of strain gages would
have probably enhanced the test results, as would have further LVDT measurements
along additional horizontal lines.
The six specimens were made from sheet steel by cold-forming on a brake press.
All the specimens were manufactured from the same sheet of steel and therefore
had practically the same thickness and material properties. Though the (b/t) ratios
of their main plate elements were nearly identical b/t D 153 155, each column
section was different, from C channel to lipped channel, trapezoidal channel and
320 Columns, Beams and Frameworks

one lipped channel closed by a plate spot welded to its lips. Duplication or tripli-
cation of each shape would have probably given more weight to the experimental
conclusions.
In order to prevent lateral movement of the loaded ends of the specimens during
the tests, while offering only little restraint to out-of-plane rotation of the plate
elements, these aluminum end plates were glued to both ends of the columns
with Araldite glue, as shown in Figure 6.23 (Section BB). Since the glue prevents
translation of the specimen ends, while permitting angular movement of the plate
elements, the ends were considered to be a close approximation to simple supports.
The failure load predictions of BS 5950 Part 5 were found to be in good agree-
ment with experimental results. All the tests continued far into the plastic unloading
range.
One should note that though relatively simple, this test rig provided the means
for the required controlled uniform compression loading.

6.3 Torsional-Flexural and Distortional Buckling


6.3.1 Torsional Buckling

Whereas crippling relates to the collapse behavior of very short thin-walled


columns, torsional buckling is a possible mode of buckling and failure for thin-
walled columns of medium length, whose torsional stiffness is relatively low, as
already pointed out in Section 2.1.3 of Chapter 2.
The phenomenon of torsional buckling and torsional-flexural buckling of
columns was recognized in the twenties, when the first all-metal airplanes were
designed and built. Open-section columns, such as channels, zees or angles, were
then widely used in aircraft design, since they were easily connected and could be
conveniently inspected, but due to their small torsional rigidity they were prone
to torsional failure. Wagner in Germany was the first to present in 1929 a theory
for arbitrary thin-walled sections [6.39]. His work was further developed in 1937
by Kappus in Germany [6.40] and by Lundquist and Fligg in the US [6.41]. The
problem was extensively studied in the thirties and forties by many investigators
(see for example [6.42] [6.44]), and since the forties and fifties it has been an
important topic in most relevant text books (see for example [2.1], [2.3], [4.18],
[4.19], [6.45] [6.47]).

6.3.2 Torsional-Flexural Buckling Tests

The earliest experimental study of torsional buckling was that carried out in 1934
by Wagner and Pretschner on plain and flanged aluminum alloy angles [6.48].
Significant experimental data on torsional buckling was also provided by a series
of tests on about 500 equal angle section steel and aluminum alloy columns, made
by Kollbrunner in Zürich in 1935.
Torsional-Flexural and Distortional Buckling 321

Another series of tests on 33 aluminum alloy channels, carried out for NACA
by Niles at Stanford University in the late thirties ([6.50], or see [6.45], pp. 316,
351 355), set the pattern for torsional buckling experiments on columns in the
following decades. The columns were loaded by an hydraulic jack in a universal
testing machine, through special sophisticated end fittings. These were designed
to apply the resultant load through the centroids of the end cross-sections, while
permitting free warping of the three main elements of the column cross-section,
except at their mid-points. This was achieved by a system of three knife edges,
one acting at the mid-point of each of the three elements, which supported three
bearing blocks, beveled to allow a š 5 degree uninterfered rotation about the knife
edge, which thus ensured the desired free warping.
Rotations and translations of cross-sections were measured with the aid of
“antennas”, constructed from round steel rods attached at different heights in the
center of the back of the specimens (see Figure 6.24). As shown in the figure,
the movement of each antenna (1) was determined by measuring the distances
from reference points (2) about 0.5 inch from the end of each of its arms to fixed
reference points (3) on the wooden scaffolding (4). As can be seen, the distances
between the reference points were measured by ordinary vernier calipers (5) with
special lozenge-shaped attachments (6) on their jaws. Though rather unsophisti-
cated (today probably LVDTs or optical means would have been employed), the
accuracy of the rotation measurements was quite good, with an error of less than
0.04 degrees, but the measurements of the translational movements of the cross-
sections were less satisfactory. The change in length of the specimens under load

Figure 6.24 Stanford University experiments of torsional column failure: “antennas” for
indication of rotation of cross sections and the positioning of calipers for
measurement of distances from reference points on the scaffolding (from [6.50]):
1. “antenna”, 2. reference point, 3. fixed reference point, 4. wooden scaffolding,
5. vernier caliper, 6. lozenge-shaped attachment
322 Columns, Beams and Frameworks

was measured by dial gages attached to the end fittings, as was and still is customary
in many tests.
A very detailed description of the test procedure was presented, including
enumeration of checks to reduce the human errors, as well as a detailed description
of the calibration of the Bourdon pressure gage of the jack for load measurements.
The precision of the results was also carefully examined. These discussions
presenting important experimental details are worth reading even today.
The critical (failure) loads for all the columns are plotted in Figure 6.25 against
their length, and are compared with theoretical predictions based on the analysis of
Lundquist and Fligg [6.41], shown as curve Pth . The Euler load PE , also depicted
in the figure, emphasizes the importance of torsional buckling as a failure mode
that can result in much lower critical loads. It may be noticed that for lengths
in excess of 24 in., when failure is by torsional buckling, the agreement between
prediction and experiment is very good.
It may be noted that, contrary to the sophisticated free warping ends of the
Stanford University tests [6.50], most other torsional or torsional-flexural buckling
experiments of the forties mentioned below, used simpler “flat ends”. For example
in Ramberg and Levy’s tests [6.54], the ends of their 125 extrusions were ground
flat and perpendicular to the axis and the specimens were compressed in a testing
machine between ground steel blocks. To obtain uniform loading, a plaster of
Paris cap was placed between the top steel block and the head of testing machine.

Figure 6.25 Stanford University experiments of torsional column failure: critical loads against
length comparison of experimental buckling loads with theoretical prediction Pth
and Euler loads PE (from [6.50])
Torsional-Flexural and Distortional Buckling 323

The distribution of strain was measured with Tuckerman mechanical strain gages
(this was before electric strain gages were available) and when at low loads strain
divergences exceeded 10 percent, the column was reground and retested. The twist
was measured only at the center of the specimen, but in some of the tests quite
accurately, by optically measuring the relative rotation of two prisms with an
autocollimator.
In the late thirties and forties further extensive experiments were carried out
to lend support to the theory and design methods for torsional and flexural-
torsional buckling of columns and stringers attached to sheets, primarily for aircraft
structures. For example, at NACA and at the US National Bureau of Standards,
aluminum-alloy panels stiffened by Z-, S-, C- and U-section stiffeners were tested
([4.24], [6.51] and [6.52]), as well as similar panels stiffened by bulb angles [6.53].
Later many tests were also performed on extruded sections made of aluminum
and magnesium alloy [6.54], which included also inelastic buckling. At about
the same time, a series of tests on folded mild steel and aluminum alloy angle
section columns, was carried out at Battersea Polytechnic, London [6.55], for civil
engineering applications.
By the fifties, the aeronautical engineers seemed to have a sufficient data base
for design against torsional and torsional-flexural buckling. But the growing use
of thin-walled open sections as load carrying structural members in other fields of
engineering, in the following decades, motivated extensive studies of torsional
and torsional-flexural buckling, primarily by civil engineers (see for example
[6.56] [6.58]). As a typical example of the experimental investigations involved,
one can consider the series of tests on columns with different cross-sections
performed at Cornell University in the mid-sixties [6.59], or the tests on buckling
of steel angle and tee struts, carried out a few years later at the University of
Windsor in Ontario, Canada, [6.60].
The Cornell experiments (see [6.59]) included columns with lipped and plain
angles, channels and hat sections, all tested with fully restrained ends. Fixed ends
were preferred at Cornell, to avoid the complex end fitting required for accurate
simple supports. The test setup is shown in Figure 6.26. The ends of the test
column, to which steel end plates were welded was set in hydrostone (a type of
quick-setting cement), which served two distinct purposes. While it was wet and
plastic, the column could be tilted and brought into vertical alignment. After it had
dried and hardened the hydrostone served as a means of evenly distributing the
load from the testing machine to the specimen. At Cornell at the time, any further
refinement in the alignment procedure was considered superfluous, in view of the
shape imperfections inherent in any cold-formed member. Today, the imperfections
would be measured and their effect be compensated by some means of adjustment
of the alignment during the test.
The twist and lateral displacements of the column tested were measured by a
simple set of pointers, scales and two transit theodolites, which provided fairly
accurate measurements of twist and local distortion.
One may note that the Southwell plots (discussed in detail in Chapter 4,
Section 4) employed here, again yielded good agreement between experimentally
324 Columns, Beams and Frameworks

Figure 6.26 Cornell University torsional-flexural buckling tests on “flat end” steel columns:
test setup, schematic (from [6.59]). The hydrostone assists alignment while wet
and plastic, and helps to even load distribution after hardening

observed buckling loads and those predicted by linear theory. The plain and lipped
equal-legged angles and hat section columns had a stable postbuckling curve and
failed only at loads about 15 percent above the theoretical buckling load. This
postbuckling strength was attributed to the axial membrane stresses caused by the
large twisting deformations (20 40 degrees at the center of the column).
The Windsor tests set out to eliminate the limitations of the then current US AISC
and Canadian CSA specifications, which did not take torsional-flexural buckling
into account. The tests included 72 angle struts with both hinged and fixed end
conditions, and 27 T-sections with hinged ends.
In order to minimize end effects, the specimens were made as long as possible,
consistent with required slenderness ratios (22 to 114) and the capacity of the
available testing equipment. The angle struts were about 4 ft. (¾1.2 m) long and the
T-section struts 4 7 ft. (¾1.2 2.1 m). The test pieces were fabricated by regular
production processes (in line with the “as fabricated” specimen philosophy that
aims at providing reliable empirical data that can be used for design), and three
specimens of each configuration were tested.
Torsional-Flexural and Distortional Buckling 325

The test setup was a regular horizontal test frame, comprising two nearly 10 ft.
(3 m) long 12 in. channels placed back to back, 14 in. (356 mm) apart, and bolted
to a 1/4 in. (6.35 mm) thick steel plate at the bottom, with three batten plates at the
top. The load was applied by a 120 kip (534 kN) capacity hydraulic jack through
a flat precalibrated load cell. Some columns required a larger, 200 kip (890 kN)
capacity jack. In placing the specimens in position, shims were used to ensure truly
axial load, and friction was minimized by lubricating the contact surfaces. Deflec-
tions were measured at quarter points of the column with dial gages mounted on a
separate rigid frame. For a few tests on the fixed-end angle struts, the deflections
were also measured at the ends to ensure that there was no significant end rotation.
The critical loads were determined by the “top-of-the-knee” method, developed
at NACA in the mid-forties for plates and discussed in Chapter 8, Section 8.3
(see Figure 8.57). There was good agreement between the experimental buckling
stresses and the theoretical ones, with an (Pexp /Pth ) range of 0.95 1.19 and an
average ratio of 1.05 with a standard deviation of 0.06. In the majority of cases
the actual modes of failure were also predicted correctly. Nearly all the single
angles failed by flexural buckling, more than half of the double angles failed by
torsional-flexural buckling with the remainder divided between inelastic flexural
or plate buckling, and all the T-struts failed by torsional-flexural buckling. Typical
torsional-flexural failures of double angles are shown in Figure 6.27 (reproduced

Figure 6.27 University of Windsor torsional-flexural buckling tests: typical torsional-flexural


failures of double angles (from [6.60])
326 Columns, Beams and Frameworks

from [6.60]). Practically always, the three specimens of each configuration failed
in the same mode. Calculations showed significant discrepancies between the theo-
retical predictions (that were verified by the tests) and those of the AISC and CSA
specifications, justifying their possible modification.
There have been many other experimental investigations of torsional-flexural
buckling of open section thin-walled columns in the last decades, as is evident for
example in the comprehensive 1982 review on buckling of angles by Kennedy and
Madugula [6.61]. The results of these studies have been incorporated in the various
design recomendations for thin-walled columns, as for example in Chapter 13 of the
SSRC Guide [6.3] or the ECCS Recommendations for Steel Constructions [6.5].
Before leaving the topic of torsional-flexural buckling, it may be of interest to
mention that the widely discussed 1978 collapse of the space-truss roof of the
Hartford Coliseum in the USA (see for example [6.62] or [6.63]) was originally
attributed to torsional buckling failure of its columns, with a four equal leg angle
cruciform cross-section. The disagreements about the cause of this collapse moti-
vated a review of the buckling analysis of four angle cruciform columns [6.64],
which however lacks experimental verification.

6.3.3 Distortional Buckling

The columns of industrial steel rack structures are usually cold-formed open
sections, most commonly thin-walled lipped channels. In experiments on these
channel columns, another failing mode distortional buckling was observed in
addition to flexural-torsional or local (plate) buckling (see [6.65] and [6.66]).
The distortional mode (sometimes called also local-torsional mode) is shown in
Figure 6.28. It involves a rotation of the flange and lip combinations (A), about the
flange/web junctions (B). The stiffness of the web element of the channel (C, often
called “front face” in rack columns), provides a restraint to this rotation, which
depends upon the slenderness of the web and the destabilization of the web caused
by the compressive stress present. Note that the distortional mode shown in the
figure does not involve any rotation of the whole cross-section which characterizes
torsional buckling.
The more recent series of tests [6.66] included 68 thin-walled steel channel
columns of four different section geometries, with dimensions commonly used in

Figure 6.28 Distortional mode of lipped channel columns (from


[6.65])
Torsional-Flexural and Distortional Buckling 327

the rack industry, and made of three different steels (see Figure 6.29). A finite
strip elastic buckling analysis indicated that the intermediate slenderness columns
of all four sections would fail mainly in the distortional mode, whereas the longest
specimens would fail by flexural-torsional buckling and the shortest ones by local
buckling. These buckling modes were indeed confirmed in the tests. The specimen
lengths ranged from 0.3 in. to 1.9 in. Their ends were milled to provide flat loading
surfaces.
The experiments were performed in a 250 kN capacity Instron TT-KM testing
machine, except for the thick short columns which were tested in a 2000 kN
capacity Avery testing machine. All specimens were tested with fixed ends. The
load was applied at the top end, through a rigid end platen fixed against rotation,
whereas the bottom end rested on a spherical bearing, which was restrained from
rotation about both horizontal axes and the vertical axis during the test (although
free to move for adjustments prior to loading). The purpose of the spherical bearing
was to ensure parallel ends prior to loading, in order to promote uniform loading
across the sections. Two of the specimens had 11 pairs of strain gages attached
at the mid-height position, to measure the stress distribution during loading, and

Figure 6.29 University of Sydney distortional buckling experiments on channel columns: test
sections (from [6.66]). Typical dimensions: t D 1.6 2.4 mm, bw D 76 85 mm,
bf or b1 D 30 80 mm: (a) simple lipped channel, (b) rack column upright,
(c) rack column upright with additional lip stiffeners, (d) hat
328 Columns, Beams and Frameworks

Figure 6.30 University of Sydney distortional buckling experiments on channel columns:


distortional buckling modes of test specimens (courtesy of Prof. G.J. Hancock)

it was indeed found fairly uniform till 0.85 0.95 of the maximum load. Typical
distortional buckling modes obtained in the tests are shown in Figure 6.30. More
recent experimental studies on distortional buckling by Hancock and his students
at the University of Sydney can be found in [6.176] and [6.177].

6.4 Lateral Buckling of Beams


6.4.1 Lateral instability of beams
Slender or thin-walled beams, subjected to bending loads in the plane of their
greatest flexural rigidity, can buckle by combined twist and lateral bending,
Lateral Buckling of Beams 329

called lateral buckling or lateral instability, as discussed earlier in Chapter 2,


Section 2.1.6. Due to the low torsional and lateral flexural stiffness of slender
beams, with narrow rectangular sections, I-sections with narrow flanges or thin-
walled open sections, their cross-sections in the center of the beam rotate and
deflect laterally, as in torsional-flexural instability caused by axial compression.
The moment of inertia in the plane of bending therefore decreases till the reduced
bending stiffness together with the torsional stiffness are insufficient to resist the
bending loads, leading to lateral buckling failure.
The phenomenon of lateral buckling was already known in the last decades of
the 19th century and had been observed in some tests on wrought iron and mild
steel beams (see for example [6.67]), but the first rigorous theoretical analyses
were those published simultaneously in 1899 by Prandtl [4.5] and Michell [6.68].
Extensive theoretical and experimental studies continued in the first half of this
century, motivated both by civil and aeronautical engineering applications (see for
example, [6.69] [6.78], [9.45], Volume 2, and Lee’s 1960 review [6.79]).

6.4.2 Prandtl’s Lateral Buckling Experiments

Prandtl’s theoretical analysis [4.5] was accompanied by careful experiments. These


1899 tests show, as do the 1910 von Kármán experiments discussed in Chapter 4,
the methodology of careful buckling experiments, with awareness of the essential
elements of a successful test already in the planning stage, as well as a realization
of many important factors that have worried experimenters in the 20th century and
even today still present pitfalls to the experimentalist.
Consider an example of Prandtl’s reasoning: Motivated by the possible use of
vibrations for determination of the stability of the beam, Prandtl sets out to mini-
mize friction in his test setup. He therefore chose a type of loading that promised the
smallest friction, a cantilever loaded by weights at the free ends. Then he demanded
very high rigidity of his clamping arrangement, but realized that it could not be
completely achieved in practice, since some small movements remained that gave
rise to small frictions, which were noticeable near the critical load in spite of their
small magnitude. Prandtl reported that these frictions “played many tricks” in his
early tests. They were found to be primarily due to the twisting of the cross-sections
during torsional displacements, that extended into the clamping block and resulted
in a “residual” torsional deformation at the root. He suggested that the incomplete
clamping could be compensated by appropriately increasing the effective length of
the beam, at least for elastic buckling.
Or consider the simple system of loading and of monitoring the deflections of
the free end. As shown in Figure 6.31, reproduced from Prandtl’s thesis, a scale
pan suspended from a curved bar (b), which rested on a knife edge, applied the
load and thus ensured that it was acting in the midplane of the beam. The position
of the free end of the beam at any moment was indicated by marks made on a
graph paper attached to a wooden plate by the two sharp corners S. The wooden
plate, which was moved by hand on three guides, parallel to the initial midplane of
the beam, was presented against the sharp corners to make the marks. From these
330 Columns, Beams and Frameworks

3
1

3 mm
b

39.9
l 5

Figure 6.31 Prandtl’s 1899 lateral buckling experiments: loading and deflection monitoring of
free end of cantilever beam (from [4.5])

marks the angle of twist could be easily found, since the distance between the
two corners was known. In order to follow the deflections of the free end as they
became large, the wooden plate sat on small metal knife edges that permitted it to
tilt as required. With this simple system Prandtl obtained consistent prebuckling
and initial postbuckling records and from them the buckling load.

6.4.3 Other Early Lateral Buckling Tests

Michell, in his 1899 paper [6.68], also included an experimental verification, or as


he put it “made the attempt to verify one or two of the results given by the theory”.
As a test beam he used a 4 ft long engineer’s steel straight-edge, from which the
feathered edge was removed by planing.
The elastic constants of the steel were determined by bending and twisting the
specimen itself. The bending deflections were measured by a screw-micrometer,
with an error of about 0.05 percent. The angles of twist were determined by setting
a small mirror fixed to the vertical circle of a theodolite normal to a line of sight
attached to the specimen, with an average error of 0.17 percent.
In order to eliminate the effects of the weight of the specimen, “counterpoises
(counterbalancing weights) were used to apply forces directed vertically upwards
at (11) points 10 cms. apart along the axis, each force being equal to the weight
of 10 cms. of the specimen. These forces, as well as the test loads, were applied
by means of steel hooks fitting in small double-counter-sunk holes drilled through
the specimen.”
“The specimen was adjusted before each experiment so that no lateral deflection
occurred with a moderate load. The test-load was then gradually increased until the
point of application would remain at rest in contact with either of two stops placed
Lateral Buckling of Beams 331

about 1 cm on each side of the initial position”, this being considered the critical
load. Since a marked sideways movement was observed in every case already with
test-loads 1 or 2 percent less than the critical one (the minimal load that would
maintain a deflection to either side), the experimental load was expected to be
slightly in excess of the true one.
“The specimen was tested as a cantilever (of 110 cms net length) in four posi-
tions, being twice inverted and once reversed end for end.” The mean measured
load was 2.9 percent above the calculated one. The specimen was then tested as a
simply supported beam with a single central load, with the beam being inverted for
the second test. The mean experimental buckling load was found to be 0.2 percent
less than the calculated one. Finally, “as a check on the methods and apparatus”
the specimen was tested as an Euler column, yielding a buckling load 2.2 percent
above the theoretical one. Michell concluded that “the chief source of error in
the experiments was the want of uniformity in the thickness of the specimen”
(being C1.4 to 1.6 percent of the mean), since the torsional and lateral bending
stiffnesses varied as the cube of the thickness.
One should note the relatively simple, yet accurate, test setup, the careful exper-
imental procedure and the lucid discussion of measurements and results of these
experiments, carried out nearly a century ago, which warrant their detailed presenta-
tion even today and justify the verbatim quotation of some paragraphs of Michell’s
paper.
The early civil engineering oriented lateral buckling investigations dealt primarily
with deep I-beams subjected to transverse loading. One of the most important of
these was the test series of 31 full size steel I-beams, carried out by Marburg at
the University of Pennsylvania in 1909 [6.69]. The beams were of 15 in., 24 in.
and 30 in. depth and of three shapes: American Standard I-beams, I-beams with
specially wide flanges rolled by the Bethlehem Steel Company, and Bethlehem
broad-flanged girder beams. The beams were tested with a minimum of lateral
support and in many cases failure was clearly by lateral buckling.
The next comprehensive investigation was performed by Moore at the University
of Illinois in 1910 1913 and included 40 full size I-beams [6.71]. Moore also
tabulated and assessed the available earlier test results. Moore’s tests were nearly
all made on a four-screw 200 000 lb Olsen testing machine with a long table for
beam tests. The method used for supporting unrestrained beams and preserving
freedom in respect to sidewise buckling (the term used for lateral buckling) is
shown in Figure 6.32. The unrestrained test beam in the figure rested on the long
weighing table of the testing machine via a sphere and plate bearing at one end
and a rocker bearing at the other end. The load was applied via sphere and plate
bearings and a roller bearing, at about 1/3 and 2/3 the span, with the rocker
and roller bearings permitting axial displacement. In two of the test series, the I-
beams were restrained against end twisting, by heavy angles at the ends, or against
sidewise buckling by fastening two beams together along their compression flanges
with batten plates.
Moore’s Illinois experiments provided considerable insight into the phenomenon
of lateral buckling, about which he stated, for example, that “the resistance of a
332 Columns, Beams and Frameworks

Figure 6.32 University of Illinois 1913 lateral buckling experiments: test setup showing
method of support and loading which permits unrestrained lateral buckling (from
[6.71])

beam against buckling depends on the stiffness of the beam and on the amount
of torsional fixity of the bearings”. They also yielded substantial empirical design
data and recommendations, whose influence was felt in many countries for decades
(see for example [6.75]).
Aeronautical engineers encountered lateral buckling in the twenties, when
new airfoil sections permitted the use of deeper spars. At MIT in Cambridge,
Massachusetts, a small series of tests on deep spruce spars, with depth to breadth
ratios of 2.5 17, were therefore performed in 1926 for NACA [6.74]. In these tests,
beyond a critical span or depth-breadth ratio, failures by lateral collapse occurred,
but at stresses below the elastic limit of the material, making repeated tests on a
single specimen possible.
A decade later, the lack of experimental results motivated NACA to sponsor
extensive tests on deep rectangular aluminum beams at the Aluminum Company of
America Research Laboratories [6.76]. The investigators, Dumont and Hill, pointed
out that the experiments were necessary to verify the apparently rigorous theoretical
elastic solution, in order to increase the confidence in it; but also to provide data
for an empirical extension of the elastic analysis to the plastic regime (for which
at the time no analytical solutions existed).
The study covered 26 deep rectangular beams, including also two of hollow
sections. The rectangular bars were tested in pairs, each pair being securely bolted
to channel spreaders, as can be seen in Figure 6.33, which shows the arrangements
of the ends. The spreaders held the ends of the bars vertically and prevented lateral
deflections at the ends. The central transverse load was applied via a loading beam
and rollers, and subjected the beams to a constant bending moment in their vertical
planes. The ends of the test beam pair rested on perpendicular stub I-beams, which
were attached to the base table, made of a pair of heavy channel beams.
In the following years, Dumont and Hill carried out further lateral buckling
experiments at ALCOA, for example a series of tests on equal-flanged Aluminum-
Alloy I-beams [6.77]. The specimens were extruded 27 ST aluminum alloy, rela-
tively light I-beams. They were again tested in pairs in a 40 000 lb. capacity Amsler
Lateral Buckling of Beams 333

Figure 6.33 ALCOA lateral buckling experiments of deep rectangular aluminum beams:
loading arrangement for a pair of specimens (from [6.76])

testing machine. The test setup was similar to that shown in Figure 6.33, except
that here the ends of each pair of beams were laterally restrained by means of rela-
tively rigid end restraining frames. The lighter beams and heavy end frames were
to achieve a high degree of end fixity, but complete lateral restraint of the ends of
the beams was not attained. Calculations based on the experimentally determined
critical stress indicated that the degree of end fixity varied between 96 percent for
the longest specimens to 81 percent for the shortest ones. It may be pointed out
that the difficulty to obtain complete end fixity, found in these tests, has since often
troubled experimenters in many types of test setups for various structural elements.
One should also note, that the Southwell method for determination of the crit-
ical loads, presented in Sections 4.5 and 4.6 of Chapter 4, was applied here, very
successfully, for the first time to lateral buckling.

6.4.4 Recent Lateral Buckling Investigations

By the middle of the century, lateral buckling had been recognized as an important
factor in the design of many types of structures. In the last decades it became
the subject of many studies, in particular on the influences of the conditions of
loading, end conditions and lateral restraint, of monosymmetry and of inelastic
buckling (see for example the reviews of Lee, Galambos, Trahair and Nethercot,
[6.79] [6.81] and [2.19] respectively). Today, lateral buckling also occupies a
significant chapter in texts on stability or design of structures (for example, [2.1],
[2.3], [4.11], [4.19], [6.47] and [6.82]). It has, however, also been realized that, due
to the different types of imperfections affecting the lateral buckling behavior, which
are not completely accounted for in the analyses, comprehensive experimental
data is required for reliable design methods. Extensive experimental studies have
therefore been carried out in the seventies and eighties (see for example the review
334 Columns, Beams and Frameworks

of Nethercot and Trahair [6.83], or of Fukumoto and Kubo [6.84]), many of them
focused on the effects of local concentrated loads.
A typical example of these are the lateral buckling tests carried out by
Fukumoto and his co-workers at Nagoya University in Japan in the early
eighties ([6.85] [6.87]). One of the series of experiments, that on lateral buckling
of welded continuous beams [6.87] is discussed in detail. The purpose of this study
was to determine the effect of an adjacent nonloaded span L2 on the lateral buckling
strength and deformation of the critical loaded span L1 (see Figure 6.34). Twenty-
one two-span continuous welded beams were tested under a midspan concentrated
load, as shown in the figure. Their buckling behavior was compared with that of
similar simple welded beams tested earlier ([6.85] and [6.86]), as well as with
calculated results.
The 21 mild steel beams had a nominal identical cross-section I-250 ð 100 ð
6 ð 8 mm, as shown in Figure 6.34, which was also identical to that of the earlier
simple welded beams. They were treated in seven groups of three, each group
having a different span length (3.5 5.5 m), with span ratios (L2 /L1 ) varying from
0.287 to 0.909. The beams were built-up from flanges and webs, made from
flame-cut plates, with single-run welds. A diagram of the test setup is shown in
Figure 6.35. The loaded span L1 was supported on simple supports and the contin-
uous extension L2 had a plate welded to its end, which was bolted with 18 bolts
to a fixed support C. The test beams were restrained at the supports A and B
against twisting about the longitudinal axis and lateral movement, but were free to
warp at support A. Longitudinal displacement was free at both A and B, whereas
support C provided fixed-end conditions vertically, laterally and torsionally. The
concentrated load was applied vertically at midspan of AB, through a Lehigh-type
gravitational load simulator, designed to eliminate any restraining effects of the
applied load. This gravity load simulator is discussed in Section 6.6.2 (see also
[6.118] or [6.120]).

Figure 6.34 Nagoya University lateral buckling tests on two-span continuous welded beams:
test specimen and buckling mode (from [6.87])
Lateral Buckling of Beams 335

(a)

(b)

Figure 6.35 Nagoya University lateral buckling tests on two-span continuous welded beams:
test setup: (a) diagram of test rig (from [6.87]), (b) view of a test (courtesy of
Prof. Y. Fukumoto)

The points of strain and deflection measurements on the specimens are shown in
Figure 6.36. The strain gages are located so as to separate in-plane and out-of-plane
strains and to experimentally yield the inflection points in AB. Initial geometrical
imperfections were carefully measured on all the beams and their initial lateral
deflections u0 and twist 0 at the centroid are shown in Figure 6.37 (where they
are drawn as if all groups had the same span length). The nondimensionalized
values of the initial lateral “crookedness” u0 of the continuous beams were found
to be slightly greater than those of the simple one-span beams, whereas the initial
twists were similar.
The mechanical properties of the original plates, from which the beam sections
were flame-cut, were determined from 30 tensile coupon tests. The longitudinal
336 Columns, Beams and Frameworks

Figure 6.36 Nagoya University lateral buckling tests on two-span continuous welded beams:
measurement points for strains and deflections (from [6.87])

Figure 6.37 Nagoya University lateral buckling tests on two-span continuous welded beams:
initial crookedness of test beams, lateral displacement u0 and twist 0 (from
[6.87])

residual stresses were measured by the sectioning method and found to be similar
to those obtained for the earlier simple beams. The curves of load versus measured
vertical strain for four arbitrary test beams were compared with computed values
and found to agree well; indicating that the loading and support conditions were
very close to the designed ones.
In Figure 6.38 (taken from [6.83]) the lateral buckling test results of the contin-
uous welded beams tested by Fukumoto et al. [6.87] are compared with those of
the earlier hot rolled continuous beams studied by Poowannachakai and Trahair
[6.88] and theoretical approximate curves for simply supported hot-rolled beams
with equal end moments. The results are presented in the figure as a nondimen-
sional test strength, the ratio of the failure
 load Pu to the full plastic collapse
load Pp , versus the modified slenderness, Pp /PE , where PE is the elastic buck-
ling transverse load. One can note that the Fukumoto et al. beams were relatively
slender and buckled primarily in the elastic range, where they agreed well with
earlier results. They hardly extended to the inelastic range, where the earlier results
were considerably below the predictions.
Lateral Buckling of Beams 337

Figure 6.38 Test results for lateral buckling of continuous beams: comparison of hot-rolled
beams [6.88] with welded beams [6.87] and approximate inelastic predictions
(from [6.81])

Another typical example are the lateral buckling experiments carried out by
Yura and his students and co-workers at the University of Texas at Austin in the
seventies and eighties (for example [6.89] [6.92]).
One of the special topics studied at Austin was the lateral buckling of coped
steel beams. In steel construction, beam flanges must often be notched out to
provide clearance for the supports when framing beams are at the same elevation
as the main girders, or when the bottom flanges of intersecting beams have to be
at the same elevation (see Figure 6.39). Such notches or cutouts are called copes.
They can be at the top, bottom or both flanges in combination with different types
of shear connections, as shown in the figure. Since theoretical studies [6.93] had
shown that coped connections could significantly reduce lateral buckling strength,
and only a few tests with one cope detail had been reported [6.94], a series of
experiments with different copes were carried out [6.92].
A single length of W12 ð 14 section beam was used for six elastic lateral-
torsional buckling tests on coped connections. The varying cope details, as shown
in Figure 6.40, were cut successively into the two ends of the beam, and test LTB1
with no cope served as reference. After each test, the beam was either rotated to
connect the other coped end to the stub column, or further coped while in place,
when feasible. Then a new test was performed.
In order to minimize the end restraint, and thus approach a pinned end condition,
12 in. ð 6 in. and 0.133 in. thick shear end plates were welded onto the test beams
and attached to the supports by four bolts. At each bolt location, two washers
were placed between the end plate and the supporting column, which reduced
the in-plane and lateral-end restraint of the beam. Standard bolt-hole clearances
were used.
338 Columns, Beams and Frameworks

Figure 6.39 Types of coped beam connections (from [6.91])

The test setup, shown schematically in Figures 6.41 and 6.42, was designed to
apply specified forces to the coped connection to be tested. The beam was loaded
upside down, with a reaction floor and wall system serving as a loading frame. One
coped beam end A was bolted to a heavy stub column, which itself was attached
with large bolts to the vertical reaction wall.
Load was applied upwards to the test beam at B, 114.3 in. (D 2.903 m) from the
face of the end plate, by a 60 ton hydraulic ram via a load cell. The load position
B was chosen so as to produce elastic buckling in the test span AB, while mini-
mizing restraint of the adjacent span. The hydraulic pressure was also monitored
to provide a second measure of the load. A roller assembly (see Figure 6.43) was
placed between the bottom flange of the beam and the ram to permit longitudinal
displacement. A tension load cell was placed at C, 108.9 in. (D 2.766 m) from the
center of the jack B, to measure the reaction. This load cell was connected to the test
beam and the floor beam by a bracket arrangement with pin joints to allow longi-
tudinal displacement also at C. The test beam was supported laterally at the load
and reaction locations, B and C respectively by out-of-plane bracing systems. These
included adjustable brace plates with slotted holes, which prevented lateral move-
ment but permitted vertical movement of the beam, as shown in Figure 6.42 for the
load location B. An additional adjustable lateral stop was placed near the midspan of
the test span (see Figure 6.41) to prevent large lateral movements (beyond 0.75 in.)
that could cause yielding on the compression flange of the test beam.
Lateral Buckling of Beams 339

Figure 6.40 University of Texas at Austin lateral buckling tests on coped beams: details of
coped connections (from [6.92])

The load applied by the hydraulic jack was determined by the load cell and
verified by two pressure transducers, one linked to a strain indicator and the second
connected to an X Y plotter. Six dial gages with 0.001 in. (D¾ 0.025 mm) intervals
were used to measure the in-plane deflection at the coped connection A, at the
load point B and at the reaction C (see Figure 6.43), and an inclinometer with a
0.00003 radian accuracy was used to measure the rotation of the coped connection.
The out-of-plane deflection instrumentation consisted of the simple device of a
string stretched parallel to the beam and a scale with graduations to 0.02 in. (D ¾
0.5 mm) placed at eight locations along the compression flange, for measurement
of lateral displacement. A potentiometer was also placed near the center of the test
span for monitoring of the load-lateral-deflection response with an X Y plotter. A
inclinometer was also placed near the middle of the test span, 7 in. (D ¾ 180 mm)
away from the potentiometer position, to measure the twist of the compression
flange.
Prior to the main tests, four connection restraint tests were carried out with
dead weights on 40 in. (D ¾ 102 cm) long W12 ð 14 beams, identical in section
and material to the test beams, and with the same size end plate connection to
340
Columns, Beams and Frameworks

Figure 6.41 University of Texas at Austin lateral buckling tests on coped beams: (a) schematic of test setup, (b) view of test setup (from [6.92])
Lateral Buckling of Beams 341

Figure 6.42 University of Texas at Austin lateral buckling tests on coped beams: schematic
of loading system and out-of-plane bracing (from [6.91])

Figure 6.43 University of Texas at Austin lateral buckling tests on coped beams: in-plane
deflection instrumentation (from [6.91])
342 Columns, Beams and Frameworks

Figure 6.44 University of Texas at Austin lateral buckling tests on coped beams: out-of-plane
deflection instrumentation (from [6.91])

the stub columns. From these preliminary tests, in-plane and out-of-plane moment
rotation curves were found for the connections with and without washers, showing
a much smaller in-plane stiffness (about 2.5 to 3.5 times) with washers, whereas
the out-of-plane rotation curves were almost the same with and without washers,
and about half the stiffness as the in-plane one with washers.
In the test procedure, before applying loads, the first loading stage data were
taken as the self-weight of the beam, which was supported by the connection to
the stub column and the far end reaction. Then loading was applied in increments
till about 85 percent of the buckling load, determined by the Southwell method in
order to avoid yielding due to large lateral deflections.
The test results of the lateral buckling investigation showed indeed that the
reduction of buckling strength due to coping could be very significant. They also
showed that the end restraints from connections and the restraints from adjacent
spans could significantly increase the buckling strength of coped beams, especially
for large copes.
The Southwell method, employed here and discussed in Chapter 4, Sections 4.5
and 4.6, usually yields a very good estimate of the experimental buckling load of
columns or beams without testing them to failure, by plotting the measured out-
of-plane deformation divided by the measured load versus that deformation. Two
typical Southwell plots for Test LTB4 are presented in Figure 6.45, (a) using the
measured lateral deflections, and (b) using the measured twists. The buckling load
Pcr , determined by a straight line passing through the data points (here, using the
least square method), was nearly the same for lateral deflection and twist. However,
when the coped depth increased, localized distortion of the coped region affected
the flange twist data and hence the lateral deflection data was preferred here for
the analysis of the test results.
Lateral Buckling of Beams 343

Figure 6.45 University of Texas at Austin lateral buckling tests on coped beams: typical South-
well plots, for Test LTB4, (from [6.91]): (a) employing lateral deflection data,
(b) using flange rotation data

In a typical test, one would plot the data as the load was gradually increased,
and (after disregarding the unreliable initial data points) would obtain estimates
for the buckling load during the test, “in real time”. One could therefore plan
where to terminate the experiment to remain non-destructive, even without good
theoretical predictions, at least in the case of columns, beams or frames. As pointed
out in Chapter 4, for reliable Southwell estimates the test load should approach
the buckling load, at least reach 80 85 percent of it. In the University of Texas
tests on coped beams the Pmax reached at least 87 percent of Pcr yielding accurate
estimates of Pcr .
One may note that Cheng and Yura also used other plotting techniques in their
work: the 1938 Lundquist plot [4.25] discussed in Chapter 4, and the method
proposed by Meck in 1977 [6.95]. However, all three methods yielded almost the
same results, and therefore the University of Texas coped beam lateral buckling
tests represent a good example of reconfirming the applicability and usefulness
of the Southwell method. Yura and his co-workers employed Southwell plots
successfully also in other studies (see for example [6.89], [6.90] and [6.96]).
344 Columns, Beams and Frameworks

It may be mentioned here that for beams another plot (which is essentially
also an adaptation of the Southwell plot) has been extensively used, the Massey
plot ([6.97] or [4.41]), and other methods have been proposed (see for example
[6.95] [6.99]). It appears, however, that the Southwell method is preferable.

6.5 Interactive Buckling in Columns and Beams


6.5.1 Mode Interaction and Early Studies
In Section 2.14 of Chapter 2, the importance of mode interaction, or modal
coupling, which can significantly reduce the collapse load of a built-up structure,
was demonstrated with the aid of van der Neut’s analysis of an idealized thin-walled
column [2.17].
The primary mode interaction for columns is that between the overall column
buckling (Euler buckling) in one half wave and the local plate buckling in shorter
waves. It was first studied by Bijlaard and Fisher at Cornell University in the early
fifties under the sponsorship of NACA, [6.100] and [6.101], and included also tests
on aluminum alloy drawn square tubes and extruded H-sections. Though carried
out over four decades ago, these Cornell experiments exhibited some ideas and
techniques, which were reported in detail, and are worth considering even today.
The specimens were carefully measured for deviations from flatness, straight-
ness, squareness, twist and thickness variations, which were generally found to
be well within tolerable limits. The compressive stress-strain characteristics of the
square tubes were measured directly on 8 in. long square tube specimens whose
walls were supported to prevent their premature local buckling. The walls were
supported by blocking inside and outside, such that unsupported portions of wall
had a width to thickness ratio b/t < 12.5.
The external blocking arrangement consisted of three square clamping frames,
which held four vertical steel supporting blocks, one against each face of tube. For
internal blocking, a special octagonal expanding fixture operable from the ends of
the specimens was used, which consisted of two semi-octagonal supporting blocks
and a screw driven wedge system. Since the range of expansion was about 1/4 in.,
auxiliary blocks had to be used for the larger tubes.
To prepare for a stress-strain test, the internal expander, slightly shorter than
the specimen, was inserted first with the necessary auxiliary blocking and centered
on the length of the tube. All block surfaces contacting the tube were lubricated
to avoid frictional restraint. Next the external blocking was applied. The steel
supporting blocks, also lubricated, were centered vertically and laterally, while
being supported on sponge-rubber pads and held in place by the center square
clamping frame. All blocking was then drawn up to the tube, a light seating load
was applied and the other two square clamping frames set in place. The appropriate
clamping pressure, selected in preliminary tests, was then applied.
Strains were measured with (the then relatively new) SR-4 electrical resistance
strain gages. In the stress-strain tests eight gages were used, two to a face, outside
the supporting blocks.
Interactive Buckling in Columns and Beams 345

It is of interest to note how centering the specimen and providing uniform stress
distribution, considered by Lundquist and Fisher to be “perhaps the most difficult
and persistent problem encountered”, was achieved. First, nearly perfect flatness of
ends was obtained by squaring and sanding of the sawed specimens on a disk sander
(today one would probably use a milling machine), followed by hand-lapping on a
surface plate with oil and emery. Then, in the testing machine, “use was made of
tissue paper shims, 0.0015 in. thick, slipped between the upper machine head and
the corners of a hardened-steel bearing block on the upper end of the specimen”
in order to correct for nonparallelism of ends and/or machine heads. Paper shims
were applied or relocated until strain readings on the eight gages showed a total
high-to-low deviation of less than 3 percent. Usually in the short column stress-
strain tests they could “hold this to less than 1.5 percent at each of three widely
separated loads in the elastic range”!
For the longer actual column test specimens, which were supported at the ends
by knife edges, centering was done with both strain readings and lateral deflection
readings. Eight Tuckerman optical strain gages were applied to the corners of the
column at two appropriate stations, and centering adjustments were determined
from these two sets of strain readings. “Final centering, at about two-thirds of the
predicted critical load, was done by adjusting the column ends until the lateral
deflections at mid-length and both quarter-points were negligible, and then making
a final check of strain distribution. Centering by deflection proved to be consider-
ably more accurate than strain readings for the final adjustments.” Differences in
the average strain on opposite faces of the column could be held to a maximum
of 1 2 percent.
Another interesting experimental point was the separate measurement of local
and overall column (Euler) buckling deformations. Bijlaard and Fisher felt that
electrical resistance strain gages on the column faces would have difficulty in
“sorting out” the proportionate effects of the two types of deformation. (Today
such a “sorting out” would probably be done by computer from data acquired with
a large number of small strain gages.) They therefore developed a mechanical local
buckling gage, consisting of a suspended blade in contact with the column face, two
accurate dial gages measuring the blade movement due to the buckle formation,
and a collar to fit on the column at a desired location and carry the gage elements,
without affecting the local buckling characteristics. This local buckling gage could
“ride” with the column during its primary deflection, and thus measure the net
local buckling deflections.
Both calculations and experiments in [6.100] indicated that for box sections and
common size I-, H- and channel sections the interaction effects were negligible,
but could be significant in sections prone to torsional buckling like T- or angle
shapes.

6.5.2 Interactive Buckling Experiments


The nonlinear interactive buckling phenomena have been widely studied
analytically and experimentally in the last two decades (see for example [6.12],
346 Columns, Beams and Frameworks

[6.34] [6.36], [6.102] [6.113]). Many theoretical studies dealt with the general
nonlinear problem of compound instability, which can arise in optimized structures
when local and overall buckling loads of the perfect structure are equal or
near-equal, and then the imperfection sensitivity is magnified (see for example
[6.102] or [6.103]). Most experimental studies, on the other hand, focused on
the practical problem of interactive buckling failures in slender section columns.
Typical examples were the tests on I-section columns, [6.34], [6.36] or [6.109],
on box sections, [6.12] or [6.107], or on channels and lipped channels, [6.104],
[6.110] or [6.111].
The test setups and techniques used in these investigations were in general
similar to those considered in the earlier sections of this chapter. Hence only some
particular aspects of typical interactive buckling tests will be discussed here.
At the University of Sydney, Australia, a comprehensive research program on
the interaction of local and overall column buckling of fabricated I-sections was
in progress in the mid-eighties. One typical group of tests in the program were the
compression tests on short and long welded high tensile steel I-sections carried out
by Davids and Hancock [6.108] and [6.36]. An I-section, whose flange width was
equal to the web depth and with web and flanges of the same nominal thickness,
was chosen for all the tests, because it appeared from previous theoretical studies
[6.106] to be rather sensitive to interaction buckling. These proportions were there-
fore expected to accentuate interaction effects, though they would usually not be
chosen in practical designs.
The purpose of the short column tests was to study the local buckling and
post-local buckling behavior of this I-section. A test length corresponding to a
nominal three local buckling half-wave-length was therefore chosen for all the test
specimens in the series, referred to as Series I. Nine specimens, three of each size,
were fabricated. One of each three was used for residual strain measurement by the
method of sectioning. The remaining two of each size were tested to destruction
in a DARTEC 2000 kN servo-controlled vertical universal testing machine. In the
test setup, shown in Figure 6.46, a pair of freely moving universal joints, with
spherical bearings and rigid end platens, were installed at the top, on the end of
the hydraulic actuator, and at the bottom, on the base bed of the testing machine.
Each specimen was centered on the end platens so as to distribute fabrication errors
uniformly about the geometric centroid.
One specimen of each size was fitted with strain gages around the perimeter at
column midlength. The, perhaps regrettable, economizing with strain gages was
probably due to the focusing on displacement measurements. These were performed
with linear displacement transducers (LDTs), supported from an instrumentation
frame, which measured the flange tip and web centerline deformations on all the
specimens tested (see Figure 6.46). The experimental loading was controlled by a
signal from the extensometer (a high resolution LDT) mounted on the lower bearing
and connected to the upper one by an axially stiff arm, as shown in the figure.
Geometric imperfections were measured by optical survey leveling of the spec-
imens (standard civil engineering practice) to an accuracy of 0.002 in. (0.05 mm)
per reading (with a Zeiss Koni 007M precise level). All component plates (flanges
Interactive Buckling in Columns and Beams 347

Figure 6.46 University of Sydney local and interaction buckling tests on short I-section
columns: test configuration (from [6.108])

and web) of the test specimens were divided into grids consisting of four equally
spaced lines on each flange outstand and five on each web, and 19 stations along
them, as shown for a flange in Figure 6.47a. The observed surface profile can be
decomposed into overall plate twisting, as shown in Figure 6.47b, and a net plate
ripple which is the local geometric imperfection. This local ripple measured rela-
tive to lines A,B,C,D of the twisted plane of Figure 6.47b is shown in Figure 6.47c
for one flange of a typical specimen. A cubic polynomial curve fitting program was
used to draw the curves in the figure. It should be noted that the local imperfec-
tions at the flange tip (line A) are significantly larger than those nearer to the web
junction. The local ripple component of the imperfection was then modeled as a
finite Fourier series. In Chapter 10, Volume 2, different techniques of reducing
geometric imperfection data are discussed in detail. Here the Fourier term of
the same half-wavelength as the predicted local buckling mode was averaged
for the local ripple component over the four flange tips and the web centerline,
yielding a maximum amplitude of these terms for the three sizes of specimens of
0.26 0.49 mm (D 0.05 0.10t, where t is the plate thickness).
All specimens in this series buckled with three local buckling halfwaves. Local
buckling was clearly observed visually and by the rapid changes in the linear
displacement transducer readings.
348 Columns, Beams and Frameworks

Figure 6.47 University of Sydney local buckling tests on short I-section columns measurement
of geometric imperfections (from [6.108]): (a) measuring points on one flange
outstand, (b) twisted plane, (c) variation of local ripple imperfections along flange
length
Interactive Buckling in Columns and Beams 349

The change in the axial stiffness SŁ (nondimensionalized by division with the


original stiffness), derived from measurements taken by the high resolution LDT
mounted on the lower bearing, was employed as an indication of buckling. As was
expounded in [6.108], for each particular column geometry a different value of the
axial stiffness ratio SŁ would define the point local buckling. Here SŁ D 0.65 was
the appropriate value.
The short column tests of Series I demonstrated that the specimens buckled in
the local mode predicted by the local buckling theory. The stage was therefore
set for the long column tests of the same I-section, denoted Series II [6.36]. The
specimens were fabricated from a similar high strength hot rolled steel plate as
the short columns and in a similar manner, though some variability of welding
technique between Series I and II was indicated by the residual strain gradients.
One 700 mm long specimen of each of the three cross-sectional sizes was fabricated
for residual strain measurement, again by the method of sectioning.
The local geometric imperfections of the Series II specimens were measured by
high precision optical leveling, as in the case of Series I, and the data was also
reduced in a similar manner. The magnitudes of the relevant Fourier term were
0.01 0.02 t, less than those in Series I. The overall geometric imperfections about
the minor flexural axis of the long columns were determined by measuring, with
a high precision optical level, the in-plane displacements relative to a line passing
through the ends of the specimens. A Fourier analysis was also carried out on
the overall geometric imperfections measured, and the magnitudes of the relevant
Fourier terms (of the same half-wavelength as the length of the column L), were
found to be less than 0.0002 L.
The long columns were tested in a horizontal reaction rig with a 2000 kN
capacity servo-controlled hydraulic ram, shown in Figure 6.48. The test rig could
take specimens up to 10 m long. The ends were pin-ends similar to those used
in the vertical setup for the short columns. The specimen was supported on rigid
end platens, which were mounted on spherical bearings. The end bearings allowed
rotations about two perpendicular axes located in the same plane by the use of
a shear box as can be seen in the detail of the end bearing (Figure 6.48b). Two
adjustable ball bearings were bolted onto the end plate mounted on the extended
ram (as shown in Figure 6.48d). These bearings moved longitudinally on either side
of a fixed rail (not shown in the plan view Figure 6.48a) and thereby eliminated
the possibility of a specimen failure in the flexural-torsional buckling mode.
A transducer was mounted on a plate supported at the center of the rigid end platen
near the ram (see Figure 6.48c). A rigid bar, extending between the two end platens,
was attached to a bearing, supported at the center of the opposite rigid end platen
(see Figure 6.48a). Since the center of the end platens coincided with the centroid of
the specimen, the transducer measured the centroidal specimen shortening, between
the faces of the end platens. Another transducer measured lateral deflection at the
column midlength (see Figure 6.48a). Additional transducers were attached at each
end bearing to measure horizontal and vertical end rotations (see Figure 6.48b).
The plate deflections at mid-length were measured by three transducers supported
on a special frame which was attached to the specimen (see Figure 6.48e). Thus
350
Columns, Beams and Frameworks

Figure 6.48 University of Sydney interaction buckling tests on long I and channel section columns: schematic plan view of horizontal test config-
uration. The test rig is based on a 2000 kN capacity servocontrolled hydraulic loading ram located in a reaction frame which can
accommodate specimens up to 10 m long (from [6.112]): (a) plan view of test configuration, (b) detail of an end bearing, (c) Section
A-A, (d) Section B-B, (e) special frame, attached to specimen at midlength, to support plate deflection transducers
Interactive Buckling in Columns and Beams 351

the plate deflections could be measured at the same point on the specimen during
overall buckling. The frame was attached to the specimen at the flange-web junc-
tions (where local buckling deformations are small), with four set screws (see
Figure 6.48e).
The axial shortening and lateral deflection transducers, which formed part of the
control loop for controlling the tests, were (stepless) LVDTs, whereas all the trans-
ducers to measure the plate deflections (the cross-section distortions) were LDTs
(potentiometer based displacement transducers). The tests were controlled electron-
ically, usually by adjustment of the ram position to a required axial displacement,
as in Series I, with the load following automatically, also in the post-ultimate range.
In these tests, the lengths of the specimens, between the centers of the pin-ends,
were L D 2.45 6.25 m, yielding slenderness ratios of 54.7 117.2. As mentioned,
displacements and axial strains were measured during each test with displacement
transducers and strain gages, respectively.
Two tests were performed for each column size and length. One, concentrically
loaded to find the overall bifurcation load, and one with eccentrical loading, to find
the maximum strength of an imperfect specimen, whose geometric imperfection
was augmented by a nominal load eccentricity of 0.001 L. During the tests, at
loads below the local buckling load, the strain gages and displacement transducers
were used to calculate the actual equivalent load eccentricity. This was found to
be less than 0.00025 L for the “concentrically loaded” columns and of an average
magnitude of 0.00091 for the eccentrically loaded ones.
The load versus axial shortening for a typical pair of nominally identical
specimens, one concentrically loaded and one eccentrically, is shown in
Figure 6.49. After local buckling (the mean experimental local buckling load is
indicated as P exp ), the concentrically loaded specimen developed local buckling
cells of approximately uniform wavelength and amplitude along the full length of
the column. The local buckling resulted in decreasing axial stiffness with increasing
load, appearing as a change in the slope of the path shown in the figure. The
buckling mode and change in axial stiffness were similar to those observed for the
short columns of Series I. The nonlinear post local buckling path in Figure 6.49
indicates the occurrence of interaction buckling. The initial loading eccentricity
in the second specimen caused lateral deflections from the beginning of loading,
a reduced local buckling load and a lower ultimate load. However, as is usually
the case for large imperfections, the initial rate of post-ultimate load shedding was
milder than that for the concentrically loaded twin.
The experimental studies on distortional buckling by Professor Hancock and
his co-workers at the Centre of Advanced Structural Engineering of the Univer-
sity of Sydney have continued vigorously (see for example [6.176] and [6.177]).
Recently three test series on thin-walled I-sections and square hollow sections,
subject to combined compression and bending have been initiated. The group has
also embarked on another test program on plain and lipped thin-walled channel
section columns, focusing on the effect of end conditions on the interaction buck-
ling behavior. Hence one can expect a flow of a wide range of data and improved
test techniques on distortional buckling from this research center.
352 Columns, Beams and Frameworks

Figure 6.49 University of Sydney interaction buckling tests on long I-section columns: typical
measured load versus axial shortening for a pair of specimens (240 4200 A
and B, from [6.36])

Another example of interactive buckling experiments, which dealt with the


interaction between local buckling and lateral-torsional buckling of beams in
bending, were the tests carried out recently (in parallel with theoretical studies)
by Menken and his co-workers at Eindhoven University of Technology in The
Netherlands, [6.113] [6.115]. In these experiments a simply supported prismatic
aluminum T-beam was loaded in pure bending, with the flanges in compression
(see Figure 6.50a). The beam was built up from a thin flange, carefully machined
from sheet metal and glued to a relatively stiff web (Figure 6.50b). In this manner
the flanges were assumed to have a practically uniform thickness. It was then
experimentally verified that the glue had no effect on the bending stiffness.
The test rig is shown in Figure 6.51, schematically in (a) and by a general side
view in (b). The beam was simply supported by being suspended from two thin
strips, whose in-plane rigidity prevented both vertical displacement and rotation
about the longitudinal axis at the ends of the beam. On the other hand, the out-
of-plane bending and torsional flexibility of these strips permitted the ends of the
beam to rotate freely with respect to their principal axes, as required for simple
supports.
In mounting a test specimen in the test rig, twisting of the beam could occur,
which had to be prevented. A cylindrical boss made of Araldite was therefore
fixed to each end of the beam, as can be seen in Figure 6.51. Each boss was
Interactive Buckling in Columns and Beams 353

Figure 6.50 Eindhoven University of Technology experiments on interactive buckling of


beams in bending (from [6.113]): (a) test configuration, (b) cross-section of beam

inserted into a holder attached to one of the suspension strips, and the jaws of
each holder were tightened by turning a tapered nut. A lever was then attached to
each nut to apply the bending moment. A simple dead-loading device was used in
the earlier test series to apply the bending moments, which meant, however, that
descending equilibrium paths could not be followed during the tests. In later test
series [6.115], the dead-loading was therefore replaced by a device prescribing the
vertical displacement at midspan. An air bearing permitted nearly frictionless lateral
movement, while keeping the direction of loading vertical. Hold-ups (which can
be seen in Figures. 6.53a and b) prevented the lateral deflections of the specimen
from becoming excessive.
The overall buckling components (v the lateral displacement of the center of area
of the cross-section, and  the rotation of the relatively stiff web) were determined
by measuring the lateral deflections of two points on the web at the midspan of
the beam with displacement transducers. Their average was the lateral deflection v,
whereas their difference indicated the rotation  of the cross-section. The compo-
nents of local buckling were measured by four or five lightweight displacement
transducers attached to the web (see Figure 6.52b), which yielded the amplitude
of the local buckles a, and their half wavelengths , as well as the phase shift and
the average transverse displacement of the edge of the flange.
In Figure 6.53a, one can see an example of isolated local buckling, which could
only be obtained by restraining the lateral bending of the beam. When the beam was
free to deflect laterally, interactive buckling would occur as shown in Figure 6.53b,
representing a good example of combined local flange buckling and overall buckling.
354 Columns, Beams and Frameworks

(a)

Figure 6.51 Eindhoven University of Technology experiments on interactive buckling of


beams in bending test setup: (a) schematic (later test series configuration),
(b) side view (from [6.113])

In the later test series, the non-periodic local buckling was also measured by
means of a video tracking system (see Figure 6.52), which recorded the position
of an array of retro-reflective markers glued to the rim of the flange, which after
the appropriate processing yielded the true local buckling components.
These experimental studies provided needed insight into the phenomenon of
interaction between local buckling and lateral torsional buckling as a result of
Interactive Buckling in Columns and Beams 355

Figure 6.52 Eindhoven University of Technology experiments on interactive buckling of


beams in bending video tracking system for measuring the non-periodic local
buckles: (a) camera arrangement (from [6.115]), (b) the displacement transducers
and the retro-reflective markers used to determine the shape of the non-periodic
local buckles (courtesy of Dr. C.M. Menken)
356 Columns, Beams and Frameworks

Figure 6.53 Eindhoven University of Technology experiments on interactive buckling of


beams in bending (from [6.113]): (a) local buckling only, (b) combined local
and overall buckling

bending, and provided support and verification to concurrent analytical and numer-
ical studies. Theory and experiments show that interactive buckling will occur when
the local buckling load is smaller than the overall buckling load, and that the ratio
of overall to local buckling load is the primary parameter influencing the interactive
behavior. The manner of supporting the beam, the measurement of local buckling
components and other experimental techniques employed are worth noting.

6.6 Beam-Columns

6.6.1 Beam-Columns as Structural Elements

As mentioned in Chapter 2, beam-columns are structural members subjected to


combined axial compression and bending. The bending stresses may be due to
transverse loads, applied moments or eccentric loading. Most members in typical
structures, in particular in civil engineering practice, can be classed as beam-
columns, and hence beam-columns have been subjected to extensive theoretical and
experimental studies. Beam-columns occupy an important place in most structural
engineering and stability textbooks (see for example [2.1], [2.3], [6.3], [6.47] or
[6.116]), and some texts have been entirely devoted to them, (e.g. [2.22]).
Beam-Columns 357

The experimental setups and techniques employed for tests on beam-columns do


not differ significantly from those used on columns and in lateral buckling tests on
beams, except that they are sometimes a little more complicated (see for example
[6.117] or [6.118]). Hence the discussion here will focus on a recent example of
a large scale experimental study on fabricated tubular beam-columns, as used in
offshore structures.

6.6.2 Recent Experiments on Tubular Beam-Columns

The example is part of an experimental investigation on large scale unstiffened


fabricated tubular steel members carried out by Birkemoe and Prion at the Depart-
ment of Civil Engineering of the University of Toronto in the eighties [6.119]. The
goal of the experiments was to study the beam-column behavior of full scale fabri-
cated steel tubes with various ratios of axial load to bending moment, and to relate
the test results to the design and fabrication of typical members in offshore struc-
tures. In order to monitor and keep control of induced fabrication imperfections,
the specimens were fabricated in the Structures Laboratories of the University of
Toronto (except the cold rolling of the plate material into cylindrical tubes), but the
fabrication process itself followed the convention of offshore field practice. The
fabrication induced geometric distortions and residual strains, caused by welding,
were rigorously recorded during the fabrication process.
The size of the test specimens was chosen to be as near as possible to full scale,
within the capacity of the testing equipment of the laboratory, resulting in an inside
diameter of 430 mm, with wall thickness of 4.5 8.8 mm, or R/t ¾ D 25 49. The
internal diameter was kept constant to facilitate welding to reusable extension
tubes. The scale of the specimens was approximately that of full scale diagonal
bracing members and of the order of half to one fifth scale of the main jacket legs.
Three different lengths were tested 1.5 m, 5 m and 10 m, yielding approximate
slenderness ratios L/ ¾ D 10, 30 and 60, in four types of tests: concentric stub-
column tests, eccentrically loaded short column tests, 10 m long beam-column tests
(called Type C) and 5 m long beam-column tests (called Type D). Only the 10 m
long beam-column test (Type C) will be discussed here, since it was the most
challenging experimental setup. The interested reader should, however, also study
the other three types of tests described in detail in [6.119].
The specimens were fabricated by cold rolling and welding hot-rolled steel plate.
The plates (from a special batch of CSA G40.21-M350W steel) were flame-cut
and then cold formed by rolling to form cylindrical tubes (called “cans”) of length
750 mm and inside diameter of 430 mm. The roll formed tubes were then welded
longitudinally by an automated arc welding process in a single pass procedure.
Weld induced surface residual strains were measured on some of the “cans”,
with the aid of a mechanical extensometer (Pfender gage) mounted ball targets and
reference length gages. The targets consisted of 1 mm diameter hardened steel balls
set into the parent plate material with a special punch, 100 mm apart, around the
circumference of the tube, on the inside and outside, at midlength. The distances
between the targets were measured with the mechanical extensometer, before and
358 Columns, Beams and Frameworks

Figure 6.54 University of Toronto tubular steel beam-column experiments: 10 m


beam column test setup (from [6.119])

after welding to an accuracy of 1 ð 105 strain units. On a few “cans” these strain
readings were verified with resistance strain gage data.
Complete surface profiles of some of the single “cans” were recorded, from all
test specimen types, before and after the seam weld process, to determine the distor-
tion and to correlate it with the welding parameters. Then the “cans” were manually
welded together by circumferential welds, using a Metal Inert Gas welding process.
For the long beam-column tests, the 3 m long test sections consisted of four cans
each, and two 3.5 m long stiffer reusable extension tubes of equal internal diameter
and a wall thickness of 10 mm were welded to their two ends (see Figure 6.55).
To obtain an accurate geometric record of each specimen, as well of some single
cans for assessment of welding distortions, a special profiling rig was built (see
Figure 6.56) in which the specimen, or single can, was mounted vertically and
rotated about an axis close to its longitudinal centroidal axis. After approximate
centering of the specimen with adjustable screws on the top and bottom spider
clamps, the variation in its radius was measured with an LVDT mounted on a
carriage that could slide along a vertical aluminum rail. The angular position of
the rotating specimen was indicated by a rotational potentiometer attached to the
bottom shaft and the vertical position of the LVDT was obtained from a second
rotational potentiometer activated by a chain linkage. At each vertical position, a set
of 72 radius-angle readings for a complete circumferential scan were automatically
Beam-Columns

Figure 6.55 University of Toronto tubular steel beam-column experiments: 10 m beam column test layout (from [6.119])
359
360 Columns, Beams and Frameworks

Figure 6.56 University of Toronto tubular steel beam-column experiments: profiling rig for
specimens, test sections and “cans” (from [6.119])

measured and recorded electronically on a tape data file. On-line analogue plots
were produced independently, as well as analogue plots of vertical profiles which
served as back-up to the circumferential data.
Such surface profiles were obtained for all test specimens prior to testing, but for
the long beam-column specimens only their four “can” test sections were measured.
For the short specimens, the straightness of their longitudinal axis could also be
obtained from the profiling data. For the long beam-column specimens, however,
Beam-Columns 361

the straightness was determined from simple measurements of the distance between
the specimen and a taut string at four quarter circumference positions.
The complete surface profiles of the specimen test sections provided four types
of information (as discussed in detail in [6.119]):
1. a general visual impression of the specimen geometry and its initial geometric
imperfections
2. out-of-roundness figures at each circumferential profile plot
3. mismatch of abutting edges at circumferential welds
4. out-of-straightness of the test section.

Further data reduction along the lines outlined in Chapter 10, Volume 2, could
provide additional assessments on the influence of the imperfections on the buck-
ling behavior of the beam-columns.
The 10 m beam-column (Type C) horizontal test arrangement is shown in
Figure 6.54 and its layout in Figure 6.55. As seen in the layout, the 3 m long
test section of each specimen consisted of four cans, with their longitudinal welds
staggered at 90 degrees. The reusable extension tubes welded to the test section
had flanges at their outer ends, which were bolted to spherical bearings, that will
later be discussed in detail. In the tests, the axial load was first applied to a pre-
determined level and then the bending moment was introduced by application of
two equal vertical upward point loads at a distance 2.4 m apart.
The axial load was applied with a 10 MN (2000 kip) hydraulic actuator, at the
east end of the test setup, which was controlled manually through a hydraulic pres-
sure maintainer (see also Figure 6.58). The actuator was anchored to the test floor
through a reaction frame system, which was held down against overturning forces
by heavy beams bolted to the test floor. The horizontal shear was transmitted by
friction between the floor and a 3 m ð 3 m steel plate that covered 24 preten-
sioned floor bolts. Half of the anticipated horizontal force could be resisted by the
frictional force, and the remaining resistance was produced by pretensioning the
reaction system with four high strength rolled thread tension bars (DYWIDAG)
to approximately balance the frictional resistance. The horizontal resistance of the
reaction system was thus doubled and brought up to the required magnitude.
The vertical forces, which introduced the bending moment in the test section,
were applied with two pairs of hydraulic actuators (see Figure 6.55). To maintain
a constant bending moment in the test section, these vertical forces had to be kept
equal throughout the test, and were therefore controlled, using real-time control
of the servo-actuators through a micro-computer, and constantly monitored during
the test. Each vertical actuator was attached to a Lehigh gravity load simulator,
which permitted horizontal movement (shortening) of the specimen with only a
very small change in verticality of the applied loads.
This gravity simulator, developed at the Fritz Engineering Laboratory, Lehigh
University [6.121], is a mechanism that ensures vertical alignment of the load in
structures that are allowed to sway laterally under load. The basic layout of the
mechanism, as used in the Toronto beam-column tests, is shown in Figure 6.57. By
choosing appropriate ratios for the member lengths, the locus of the load point A
362 Columns, Beams and Frameworks

Figure 6.57 The mechanism of the Lehigh gravity load simulator, which ensures vertical
alignment of the load in structures that are permitted to sway laterally (from
[6.119])

(shown highly exaggerated by the full line in the figure) can be kept very close to
a horizontal line. For static equilibrium the line of action must pass the intersection
O of the long members, but for stability of the load its line of action must also be
perpendicular to the locus of A. Due to this perpendicularity the line of action is
not always vertical, but for the amount of sway allowed its deviation from vertical
was shown to be negligible. Further details on the gravity load simulators (which
have been used at Lehigh University, at the University of Wisconsin in Milwaukee
and at Nagoya University in Japan) can be found in [6.121] or in Appendix D of
[6.119]. Three such gravity load simulators were used in the 10 m beam-column
tests, two for the vertical loads and one for the vertical reaction (see Figures 6.55
and 6.58, which also show the actual construction of these simulators).
The vertical reactions at either end of the 10 m specimen were provided by reac-
tion collars (see for example Figure 6.58) which were lined with neoprene rubber
to give a more uniform load distribution and to offer the least amount of restraint
to distortion of the specimen at their locations. Similar collars were employed for
application of the vertical loads at the test section. The reaction collars, which were
instrumented and calibrated as load cells, were connected, via single high strength
steel threaded connecting rods, to hand operated hydraulic jacks that were adjustable
at each of the two reaction points. At the east end, the connecting rod and actuator
were attached via a gravity load simulator (see Figure 6.58) to allow horizontal
displacement as the specimen shortened. At the west end (not shown here in detail),
the rod and jack were directly connected to the test floor, permitting only the little
displacement arising from deflection of the reaction frame.
The end plates of the extension tubes were bolted, via adaptor plates, to spherical
bearings (see Figure 6.58), which were designed to have low friction, as opposed
Beam-Columns

Figure 6.58 University of Toronto tubular steel beam-column experiments: load application and instrumentation, east end of test rig (from [6.119])
363
364 Columns, Beams and Frameworks

to the usual spherical bridge bearings, mentioned in Sub-section 6.1.2 (and [6.11]).
The contact surfaces of these special bearings were covered with a layer of lightly
greased low friction Teflon, bearing on a polished stainless steel surface with a
friction coefficient of about 0.005.
To ensure the proper smooth functioning of these special spherical bearings (used
for the first time in large scale laboratory applications to model near moment-free
boundary conditions) in the experimental setup, a performance friction test was
carried out (see Appendix E of [6.118]). As seen in Figure 6.59 there, the two

Figure 6.59 University of Toronto tubular steel beam-column experiments: friction test rig for
spherical bearings (from [6.119])
Beam-Columns 365

spherical bearings were arranged, with a spacer block, as parts of a complete sphere.
The vertical (axial) load was applied with a 2750 kN MTS servo-controlled testing
machine. Rotation of the spherical core was achieved by application of a vertical
force, via a small spherical bearing, to a lever arm welded to the spacer block,
with a hand operated hydraulic jack in series with an electronic load cell. The jack
load F corresponded to the initial static friction load P. Since P was known, the
friction coefficient  could be found. Eleven tests yielded an average static friction
coefficient  D 0.0049 with a standard deviation of 0.00063. Load-displacement
plots of these tests showed a marked increase in friction load as the rotation of
the bearing increased, probably due to a small misalignment or to a deviation in
the radius of the spherical surface. In the Toronto test setup these increases would
not occur.
Returning to the test layout, one may note that all load application and reac-
tion points were horizontally free floating (east-west), except the western spherical
bearing which was connected directly to the reaction frame, and represented there-
fore the reference point for the displacement of the specimen.
As can be seen in Figures 6.58 and 6.60, many linear variable differential trans-
formers (LVDTs) were used in the instrumentation of the tests. The shortening of
the specimen was measured by two LVDTs at the ends of the specimen, the strain
distribution in the extension tubes was determined by a set of four mounted at
top, bottom and sides of the tube, and similarly the longitudinal strain of the test
section was measured by four LVDTs, though with a much longer gage length.
Rotation of reaction points (with reference to the longitudinal axis), was, however,
measured with rotational potentiometers attached at the load collar axes.
At three circumferential weld locations in the test section, the change in diameter
of the tube was measured vertically and horizontally. Steel rods were connected to
diametrically opposite holes in the tube and attached to the tube wall on one side
and to an LVDT (mounted to the tube via a bracket) on the other (see Figure 6.60).
Strain gages were located on the inside and outside of the test section wall
in the compression region where local buckling was expected. Additional strain
gages were also placed on the extension tubes in some of the tests to determine
the moment gradient outside the test section.
The aluminum profiling rail, that had been used in the profiling rig of
Figure 6.56, was mounted horizontally below the test section (see Figure 6.60) to
monitor the curvature changes and the formation of buckles. Profiles were recorded
as analogue plots, as the carriage was moved manually along the rail, at regular
intervals until the buckling deformations became too large.
Most of the data acquisition was done automatically. Full sets of readings were
processed at predetermined increments, but instantaneous data readings could also
be initiated by the operator. The longitudinal profile of the compression side of the
test section was also intended to be recorded automatically, but due to insufficient
speed of the controlling micro-computer this option was not used and replaced by
analog plots.
A schematic layout of the computerized test control employed is presented in
Figure 6.61. Such a computerized system for test control and data acquisition was
366 Columns, Beams and Frameworks

Figure 6.60 University of Toronto tubular steel beam-column experiments: detail of instru-
mentation in central test section (from [6.119])

used for the first time in the Structures Laboratory of the University of Toronto. The
control hardware was installed and calibrated simultaneously with the development
of software for test control and data acquisition. Appropriate interrupt routines
were introduced to permit a reasonable degree of manual control when required.
The system operated satisfactorily, but the processing speed and memory capacity
of the controlling micro-computer were found to be inadequate. It was therefore
concluded that a more powerful computer was essential for future tests. It was also
found that the manual adjustment of the vertical positions of the reaction collars
was unsatisfactory, and should be replaced by continuous adjustment with software
controlled servo-actuators.
Buckling of Frameworks 367

Figure 6.61 University of Toronto tubular steel beam-column experiments: schematic layout
of computerized test control (from [6.119])

The experimental data obtained clarified the beam-column behavior of unstiff-


ened fabricated tubular members throughout the loading, ultimate and post ultimate
stages, and pointed to inadequacies in the design rules. The experimental setup and
techniques have been discussed here in considerable detail, because they represent
modern test methods for large scale column or beam-column experiments, and the
sophistication necessary to provide the information on the behavior of the structure,
required to study the effectiveness of advanced analytical techniques.

6.7 Buckling of Frameworks


6.7.1 Frame Instability
Frames, plane and three-dimensions (spatial), are one of the commonest forms
of structures in civil engineering and other fields. They and their instability have
therefore been extensively investigated in the last decades, though most of the
studies were theoretical and numerical. Frame buckling features prominently in
many textbooks (see for example [2.3], [4.11], [4.18], [6.45] [6.47], [6.82] or
[6.122]) and several monographs and volumes have been devoted to their strength
and buckling behavior (e.g. [6.123] [6.126]).
368 Columns, Beams and Frameworks

Frames (or frameworks, as they are sometimes called) can be broadly divided
into two groups: no-sway frames and sway frames (see for example [2.3] or
[4.11]). For simplicity, the discussion is usually limited to plane frames. The first
group, the no-sway frames, which include triangulated trusses and braced building
frames, use mainly the axial stiffness of their members to maintain the shape of
the frames under load. Flexural stresses may be important, as in the beams of
building frameworks, but the essential feature of the no-sway frames is that no
substantial translational displacements of their joints can occur in the plane of the
frame without axial deformations in some of the members. Alternatively, one can
just state that the joints of no-sway frames are not free to move relative to each
other. The second group, the sway frames, which include building frames whose
beams and columns are rigidly connected at the joints but not braced, resist lateral
forces entirely by flexure of the members. In sway frames substantial translational
displacements of the joints can occur without axial deformations in any member.
Alternatively, one can just state that in sway frames the resistance to lateral loads
is provided by sway moments induced in their members.
Up to the forties, it was customary to determine the buckling load of a framework
by examining each bar individually and calculating its buckling load with some
column formula, assuming some end fixity coefficient. In reality, however, the
stability of any bar, or member of a frame, depends not only on its stiffness
but also on the amount of end restraint offered to it by adjacent bars, whose
stiffnesses in turn are influenced by the stiffnesses of their neighbors. This was
realized in the late thirties and resulted in overall frame stability analyses (see
for example [6.127] [6.129]). At the Aeronautical Laboratories of the Polytechnic
Institute of Brooklyn, Hoff and his co-workers also carried out, in the late forties,
a series of careful tests on eight rigid-jointed frameworks (two welded steel frames
and six riveted 24 S-T aluminum alloy specimens, 45 60 in. long and 15 20 in.
high), concurrently with their analyses [6.129]. Each specimen was made of two
identical vertical diagonally braced plane frames, with a length-to-height ratio of
3:1, which were connected by horizontal and diagonal braces so as to constitute
a 3-dimensional framework, in order to avoid out-of-plane buckling in the plane
frames. The stiffening effect of these horizontal braces was, however, neglected in
the analysis, which considered the two vertical frameworks to be 2-dimensional.
The specimens had each bars of different cross-sections. They were tested under
the combined action of a vertical load and a bending moment, causing tension
in the upper chords of the vertical frames and compression in their lower ones.
Care was taken that the shear and bending moment be evenly applied to the two
vertical plane frames of the specimen. These frame experiments, in which (the
then relatively new) electric strain gages were extensively employed, were a good
example of the modern aeronautical buckling experiments of the period.
From the fifties onward, frame stability has been the subject of extensive research
efforts, devoted primarily to the development of reliable methods of analysis and
design for frames. Satisfactory and safe methods have been arrived at, but as to
which of them is the optimal approach, even in 1987 “there is yet no general
agreement among the leaders of the structural engineering profession. Research,
Buckling of Frameworks 369

debating, analysis and experimentation continues . . .” (see Chapter 16 of [6.3],


p. 571).
Experiments did not feature too prominently in these investigations, perhaps
because of their likely complexity, but some were of significant importance.

6.7.2 Tests on Model Frames

In the sixties, when the postbuckling behavior of frames was widely studied, a
series of noteworthy experiments on model frames were carried out by Roorda,
Brivtec and Chilver at Cambridge University and at University College, London,
some of which were discussed already in Chapter 4, Sections 4.4 and 4.7. The
series included two groups of tests: one on pin-jointed frames [6.130] and one
on rigidly-jointed ones ([4.13] [41.5] and [4.47]). As pointed out in Chapter 4,
the models frames in both groups consisted of high-tensile spring steel members
(1 in. wide, 1/16 in. thick and 10 36 in. long), that permitted buckling and large
deformations without inducing plastic strains.
As an example, one of Brivtec’s model pin-jointed frames (Example 4 of [6.130],
see Figure 6.62) will now be considered. The reader may wish to complement
the discussion with that of Roorda’s rigidly jointed Warren truss in Chapter 4
(see Figure 4.23 there). The frame here consisted of four identical members, three
compressive ones and one tensile member.
The pin-joints were obtained, in this and the other similar model frames, by
knife edges at the ends of the struts bearing on flat notches of the end attachments
of the tensile members, which assured the required freedom of rotation at the joints
in the plane of the frame.
In order to be able to study the unstable post-buckling paths of the model frame,
an “hydraulic” device was employed. The frame was loaded by means of dead
weight (lead shot) in a cylindrical container, which was allowed to sink into a
matching cylindrical water vessel. The load applied to the frame was measured
directly by a dynamometer placed between the lead-shot container and the frame.
The diameters of the container and the water vessel were chosen so that the rate
of fall-off of the applied load exceeded the rate of fall-off of the load along the
equilibrium path of the buckled frame. Due to this “hydraulic” loading device,
the whole system, including it and the frame, represented a stable system, which
permitted following the unstable equilibrium path without a motion arising in the
frame. One may recall from Chapter 4, that for the same purpose Roorda later
used in his model Warren truss a semi-rigid loading device, consisting of a spring-
balance and screw jack combination.
The vertical displacements of the frame were measured optically (to avoid
interference with the buckling behavior) by observing the joint image of a light,
graduated scale freely suspended from a point on the frame and of a vertical cali-
brated vernier in a fixed position relative to the scale on the frame. Buckling loads
for the members of the frame were estimated with the aid of Southwell plots, as
in Chapter 4.
370 Columns, Beams and Frameworks

Figure 6.62 Brivtec’s model pin-jointed frames: elastic buckling and postbuckling of a four-
member (three compressive and one tensile) frame in a mode in which only two
members buckled (from [6.130])

In the test of the four member frame, shown in Figure 6.62, all the three compres-
sive members (1, 2 and 3) were on the point of buckling simultaneously, but only
two (1 and 2) were observed to buckle, whereas member 3 remained straight and
started to unload. The experimental unstable postbuckling equilibrium path for
this mode is shown in the figure, together with Brivtec’s theoretical prediction
for a perfect frame. Good agreement was obtained, bearing in mind the initial
geometric imperfections of the model tested. Similar results were obtained for the
other model frames tested, with one of them, a two member model, exhibiting a
stable postbuckling path.
One may note that these model frame tests, as well as the similar rigidly-
jointed ones, were essentially simple demonstration experiments, but their precision
elevated them to the status of well known and often quoted verification experiments
for nonlinear postbuckling analyses.
It may also be mentioned here that model analysis, discussed in detail in
Chapter 5, Section 5.9, has often been applied to frames, as for example in
Vaswani’s experiments on plexiglass (methyl methacryloate) models of rigid-
jointed rectangular frames [6.131].
Buckling of Frameworks 371

6.7.3 Behavior of Connections


Conventional analysis and design of frameworks used to assume that the connec-
tions of beams to columns are either fully rigid or ideally pinned. Experimental
observations, however, showed that all connections used in practice have stiff-
nesses which fall between the two extremes, and should therefore be classified as
semi-rigid connections, or flexible joints. The corresponding frames are referred
to in some specifications as PR (partially restrained). Hence, though the idealized
joint behavior, of ideally pinned or fully rigid joints, simplified the analysis and
design, it was realized that the resultant predictions of frame response to loading
would be incorrect. The real behavior of beam-column connections, of flexible
or semi-rigid joints, and their influence on frame behavior had to be investigated
in order to provide practical methods of analysis and design. The seventies and
eighties saw therefore extensive research activity on flexibly connected frames
and the relevant beam-to-column moment-resisting connections (see for example
[6.125] and [6.132] [6.138]).
The beam-to-column connection flexibility can be characterized by a moment-
rotation curve, or M-
relationship, (like those in Figure 6.63 for common types of
connections shown in Figure 6.64), which is typically nonlinear over practically the
entire loading range. Since the axial and shearing deformations are usually small
compared to the rotational deformation (and torsion can be neglected for plane
frames) the rotational deformation represents the total response of the connection
and the M-
relationship defines the connection behavior.
Gerstle [6.133] noted in 1985 “that in spite of various attempts . . . no reliable
method for prediction of connection response has been accepted by the profession”
and therefore “in the absence of analytical solutions reliance must be placed on tests
results”. However, though “connection testing has been carried out sporadically
since the 1930’s . . . complete, systematic test programmes . . . are rare”. He added
“In particular, experimental data on the behavior of modern high-strength bolted
connections are sadly lacking. . . . New connection research is needed to establish
reliable stiffness data.”
In 1989 Nethercot and Zandonini (Chapter 2 of [6.134]) concurred with Gerstle’s
assessment, stating that “at present the ability to predict the moment-rotation curve
with good accuracy is rather limited” and that “test data are not usually readily
available to designers, despite the recent attempts to assemble them in usable
collections”. In the same volume, also Davison and Nethercot point out in their
overview of connection behavior (Chapter 1 of [6.134]), that the M-
relation is
“most conveniently obtained from physical tests on connections” and that “a large
body of test data is available . . ., although not always readily accessible”.
Two years later, Chen and Lui discussed the “connection data base” in their
book [6.125] and listed some of the data bases available, in which the data was
also compared with some available prediction equations. They discussed three data
bases, covering up to ten types of beam column connections: the 1983 Goverdhan
data base [6.139], the 1985 Nethercot data base [6.140] and the 1986 Kishi and
Chen data base [6.141]. These data bases were indeed fairly comprehensive, but
not readily accessible to designers.
372 Columns, Beams and Frameworks

Figure 6.63 Flexible frame joint connections: moment-rotation curves for typical beam-to-
column connections shown in Figure 6.64 (from [6.144])

Hence, though a large amount of data has accumulated, an accessible data bank,
collecting all the suitable test data in a standard and convenient form, is still badly
needed. Also additional tests on high-strength bolted connections are yet missing.
At Purdue University, such a data bank is being assembled by Chen and Kishi, a
development of the earlier one [6.141], and for its control a special program, Steel
Connection Data Bank Program (SCDB) has been developed (see [6.136]). This
may eventually fulfill the requirements.
The test carried out to obtain the M-
relationships were essentially rotational
stiffness tests of joints subjected to bending moments. These rotational stiffness
Buckling of Frameworks 373

Figure 6.64 Common types of beam-to-column connections used in flexible frames (from
[6.144])

tests usually employed either a simple cantilever arrangement or some form of


cruciform test rig. The setup shown schematically in Figure 6.65 was used in the
beam-to-column web connection studies at the Fritz Engineering Laboratory of
Lehigh University in the late seventies [6.135] and represents a typical example of
such a cantilever test rig.
After some pilot tests on small scale web connections, four full-scale web
connection assemblages with realistic beam and column sections were tested,
consisting each of a 5.5 m (18 ft) long column and a 1.5 m (5ft) long horizontal
cantilever beam, connected at mid-height of the column. Four different geometries
of welding and bolting the beam to the column (simulating actual building connec-
tions) were used in the specimens. The assemblage was placed in a 22 240 kN
(5 000 000 lb) universal testing machine and an axial load was applied to the
column. The purpose of this axial load (that affects the yielding and deforma-
tion of the connections) was to simulate realistic loading on the web connection
assemblage. The axial load P was increased till P C Vmp (where Vmp was the
beam load V calculated to cause a fully plastic moment Mp at its critical section)
was equal to 0.5 Py , Py being the load that would cause yielding in the column.
The lower end of the column was rigidly bolted to the test floor and also the upper
end was held in a fixed end condition by the head of the testing machine. Then
374 Columns, Beams and Frameworks

Figure 6.65 Lehigh University beam-to-column web connection tests: cantilever test
rig schematic (from [6.135])

an upward load V was applied to the beam by a hydraulic jack in increments of


about 110 kN (25 kips) till deflections became excessive, when instead of specified
loads, specified deflection increments were applied.
A typical load-deflection curve (for Connection 14 3) is shown in Figure 6.66.
The loading of the beam was continued till V reached about 90 percent of Vmp , the
load calculated to cause the plastic bending moment Mp at the critical section of
the beam. The plot shows an initial linear elastic slope up to about 400 kN and then
a secondary linear slope up to a load of 890 kN. Such a type of behavior of two
distinct slopes was observed also in other tests on bolted connections conducted
at Lehigh University. The second linear slope, indicating a change in rotational
stiffness, was due to many minor slips of the bolted flange plates, though no
one major slip occurred during this test or during other Lehigh beam-to-column
connection tests. As V increased further, the load-deflection curve gradually lost
stiffness due to yielding of elements within the assemblage. Towards the end of
the loading, after V reached about 1330 kN, a tear developed in the tension flange
connection resulting in a load drop to about 1100 kN. Then the connection was
unloaded. Similar, though somewhat different, behavior was observed in the other
three tests.
A recent example of a cruciform connection test arrangement is shown in
Figure 6.67, which presents the test rig used at the University of Sheffield in the
mid-eighties [6.142]. The primary objective of the Sheffield tests was a comparative
assessment of different types of beam-to-column connections in terms of connection
Buckling of Frameworks 375

Figure 6.66 Lehigh University beam-to-column web connection tests: load-deflection curve
for connection 14-3 (from [6.135])

Figure 6.67 Sheffield University beam-to-column connection tests: cruciform test setup (from
[6.142])

rotational stiffness and moment capacity. Thus all the specimens of the 17 tests
had similar beam and column sizes and were tested on the same test setup by
the same procedure. The cruciform test arrangement of Figure 6.67 was preferred
over a cantilever type, since it required a less extensive test rig and provided some
indication of the variability of nominally identical connections.
376 Columns, Beams and Frameworks

Load was applied to the centrally located column by a 500 kN screw jack.
The reaction of each beam was measured at a distance of 1000 mm from the
column face or web for the major and minor axis tests respectively. Rotations were
measured at a point on the centerline of each beam and on the column. The rotation
at each of these three positions was gauged with the aid of three LVDTs, attached
by a system of wires and pulleys to T-bars, tack welded to the specimen, as can be
seen on the right hand side of the figure. Rotation of the specimen, and the attached
T-bar, resulted in changes in the length of the three wires, which were measured
by the LVTDs. The rotation of each joint could then be computed from the new
geometry. Positioning the T-bars as near as practicable to the connection minimized
the contribution of the beam curvature to the relative connection rotation, justifying
its neglect. This measurement system yielded only the overall moment-rotation
response and did not provide any quantitative information on the contributions of
the individual components to the connection flexibility, but for the comparative
assessment aimed at this limitation was immaterial. The data were recorded and
processed by a microcomputer-based data logging system.
In the interest of the comparative assessment, all the specimens were fabricated
at the Department’s workshop and by the same technician. For better assembly
consistency, bolt tightness was controlled, first by the touch of the same technician,
and then more efficiently by applying a predetermined torque to the bolts with a
torque wrench.
The rotational stiffness of the joints, or their connection flexibility, significantly
affects the behavior of a flexibly connected frame. The joint rotations contribute to
the overall frame deformations, and in particular to the frame sway under lateral
load. The connection flexibility also affects the buckling strength of the frame,
as well as its natural periods of vibration and therefore its dynamic response to
seismic motions. The joint flexibility also affects the distribution of internal forces
and moments and thus the resulting stresses. Hence flexibly connected frames have
to be analyzed as such (see [6.133]), taking into account the connection flexibilities,
and the methods of calculation have to be verified by experiments. A few series of
tests on the strength of flexibly connected single- and two-story frames have been
performed, but without consideration of column instability or large deflections. As
a matter of fact, Gerstle [6.133] pointed out that “no testing seems to have been
carried out on flexibly-connected frames in which failure is initiated by members
or frame buckling”. Buckling and postbuckling experiments of flexibly connected
frames are therefore still missing and warranted.
Joint flexibility also affects the buckling behavior of other types of frameworks.
An interesting example is the torsional stability of a geodesic shell-like composite
framework developed for battle-damage tolerant helicopter rear fuselages. In the
investigation carried out at Imperial College, London, for Westland Helicopters
Ltd., on these (600 mm long and of 150 mm diameter) carbon fiber reinforced
cylindrical geodetic shells [6.145] it was found that the detailed behavior of the
geodetic joints significantly affected the torsional buckling load of the geodesic
shell. In the construction of this composite framework, tape prepegs were layed
up flat, since the bars were geodesics, and the orthogonal passes produced an
Buckling of Frameworks 377

interleaved construction at the joints. The stiff carbon-fiber laminates relied on


the flexible interlaminar matrix to resist a “scissoring-action” at the joints. This
flexibly resisted “scissoring” presented the joint flexibility, which was studied both
by small specimen tests and then by a finite element model incorporating the joint
flexibility measured in these tests. With this new “flexible” model the predicted
torsional buckling load exceeded the experimental one only by about 7 percent,
whereas the earlier rigid-joint model yielded values 30 percent above experimental
buckling load. The joint flexibility was therefore also here of prime importance.
Some preliminary compression tests on complete geodesic cylindrical shells,
indicated that there the joint flexibility (the scissoring) was even more dominating.

6.7.4 Seismic Loads on Multi-Story Frames

One of the most severe loading cases for building frames are seismic loads. These
horizontal loads, resulting from the acceleration of the earthquake ground motion,
present some of the overriding design criteria for building frames. Buckling and
postbuckling occur mainly in bracing members of these structures, which experi-
ence large cyclic deviations in tension and compression in the postbuckling range
during a severe earthquake (see for example [6.146]). Thus, though seismic tests
are primarily concerned with large plastic deformations, failure and energy absorp-
tion under repeated loading, buckling and postbuckling phenomena represent an
important aspect of the behavior of the members of the frame, in particular their
bracing members. Seismic tests serve therefore also as buckling experiments.
A good example are the comprehensive tests carried out in the eighties as part of
the US/Japan Cooperative Research Program in Earthquake Engineering Utilizing
Large Scale Test Facilities, under the auspices of the US National Science Foun-
dation and the Japanese Ministry of Construction and Science and Technology
Agency. The program had three phases: one focused on a reinforced concrete test
specimen, the second on a structural steel one and the third on masonry specimens.
The second phase, focusing on steel specimens, is widely concerned with buckling
phenomena and is of interest here.
The purposes of the cooperative research program in steel and reinforced concrete
buildings were to improve earthquake-resistant design in the US and Japan; to
establish relationships between full-size and reduced-size specimen test results, to
correlate static, cyclic, pseudodynamic and shaking table experimental results; to
verify analytical modeling techniques and make recommendations for inclusion in
design codes.
The base of reference for each of the associated research projects in the structural
steel phase was the full-scale six-story steel specimen tested at the Large Size
Structures Laboratory of the Building Research Institute (BRI) in Tsukuba, Japan
(see [6.147] [6.150], [6.152] [6.154]). The structure was a six-story, two-bay-
by-two-bay, steel-framed office building with a composite steel metal deck and
lightweight concrete floor system. A typical floor plan and the elevations of the
major frames are presented in Figure 6.68. The structure was 15 m (49.2 ft) square
in plan, 22.4 m (74.3 ft) high and had two bays in each direction, but was loaded in
378 Columns, Beams and Frameworks

Figure 6.68 US Japan Joint Research Program in Earthquake Engineering full-scale 6-story
test structure geometry: (a) typical floor plan, (b) frame B, (c) frames A and C,
and (d) frames 1 and 3 (from [6.152])

the north south direction as indicated. The major load frames (A, B and C) had full
moment resisting connections and the center frame (B) had K-bracing (concentric
chevron bracing) in its south bay. The outside frames (A and C) were moment-
resistant frames, and the end frames (1 and 3) had cross-bracing with simple
connections, whereas the middle frame (2) was unbraced. The cross-bracing of the
Buckling of Frameworks 379

end frames would probably be architecturally unacceptable in an office building. It


provided, however, lateral stability for the test structure in the transverse direction
and greatly increased its torsional stiffness, minimizing accidental twisting of the
floors during the tests.
The full-scale test program was divided into four phases (see [6.147]): For
phase I the test building was designed and tested with the center frame B braced
concentrically. After completion of phase I, the concentric braces were removed,
the building was repaired and eccentric braces were installed in the north bay of
the center frame B for phase II tests. At the end of phase II, the eccentric braces
were removed from the center frame B, and the moment frames A and C (with
their rigid moment-resisting connections) were tested in phase III. In phase IV,
nonstructural walls and cladding were installed on the building and an additional
series of tests was performed.
The full-scale tests, conducted in the Large Size Testing Laboratory of BRI, used
their computer on-line actuator (COLA) pseudodynamic test system. The pseudo-
dynamic method (see for example [6.151] or [6.158]) is an integrated experimental-
analytical procedure. “It is similar to standard step-by-step nonlinear dynamic
procedures in that the controlling computer software considers the response to
be discretized into a series of time steps. Within each step the governing equations
of motion are solved numerically for the incremental structural deformations. In
the pseudodynamic method, the ground motion as well as the structure’s inertial
and damping characteristics are specified numerically as in a conventional dynamic
analysis. However, rather than using a mathematical model to determine the struc-
ture’s restoring force characteristics, these are measured directly from the damaged
specimen as the test procedure progresses. Since dynamic effects are accounted for
in the equations of motion, computed displacements are quasi-statically imposed
on the test specimen.”
The pseudodynamic test idealization is shown schematically in Figure 6.69. The
principal difference between pseudodynamic testing and well-established dynamic
analysis is that the computed structural displacements d1 , d2 , . . . are actually
imposed on the test specimen, and the restoring forces r1 , r2 , . . . are measured
experimentally from the deformed specimen.

Figure 6.69 Pseudodynamic test idealization: (a) actual structure, (b) pseudodynamic test
(from [6.151])
380 Columns, Beams and Frameworks

The method permits simulation of a wide range of seismic excitations, while


allowing the use of full-scale specimens and a slow test rate that enhances test
observations.
In the BRI full-scale tests, eight servo-controlled actuators attached to a huge
reaction wall were used to load the structure. The displacements (loads) were
applied through loading beams installed at the edge of each floor. For the COLA
test method, the structure was interfaced with the computer through the actuators
and displacement measuring transducers in such a way that the response of the
building to a given earthquake was self-controlled.
Two types of data were measured and stored. The first type was that used for
the COLA testing, which included the measured actuator forces and floor displace-
ments, and the computed velocities and accelerations. This data was used for
investigation of the overall load-deflection relationships of the building tested, as
well as for assessment of the performance of the COLA testing method. The second
type of data included member strains and displacements measured with strain gages
and LVDTs, which were used for the study of moment-rotation relationships for
beams, load-deflection behavior of braces, etc. About 1000 channels of this second
type of data were recorded for each test, which also included determination of the
force-deflection curves of the braces and their buckling behavior.
Prior to testing, linear and nonlinear analyses of very detailed models of the
building were carried out with several combinations of input ground motion and
damping levels, in order to ensure the proper parameters for the COLA pseudo-
dynamic tests. These analyses also indicated the locations at which large strains
could be expected and where therefore instrumentation should be placed. Before
and after each major test, and after any repair or modification, vibration tests were
carried out to determine the damping of the structure and the frequencies of the
first few modes. Comparison of the values measured indicated the effect of the
previous test program or repair on the structural properties.
Phase I testing focused on the behavior of frame B with concentric bracing (see
[6.147], [6.149], [6.150] and [6.152]). The record of the Miyagi-Ken-Oki 1978
earthquake (with a magnitude of 7.4 on the Richter scale) was chosen for COLA
tests, with the peak acceleration scaled to an appropriate value.
Three tests were performed: an elastic one, a moderate one and a final one. The
elastic test simulated a small, frequent earthquake with a ground motion scaled to a
peak acceleration of 6.5 m/s2 (0.065 g), throughout which the structure appeared to
remain elastic. The moderate test simulated an intermediate-size earthquake, with
a peak acceleration of 25 m/s2 (0.25 g), in which limited yielding and some brace
buckling were observed. The final test simulated a major earthquake, with a peak
acceleration of 50 m/s2 (0.5 g), and indeed extensive brace buckling and yielding
were detected. Brace buckling was observed visually in seven of the 12 braces
of frame B, and beams and columns showed yielding throughout the bottom three
stories.
The behavior of the braces is of primary importance in the overall response of
the structure. In general, bracing members provide a large portion of the lateral
stiffness and strength in steel structures comprising a bracing system and moment
Buckling of Frameworks 381

resisting frames (so called dual system), and their behavior, and in particular their
buckling and postbuckling behavior, therefore governs that of the structure, in both
the elastic and inelastic ranges.
Brace buckling was also a major source of energy dissipation in the final test
(as a matter of fact, postbuckling and yielding of the braces were the dominant
source of energy dissipation as the deflections grew larger), and it ultimately led to
failure of the north brace in the third story. One of the buckled braces is shown in
Figure 6.70. After buckling, a progressive tearing occurred, which ultimately lead
to rupture of the tubular brace. As the deflection grew even larger, the flexural
yielding of beams and columns became more important for energy dissipation.
These sequential modes of energy dissipation illustrated the beneficial redundancy
of the dual bracing system. Some of the braces buckled in-plane while others
buckled out-of-plane, which motivated detailed studies of the force deflection
behavior of individual braces (see [6.152]), that explained the experimental results
and also indicated that actually only six braces buckled, whereas the preliminary
visual observations suggested that seven did.
The effect of composite action, the influence of the composite steel and concrete
floors, on the stiffness of the braces and the whole structure, was also studied and
found to be not very significant.
Prior to the phase II testing, the cracked concrete floor slabs were repaired by
pressurized epoxy injection or recast. With the installation of the new eccentric
bracing in the north bay of frame B, modifications were also made to the instru-
mentation, like relocation of some of the potentiometers and LVDTs to the north

Figure 6.70 US Japan Joint Research Program in Earthquake Engineering: A buckled brace
(a square tube) of the full scale structure tested at BRI in Tsukuba (courtesy of
Professor C.W. Roeder)
382 Columns, Beams and Frameworks

bay. The 1952 Taft acceleration record was chosen as the input excitation for
the phase II testing (see [6.148] and [6.153]), and the same damping values as in
phase I were used. Two tests were first carried out in phase II, one for examination
of the elastic behavior with the Taft accelogram scaled to 6.5 m/sec2 (0.065 g), and
one inelastic test with the Taft record scaled to a peak acceleration of 50 m/sec2
(0.5 g).
Since relatively little damage and small story drifts and displacements occurred
in the inelastic test, three additional tests with sinusoidal ground acceleration were
performed to determine the strength, ductility and final failure mechanism of the
eccentrically braced structure. All five tests were conducted with the computer on-
line actuator (COLA) pseudodynamic test technique, the input being scaled Taft
acceleration records for the first two and a sinusoidal ground acceleration for the
last three tests.
During the third cycle of the sinusoidal tests, the gusset plates, attaching the
braces in the first story to the second level floor, buckled. This caused the end
of the brace to move out-of-plane, which produced a large torque on the second
level floor girder, resulting in a large inelastic torsional deformation. The brace
was then unable to develop and maintain large axial forces, and therefore the
stiffness and strength of the first story decreased. A similar effect, though of smaller
magnitude was also observed in the second story. Eventually the moment frames,
which as pointed out before served as a redundant backup, took over large
loads, maintaining the overall performance of the structure. However, it should be
noted that the failure of the gusset plates significantly reduced the overall stiffness
of the building, as observed in the post-test vibration measurements that showed
a 19 percent increase in the natural period (indicating a 30 percent reduction in
stiffness). This emphasized the importance of connection detail design.
The tests demonstrated the significant contribution of the shear links of the eccen-
trically braced frame (EBF) to the energy dissipation and the overall performance
of the building. They confirmed the conclusions of earlier studies (for example
[6.160]) about the efficiency of eccentrically braced frames in resisting lateral
seismic loads.
Upon completion of phase II and removal of the eccentric braces, the phase III
test on the moment frames was performed. After a vibration test, which indicated
that the elastic stiffness of the phase III moment frame was 20 25 percent of that
measured in the braced frames of phases I and II, the building was subjected to the
1940 El Centro earthquake with 35 m/s2 (0.35 g) peak acceleration. Though the
phase III structure was significantly more flexible and had less resistance than the
phase I and II configurations, its inelastic behavior was stable. After having been
subjected to several major simulated earthquakes, the moment frames performed
remarkably well.
The phase IV tests on nonstructural elements, though important to the seismic
performance of the building are not relevant to buckling experiments.
Concurrently with the extensive test program carried out on the full-scale struc-
ture at BRI in Tsukuba, also some reduced-scale models of the same structure were
tested at BRI and at some structural research centers in the USA: in Tsukuba, static
Buckling of Frameworks 383

loading tests on six half-scale model frames, representing the lower three stories
of the center frame B, were performed [6.157]. At the Fritz Engineering Labora-
tory of Lehigh University, Bethlehem, PA, a 0.305 scale model of the complete
structure was tested in a quasi-static manner [6.155]. Two series of tests on a
0.305 scale model of the complete structure, first one with concentric chevron
bracing and then with eccentric bracing, were carried out on the earthquake simu-
lator at the University of California at Berkeley [6.154] and [6.159]. And, as part
of the same program, also a series of small-scale models (1:12.5 models of three
beam-columns, one braced frame and one unbraced one) were tested at Stanford
University [6.156].
The BRI half-scale model static experiments [6.157], whose main purpose was
clarification of the elastic and plastic behavior for better prediction of that of the
full-scale structure, included six frames with inverted-V-braces, two of which were
exact half-scale models of the lower three stories of the full-scale six-story test
building, except for the details of the brace connections. The other four models
were only roughly similar and had no composite floor slabs, and one of them had
no braces. From the buckling point of view, the primary predicted and observed
phenomenon was the fact that large story drifts introduced mainly large axial
compressive displacement of the braces, causing severe local buckling that often
resulted in brace failure.
The Lehigh University 0.305 scale model was tested in three configurations for
resisting lateral forces (see [6.155]), which corresponded to those of phases I to
III of the full-scale test structure:
1. a dual system with concentrical braces in frame B (CBF) phase I;
2. a dual system with eccentrical braces in frame B (EBF) phase II;
3. a moment-resisting frame system (MRF) with frame B unbraced phase III.

The structural members of the model were exact replicas of those of the BRI
prototype. The similitude laws, derived in the manner discussed in Chapter 5,
really required for dynamic similarity a higher density material, without change
in strength and stiffness. This could have been achieved by placing additional
distributed weights on the model. However, since the model was tested statically,
no such extra weights were used.
The desired lateral loading pattern was achieved in the Lehigh tests with a two-
jack loading system and a specially constructed “wiffle tree” (similar to those used
routinely in aerospace full-scale testing). The model is shown in Figure 6.71. The
model structure was instrumented to obtain the deflection response and determine
the behavior of critical structural members. The elongation of selected braces was
measured by LVDTs, and the bracing forces were measured by pairs of strain gages
mounted on them.
After some preliminary tests including a flexibility test, the testing of each phase
consisted of some 15 25 cycles of static loading, by application of controlled
displacements, in hundreds of load steps at each of which the test data were recorded.
In the phase I test, at a certain load step one brace of the first story started
to buckle. In the following load cycle this brace buckled severely and the other
384 Columns, Beams and Frameworks

Figure 6.71 US Japan Joint Research Program in Earthquake Engineering: 0.305 scale model
of the BRI prototype tested at Lehigh University (courtesy of Professor L.-W. Lu)

brace, which was under tension, fractured at its upper connection. The fracture was
repaired and both braces were reinforced by welding narrow steel strips to the four
sides, resulting in a 56 percent increase in cross-sectional area, and then testing
was resumed. As cyclic loading continued at specified cyclic roof displacements,
additional braces buckled in- and out-of-plane and later some of them cracked. At
a roof deflection (the reference deflection here) of 66 mm (2.6 in.) the phase I test
was ended. This gradual cyclic testing, with many load steps, used also in the other
two phases, permitted a careful study of the behavior of the structure and of its
members. For example, it was found in phase I (the CBF) that buckling of braces
caused a marked reduction in the stiffness of the structure, but its load carrying
capacity did not decrease till one of the braces ruptured. Or, it was observed, from
the comparison of the EBF dual system of phase II with the similar CBF one of
phase I, that the EBF system performed much better with regard to ductility and
energy absorption capacity.
Buckling of Frameworks 385

The focus in the Berkeley dynamic tests [6.154] was on the design, construc-
tion and testing of the largest possible model of the BRI prototype that could be
accommodated on the Berkeley earthquake simulator. After extensive similarity
analyses and taking into account the weight and size limitations of the simulator, a
0.3048 scale artificial mass simulation model was chosen. The model (Figure 6.72)
satisfied geometric, stiffness and loading similitude requirements, and in order to
satisfy also the mass density similarity requirement, lead ballast was fastened to
the roof and floor slabs, in such a manner that it did not affect the stiffness of the
structure.
Since the reduced-scale structural shapes were not commercially available, all
the members of the model structure were fabricated from steel plate with the cross

Figure 6.72 US Japan Joint Research Program in Earthquake Engineering Berkeley earth-
quake simulator testing of 0.305 scale model: the concentrically braced model
structure (from [6.154])
386 Columns, Beams and Frameworks

sections designed to satisfy the similitude laws. Connection detailing was the most
difficult stage in the design of the model, and since the brace-girder joint at the
second floor level failed during the full-scale tests (see [6.147]), these joints were
modified and improved. Such and other connection modifications in the model
were, however, preceded by some subassemblage tests.
As in the BRI full-scale tests, the Berkeley model was first tested on the earth-
quake simulator with concentric braces (CBF). The model was subjected to 20
simulated ground motions, with the north south component of the 1978 Miyagi-
Ken-Oki (MO) earthquake record as input. It should be recalled that the MO
earthquake record had been used for the COLA tests of the full-scale prototype.
The ground input motion here was time-scaled according to the similitude law
and the peak ground acceleration was scaled to different levels, up to 40 m/sec2
(0.4 g), to simulate different limit states of the response of the structure.
These CBF tests showed that the bracing greatly increased the lateral stiffness,
sufficiently to avoid damage during minor earthquakes. Furthermore, the structural
strength of this concentrically chevron-braced dual system was controlled by the
overall buckling of the braces, followed by severe buckling and rupture. Hence it
was recommended to limit the design slenderness ratio for such braces.
After replacement of the concentrical braces with eccentrical braces and repair of
the model the second phase of testing on the simulator commenced. The model in
this EBF configuration (which should be more appropriately denoted here EBDS D
eccentrically braced dual system) was subjected to 24 simulated ground motions
with different earthquake records as input. The peak acceleration of these inputs
was scaled to different levels, up to 0.663 g in one collapse test, and up to 0.96 g
in one after-shock test (see [6.159]).
The EBF (or EBSD) model tests on the Berkeley simulator showed that “the
eccentrically braced dual system provided a building with all the necessary charac-
teristics for it to survive severe earthquake shaking”, provided certain appropriate
design requirements are satisfied, in particular with regard to shear links (which
play a dominant part in EBF structures).
The Stanford test program with small-scale models [6.156] focused primarily on
their feasibility and limitations in earthquake engineering. The specimens tested
included three beam-column assemblies, one braced frame and one unbraced frame,
all 1:12.5 replica models of portions of the six-story BRI prototype. For example,
the braced frame model shown in Figure 6.73 represented the center portion of the
concentrically braced prototype structure.
The model fabrication presented notable challenges, some examples of which
are presented here:
The wide flanged model beams and columns were milled from a single piece of
A36 steel plate. Since this process required removal of most of the original material
and produced very thin flanges and webs, which were poorly supported during
milling, it required great efforts in the machine shop, especially as accurate scaling
also included scaling of tolerances. As the resulting model beams and columns
were relatively free of residual stress, the models were “too good” compared to the
full-scale prototype, which could throw some doubts on the simulation of failure
Buckling of Frameworks 387

Figure 6.73 US Japan Joint Research Program in Earthquake Engineering Stanford Univer-
sity small scale model tests: braced frame test setup (from [6.156])

modes, such as buckling of medium slenderness columns, sensitive to residual


stresses.
The square tubes used for the chevron braces in the model frame were made from
round AISI-4130 tubing “with an outside diameter and wall thickness resulting in
the desired inside circumference of the square tubing. The round tube was then
heat treated to reduce the yield strength, straightened, and ground to the proper
wall thickness in a centerless grinding machine. The tubing was then packed with
wires to prevent crushing and passed repeatedly between two vee-grooved rolls
until the desired square shaped was obtained.” The model tubes were accurate
geometric replicas, but due to the cold rolling of higher strength than the full-scale
brace tubes.
The model specimen connections created some fabrication problem. Scaling of
weld sizes was difficult, and the relatively more severe local heating due to welding
388 Columns, Beams and Frameworks

resulted in larger residual stresses in the model. These dictated stress-relieving that
resulted in smaller but unknown residual stresses.
Some details of the prototype structure were too small and intricate to be repro-
duced in the small-scale model. For example, the connections at the ends of the
brace tubes in the prototype structure, with multiple stiffeners in different direc-
tions, were judged to be too complicated to reproduce in the small-scale model.
Instead, a solid steel plug was inserted into each brace end and silver soldered
to the brace tube and the beam. This avoided the difficulty of model welds, but
the simplification resulted in a very rigid connection that decreased the effective
slenderness ratio and increased the brace buckling load.
These and other examples discussed in [6.159] show that successful model
construction requires proficiency, experience and often innovative solutions as well
as judicious compromises.
In the model tests (Figure 6.73), the braced frame was placed into a reaction frame,
braced laterally, and loaded with lead weights to simulate dead load stresses. It
was instrumented to measure column strains, brace elongations, and horizontal and
vertical deflections. The distribution of lateral loads to each floor, which in the proto-
type structure varied throughout each test, was simplified in the model to a constant
load pattern, that was applied with hand-operated screw jacks and whiffle-trees.
These compromises and simplifications affected some local failure modes, but
the global behavior of the model correlated well with that of the full-scale struc-
tures. As a matter of fact, a comparison of the tests of the Stanford 1:12.5 scale
concentrically braced frame model, with that of the full-size structure tested pseu-
dodynamically at BRI and that of the 1:3.28 model tested on the Berkeley shake
table showed the same global behavior for all three types of tests.
“The major conclusions drawn from all tests with regard to global behavior is
the effect of chevron type bracing systems on inelastic response. Once one brace
buckles, the beam connected to the top of that brace is pulled downward by the
corresponding tension brace and much of the lateral loading is transferred to the
moment-resisting frame surrounding the bracing system. . . . Without the additional
resistance provided by the moment-resisting frame, the lateral resistance of the
structure would have deteriorated rapidly after the first brace buckled.”
It may be useful to quote here the general conclusions on model testing presented
by Wallace and Krawinkler in [6.156], which reiterate many of the conclusions that
appear in different sections of Chapter 5:

“1. Global static and dynamic response of steel structures can be reproduced far
into the inelastic range using carefully designed and constructed reduced-scale
models. This applies to strength and stiffness as well as the identification of
critical regions that need more detailed study through component testing.
2. Localized failure modes may be affected by fabrication issues, and size and
strain rate effects. These failure modes should be investigated using full-scale
component tests.
3. In this study the reduced-scale model tests led to the same major conclusions
on structural behavior as prototype tests.
Buckling of Frameworks 389

4. Strain rate effects are of relatively small and predictable importance in


dynamic model tests, except for some localized failure modes.
5. Size effects due to fabrication and welding may be very important in the
simulation of localized failure modes.
6. Reduced-scale model testing can yield valuable and reliable information,
provided the limitations of these testing methods are clearly understood and
are considered in model design and test interpretation.”

The US/Japan Cooperative Research Program in Earthquake Engineering


Utilizing Large Scale Test Facilities of the late eighties has been discussed at
considerable length for four main reasons: (a) it was an excellent example of
internal cooperation in an important experimental topic, involving many research
laboratories; (b) it was a major effort in detailed investigation of the influence of
scale effects in static and dynamic structural testing; (c) it emphasized the primary
effect of buckling failures of structural elements in seismic collapse of multi-story
frames; and (d) it presented a novel approach to seismic loading tests with extensive
on-line involvement of computer control the COLA pseudo-dynamic test system.
This on-line earthquake response test technique has been further developed in
Japan, where it was initiated (see [6.161] [6.163]). Tanakashi, at the University
of Tokyo, recently carried out static tests and pseudo-dynamic tests on three-story
steel frames and compared the two types of tests (see [6.163]). The frames were H-
section three-story moment-resistant frames, two fabricated from plates of a widely
used structural steel SS400 and two from plates of a newly developed high quality
steel HT50, one of each being tested statically and one pseudo-dynamically.
The test layout is shown in Figure 6.74. The frame was attached to the test
floor and braced laterally to the reaction wall against lateral-torsional deformation.
Three actuators were connected to the beam-to-column joints at the levels of the
beams. At these levels transducers were installed that measured the horizontal
displacements. The loads were applied to the frame by the actuators via load cells
installed in the actuators, which sensed their magnitudes. In addition, strain gages
were mounted on the outer surfaces of the columns, which measured the strains
used for calculation of column bending moments and hence the story shear forces.
Essentially the same test layout was employed for both static and pseudo-dynamic
tests, computer control being utilized in both cases.
The flow diagrams for the two test procedures are shown in Figure 6.75,
(a) for static and (b) for pseudo-dynamic testing. In the static tests, monotonically
increasing loads were applied at the beam levels, with the ratio between floor
loads being kept constant, 3:2:1 from top to bottom. To achieve this, as shown
in Figure 6.75a, the stroke S3 of the third actuator was assigned by the computer
(PC 9801), and the other actuators were controlled to apply the prescribed loads
L2 and L1 , which were calculated (according to the predetermined ratio) from the
load L3 , measured by the load cell of the top actuator.
In the pseudo-dynamic test (Figure 6.75b) the strokes of all the actuators S1 , S2 and
S3 , are assigned by the computer, according to its calculations. Here the equation of
motion is being solved numerically by a step-by-step procedure, where the restoring
390 Columns, Beams and Frameworks

Figure 6.74 University of Tokyo static and pseudo-dynamic tests on three-story steel frames:
test setup showing actuator locations and lateral bracing of frame (from [6.163])

force of the structure brought into the calculation at each time step is measured
instantly from the test, which is conducted in parallel with the computer calculation.
The only difference between the two test techniques is therefore the loading
history imposed in the test. In a static test the entire loading history, predetermined
by the researcher, is presented prior to the test; whereas in an on-line pseudo-
dynamic test the loading history is created in parallel to the loading.
The two test techniques are complementary. The static test is the most effective
experimental tool for determination of the strength and capacity of the struc-
ture, its maximum resistance, ductility, cumulative ductility (including buckling
and postbuckling behavior) energy dissipation capacity, etc. The on-line pseudo-
dynamic test, on the other hand, provides data on the complex hysteretic behavior
of the structure under earthquake-like loading, and validates its performance under
earthquake loading.
Before leaving the topic of seismic loading of frames, it may be worth mentioning
two recent excellent surveys of design and experimental methods in earthquake
engineering, one by Professor S. Cherry of the University of British Columbia,
Canada [6.164] and one by Professors K. Takanashi of the University of Tokyo and
M. Nakashima of the University of Kyoto [6.165]. Both reviews stress the impor-
tance and indispensability of structural testing in earthquake engineering research
for verification of existing mathematical models, for development of new models,
for study of advances in technology, for investigation of the behavior of systems too
Figure 6.75 University of Tokyo experiments on three-story steel frames: comparison of flow diagrams for (a) static test procedure, and (b) pseudo-
dynamic test procedure (on-line response test)
Buckling of Frameworks
391
392 Columns, Beams and Frameworks

complex to be completely amenable to analysis, for qualification of components


and demonstration of their operability and for surveillance of damage inflicted.
Structural testing will therefore have a primary role in earthquake engineering
research for many decades to come, and buckling and postbuckling behavior will
be one of the important facets.

6.7.5 Space Structures

With the advent of spacecraft and satellites, frameworks regained their place in
aerospace structures, and their importance has been enhanced as space stations
and large space structures become a reality. Space structures have two unique
characteristics: they operate in a gravity-free (0-g) environment (except during
their launch), and they usually have to unfold from a stowed configuration to
a deployed one, or to be assembled in space. Both characteristics also present
considerable testing problems.
Space structures are usually lightly loaded but also very flimsy, and therefore
they experience buckling and postbuckling phenomena. In the case of unfolding
space frames, buckling and postbuckling are dominant features in the folding
process and are sometimes also utilized to facilitate the folding itself.
An example of the dominance of a buckling process in the folding of a space
frame is the Olympus Astromast ([6.166] or [6.167]). The Olympus solar array
deployment system employs a continuous longeron Astromast (see Figure 6.76)
stowed in a canister with motor and guide-rails for actuating the deployment and
retraction of the solar blanket. The highest stresses occur in the canister department,
where the mast changes from a straight deployed configuration into a coiled state.
In the analysis of the transition section, between the coiled longeron and the
straight mast, difficulties arose primarily due to the large deformations combined
with geometric and material nonlinearities and the internal pre-stressing state of
the deployed mast. As a result of the pre-stressing, the battens (the members of
the triangular “horizontal” frames) were in a postbuckling state and hence had a
strongly nonlinear force-displacement relationship.
Special analytical approaches were developed to compute the transfer of the
deployed mast into the stowed state (using a large strain/displacement finite element
program LASTRAN 80 [6.168], or a special program ASTRAN [6.169]), and
demonstration models were tested to verify the computed results by measurements
(see [6.166] or [6.169]).
At the European Space Research and Technology Center at Noordwijk, The
Netherlands, a 3 m long continuous longeron Astromast demonstration model was
employed, that resembled the lower part of the Olympus Astromast but was without
a canister device. Battens and longerons of the model were made of unidirectional
GFRP, as in the Olympus hardware, but the diagonals were steel cables, instead
of the very thin GFRP rods used in the Olympus, since they did not carry any
substantial compressive loads. The end plates on both ends of the model mast
were pin-jointed to the longerons to allow them to lay down at the aluminum end
discs when coiling up (see Figure 6.76b).
Buckling of Frameworks 393

Figure 6.76 A foldable space frame the Olympus continuous-longeron Astromast (from
[6.166]): (a) the Astromast transition section, (b) end plate boundary condition

The Astromast was mounted vertically on a rotary table, with the upper end plate
held in a free rotating fixture. For deformation measurements, 15 cubic mirrors,
equipped with alignment targets, were attached to the longeron hinge points to
determine their exact position (see Figure 6.77). The rotary table had an angular
read out, indicating the angular displacement of the hinges in the transition zone.
A vertically sliding theodolite, with an alignment laser, was mounted on an optical
tooling bar at a fixed distance from the rotary table and thus the vertical position
of the hinge points was measured.
To coil up the straight deployed mast, first torsion was applied at the end plate
which caused a release of pre-tension on three diagonals per bay over the whole
mast length. Further torsion then led to a kind of “torsional buckling” of single
bays, starting in the middle and progressing towards the ends. The torsion could be
increased up to a point where all bays, except the first and last one, were twisted.
Then a sudden “snap through” of the pin-jointed longerons occurred at the end
plate. Because the longerons at the end plate changed during this “snap through”
from an almost vertical position to a coiled up horizontal state, the length of the
394 Columns, Beams and Frameworks

Figure 6.77 Demonstration model for transition section of Olympus Astromast space frame
(from [6.166]): the measuring setup

mast was reduced. When compressive loading was then applied, a movement of
mast twist towards the pin-jointed end initiated, with a concurrent untwisting of
the bays at the opposite end. The transition zone was thus formed, attaining its full
length when a full bay length of the longerons was in contact with the end plate.
The battens were in a postbuckling state when moving through the transition zone
and exhibited a strong nonlinear behavior. Since the load deformation characteristic
of the battens strongly influenced the shape of the transition zone, their buckling
and postbuckling behavior was found to be a primary factor governing the folding
process of the truss.
Returning to the first characteristic of space frameworks, their operation in a
zero-g environment, this presents the more severe testing problems. Tests floating
on air bearings or on water, or in water, have been tried to simulate the zero-
g environment on the ground for static tests. For example, Figure 6.78 shows a
bending and torsion test of an unfolding coilable longeron Astromast model, with
floatation on water simulating the zero-g environment, carried out at the Astro
Aerospace Corporation in Carpinteria, California. These simulated environments,
however, distort the dynamic properties of the structure [6.170]. Tests in vacuum
chambers or zero-g aircraft flights have also been tried, with their high costs and
other limitations. For example, in 1980 a series of deployment experiments of
a 36-element tetrahedral truss module were carried out in the NASA Langley
Research Center 55 feet vacuum facility [6.171]. Whereas small scale models of
space frames could readily be deployed by suspending the model on several soft
cords, the scale of the truss investigated (with elements 2.134 m long) precluded
soft suspension, since the gravity forces and moments were of the same order of
Buckling of Frameworks 395

Figure 6.78 Astro Aerospace Corporation coilable longeron Astromast model test with floata-
tion on water simulating the zero-g environment (courtesy of Astro Aerospace
Corporation)

Figure 6.79 Deployment test of a 36-element truss in the NASA Langley 55 feet vacuum
facility (from [6.171])
396 Columns, Beams and Frameworks

magnitude as those of deployment. The method of free fall in vacuum cylinder


was therefore chosen.
The technique of the deployment tests is shown schematically in Figure 6.79,
which also indicates the dimensions of the Langley 55 feet vacuum facility (17 m
diameter and 18 m height). The packaged truss was secured by a small diameter
cable about its girth midway along its length. A pyrotechnic cable cutter was
installed to sever the cable on command. Prior to a test, the 21 deployment springs
were cocked and the packaged truss hoisted to the top of the facility by means of
a 1.6 mm cable attached to the central cluster joint of the hexagonal surface. The
support cable passed through a pulley at the top of the facility and was attached
to a wall by means of a short loop of cable containing a pyrotechnic cable cutter.
8.2 m of slack cable was provided for free-fall as illustrated. Both pyrotechnic
devices were actuated simultaneously, allowing the truss to deploy while in free-
fall. After 1.3 seconds of free-fall, the 8.2 m of slack cable was used up and a
wire energy absorbing device installed in the support cable just above the truss
brought the deployed truss to a gentle halt. Signals from strain gage bridges and
accelerometers were recorded on tape during deployment for later analysis.
Appreciably larger trusses could have been deployed by lofting the packaged
truss upward from the floor and allowing it to deploy during both its upward and
downward trajectory, but this would have required a more complex mechanism.
Another alternative is attempting to “remove” the gravity effects analytically,
or by a combination of experiment and analysis. An example of this approach
was the determination of the zero-g shape of a space structure beam/column by
a combination of experiment and analysis carried out at Caltech (see [6.172]).
The (1/25)th model beam was an appropriately scaled down prototype, designed

Figure 6.80 Experimental setup of beam model in experiments combined with analysis to
identify the zero-g shape of a space beam (from [6.172])
References 397

as part of a large space structure. The model was 2 m long, made of steel, and
was simulating a graphite epoxy prototype. It was hung from three supports (see
Figure 6.80), its deflected shape was measured with a noncontacting displacement
transducer and the support reactions were gaged by in-line force transducers (force
rings). A measurement cart, which contained the displacement transducer, a light
diode and an optical sensor, was moved along a track below the beam by a cable
and pulley system. The track was a relatively rigid 4 m long steel I-beam. Its
flatness was measured by the optical sensor on the cart, which was controlled
by a separate computer, that also reduced the data obtained. The light diode on
the cart was placed above a black and silver strip taped to the track I-beam, and
served to trigger the analog-to-digital converter and data acquisition system. The
measurements of the force and displacement transducers and optical sensor were
combined to determine the zero-shape of the model beam.
The variation in flatness of the beam, on which the measurement cart was riding,
introduced a measurement error, that had to be taken into account. Therefore images
of the laser beam were taken by the optical sensor at 51 positions along the length
of the model beam, in order to determine accurately the shape of the support track.
Of the 13 tests performed with different load conditions (obtained by adjustment
of turnbuckles in-line with the force rings), six were with the beam face-up and
seven with it face-down. From the experimental data, the zero-g shape was for
the different loadings obtained, using the measured stiffness properties of the test
beam; or alternatively the stiffness EI was deduced by assuming the zero-g shape
to remain constant.
In large spacecraft as pointed out in [6.170] and [6.173], the differences in
shape and behavior of the structure between orbital and ground environments are
magnified. Thus their structural testing becomes more difficult and will require
more sophisticated combinations of experiment and analysis. Eventually also some
on-orbit tests will have to be carried out.
To conclude this section, one should note that future space frames, in particular
large flexible ones, and as a matter of fact large spacecraft in general, will heavily
rely on control/structures interaction technology for their deployment and operation.
This will also influence the testing techniques employed and will introduce hybrid
test systems, combinations of real hardware and computer simulation of large
components and subassemblies. For example, such systems were studied at NASA
as part of the “Control of Flexible Structures (COFS)” program in the mid eighties
(see [6.174]). Future large spacecraft and their frames will no doubt be adaptive
structures and eventually intelligent structures (see for example [6.175]).

References

6.1 Tall, L., Recent Developments in the Study of Column Behaviour, Journal of the
Institution of Engineers, Australia, 36, Dec. 1964, 319 333.
6.2 Johnson, J.B., Bryan, C.W., and Turneaure, F.E., The Theory and Practice of Modern
Framed Structures, John Wiley, New York, 1893.
398 Columns, Beams and Frameworks

6.3 Galambos, T.V., ed., Structural Stability Research Council, Guide to Stability
Design Criteria for Metal Structures, (4th edition), John Wiley and Sons, New
York Chichester, 1988.
6.4 Hayashi, Tsuyoshi, ed., Handbook of Structural Stability, Column Research
Committee of Japan, Corona Publishing Co., Tokyo, 1971.
6.5 European Convention for Constructional Steelwork, European Recommendations
for Steel Construction, ECCS (Ave. Louise 326-Bte 52, B-1050) Brusells, Belgium,
1977.
6.6 Kato, B., Cold-Formed Welded Steel Tubular Members, in Axially Compressed
Structures, Stability and Strength, R. Narayanan, ed., Elsevier Science Publishers,
London and New York, 1982, 149 180.
6.7 Tebedge, N. and Tall, L., On Testing Methods for Heavy Columns, Lehigh Univer-
sity Fritz Engineering Lab. Rep. No. 351.6, 1970.
6.8 Estuar, F.R. and Tall, L., Testing Pinned-End Steel Columns, in Test Methods for
Compression Members, ASTM STP 419, American Society for Testing Materials,
1967.
6.9 Kollbrunner, C.F., Zentrischer und exzentrischer Druck von an beiden Enden
gelenkig gelagerten Rechteckstäben aus Avional M und Baustahl, Der Stahlbau,
11, (4), Feb. 1938, 25 48.
6.10 Huber, A.W., Fixtures for Testing Pin-Ended Columns, ASTM Bulletin, 234,
December 1958.
6.11 Gent, A.R., Plastic Column Behaviour and Design at High Axial Loads, in Stability
of Structures Under Static and Dynamic Loads, SSRC International Colloquium,
Washington, D.C., May 1977, ASCE New York, 1977, 649 664.
6.12 Usami, T., and Fukumoto, Y., Local and Overall Buckling of Welded Box Columns,
Proceedings of the American Society of Civil Engineers, Journal of the Structural
Division, 108, (ST3), March 1982, 525 542.
6.13 Aschendorff, K.K., Bernard, A., Buck, O., Mang, F. and Plumier, A., Overall Buck-
ling of Heavy Rolled I-Section Columns, in Stability of Metal Structures, George
Winter Memorial Session, Third SSRC International Colloquium, Toronto, Canada,
May 1983. SSRC Fritz Engineering Laboratory Lehigh University, Bethlehem, PA,
1983, 37 49.
6.14 Ellinas, C.P., Supple, W.J., and Walker, A.C., Buckling of Offshore Structures,
Granada, London, 1984.
6.15 von Kármán, Th. and Biot, M.A., Mathematical Methods in Engineering, McGraw-
Hill Book Co., New York, 1940, 320.
6.16 Chilver, A.H., End-Fitting Effects in Strut Tests, Journal of the Royal Aeronautical
Society, 60, 1956, 275 277.
6.17 Crockett, H.B., Predicting Stiffener and Stiffened Panel Crippling Stresses, Journal
of the Aeronautical Sciences, 9, (13), 1942, 501 509.
6.18 Heimerl, G.J., Determination of Plate Compressive Strengths, NACA TN 1480,
Dec. 1947.
6.19 Gallaher, G.L., Plate Compressive Strength of FS-1h Magnesium Alloy Sheet and
a Maximum Strength Formula for Magnesium-Alloy and Aluminum-Alloy Formed
Sections, NACA TN 1714, Oct. 1948.
6.20 Schuette, E.H., Observations of the Maximum Average Stress of Flat Plates Buckled
in Edge Compression, NACA TN 1625, Feb. 1949.
6.21 Needham, A, The Ultimate Strength of Aluminum-Alloy Formed Structural Shapes
in Compression, Journal of the Aeronautical Sciences, 21, (4), 1954, 217 229.
References 399

6.22 Gerard, G., The Crippling Strength of Compression Elements, Journal of the Aero-
nautical Sciences, 25, (1), 1958, 37 52.
6.23 Fahlbusch, G. and Walkner, C., Druckversuch an Wellblechprüflingen, Test Report
73/4, Institut für Leichtbau und Flugzeugbau, Technical University, Munich,
2.10.1973.
6.24 Spier, E.E., Stability of Graphite/Epoxy Structures with Arbitrary Symmetrical
Laminates, Experimental Mechanics, 18, 1978, 401 408.
6.25 Spier, E.E. and Klouman, F.L., Ultimate Compressive Strength and Nonlinear
Stress-Strain Curves of Graphite/Epoxy Laminates, in Proc. of 5th National SAMPE
Conference, Bicentennial of Materials Progress Part II, Seattle, WA. Oct. 1976.
6.26 Batista, E., Costa Ferreira, C. and Rondal, J., Stub-Column Strength of Thin-Walled
Open Profiles, in Proceedings ECCS International Colloquium on Stability of Plate
and Shell Structures, Ghent University, April 1987, P. Dubas, and D. Vandepitte,
eds., 1987, 219 223.
6.27 Batista, E., Essais de profils C et U en acier pliés à froid, Université de Liège,
Laboratoire de Stabilité des Constructions, Rapport No. 157, 1986.
6.28 Costa Ferreira, C.M. and Rondal, J., Etude expérimentale de la stabilité des
cornières à parois minces profilées à froid, Université de Liège, Laboratoire de
Stabilité des Constructions, Rapport No. 149, 1985.
6.29 Costa Ferreira, C.M., Essais de cornières en acier pliés à froid, Université de Liège,
Laboratoire de Stabilité des Constructions, Rapport No. 155, 1986.
6.30 Costa Ferreira, C.M. and Rondal, J., Flambement des cornières à parois minces,
Annales de Travaux Publics de Belgique, No. 2, 1986, 101 121.
6.31 Nagwaney, A., Mechanical Properties of Aluminum-Lithium-Superalloy in Com-
pression, Student Research Project at the Institut für Leichtbau, Technical Univer-
sity, RWTH Aachen, (under the guidance of H. Öry and S. Zurhorst), July 1992.
6.32 Graves-Smith, T.R. and Sridharan, S., Elastic Collapse of Thin-Walled Columns,
in Thin-Walled Structures, J.R. Rhodes, and A.C. Walker, eds., Granada, London,
1979, 719 732.
6.33 Hancock, G.J., Davids, A.J., Key, P.W., Lau, S.C.W. and Rasmussen, K.J.R.,
Recent Developments in the Buckling and Nonlinear Analysis of Thin-Walled
Structural Members, Thin-Walled Structures, 9, 1990, 309 338.
6.34 Loughlan, J. and Howe, D., The Influence of Local Buckling on the Behaviour of
Some Thin-Walled Compression Members, in Proceedings, International Confer-
ence on Steel and Aluminium Structures, Cardiff, July 1987, Constrado, Coventry,
England, 1987, 450 464.
6.35 Rasmussen, K.J.R. and Hancock, G.J., Compression Tests of Welded Channel
Section Columns, Journal of Structural Engineering, ASCE, 115, (4), April 1989,
789 808.
6.36 Davids, A.J. and Hancock, G.J., Compression Tests of Long Welded I-Section
Columns, Journal of Structural Engineering, ASCE, 112, (10), Oct. 1986,
2281 2297.
6.37 Ballio, G., Finzi, L. and Urbano, C., Centrally Compressed High Strength Steel
Round and Square Tubes: Theoretical and Experimental Investigations, in Prelim-
inary Report, 2nd International Colloquium on Stability of Steel Structures, Liège,
April 1977, 77 84.
6.38 Zaras, J. and Rhodes J., Carefully Controlled Compression Tests on Thin-Walled
Cold-Formed Sections, Applied Solid Mechanics-2, A.S. Tooth and J. Spence, eds.,
Elsevier Applied Science Publishers, London, 1988, 519 551.
400 Columns, Beams and Frameworks

6.39 Wagner, H., Verdrehung und Knickung von offenen Profilen, in Festschrift
Fünfundzwanzig Jahre Technische Hochschule Danzig, Kafemann, Danzig, 1929;
329 343, translated as Torsion and Buckling of Open Sections, NACA TM 807,
1937.
6.40 Kappus, R., Drillknicken zentrisch gedrückter Stäbe mit offenem Profil im elastis-
chen Bereich, Luftfahrtforschung, 14, (9), 444 457, 1937; translated as Twisting
Failure of Centrally Loaded Open-Section Columns in the Elastic Range, NACA
TM 851, 1938.
6.41 Lundquist, E.E. and Fligg, C.M., A Theory for Primary Failure of Straight Centrally
Loaded Columns, NACA Report 582, 1937.
6.42 Pugsley, A.G., Torsional Instability in Struts, Aircraft Engineering, 4, (43),
Sept. 1932, 229 230.
6.43 Goodier, J.N., Torsional and Flexural Buckling of Bars of Thin-Walled Open
Section under Compressive and Bending Loads, Journal of Applied Mechanics,
9, (3), Sept. 1942, A103 A107.
6.44 Hoff, N.J., Strain Energy Derivation of Torsional-Flexural Buckling Loads of
Straight Columns of Thin-Walled Open Sections, Brown University, Quarterly of
Applied Mathematics, 1, (4), June 1944, 341 345.
6.45 Niles, A.S. and Newell, J.S., Airplane Structures, II, 3rd ed., John Wiley and Sons,
New York, 1958.
6.46 Bleich, F., Buckling Strength of Metal Structures, McGraw-Hill, New York, 1952.
6.47 Yu, W.-W., Cold-Formed Steel Design, John Wiley, New-York, 1985.
6.48 Wagner, H. and Pretschner, W., Verdrehung und Knickung von offenen Profilen,
Luftfahrtforschung, 11, (6), 174 180, December 5, 1934; translated as Torsion and
Buckling of Open Sections, NACA TM. 784, January 1936.
6.49 Kollbrunner, C.F., Das Ausbeulen des auf Druck beanspruchten freistehenden
Winkels, Thesis ETH, Gebr. Leemann & Co., Zürich & Leipzig, 1935; or
Mitteilungen, 4, Institut für Baustatik, ETH, Zürich, 1935.
6.50 Niles, A.S., Experimental Study of Torsional Column Failure, NACA TN 733,
Oct. 1939.
6.51 Niles, A.S., Tests of Flat Panels with Four Types of Stiffeners, NACA TN 882,
Jan. 1943.
6.52 Levy, S. and Kroll, W.D., Primary Instability of Open-Section Stringers Attached
to Sheet, Journal of the Aeronautical Sciences, 15, (10), Oct. 1948, 581 591.
6.53 Dunn, L.G., An Investigation of Sheet-Stiffener Panel Subjected to Compres-
sion Loads with Particular Reference to Torsionally Weak Stiffeners, NACA TN
752, 1940.
6.54 Ramberg, W. and Levy, S., Instability of Extrusions Under Compressive Loads,
Journal of the Aeronautical Sciences, 12, (10), Oct. 1945, 485 498.
6.55 Thomas, E.W., Torsional Instability of Thin Angle Section Struts, The Structural
Engineer, 19, (5), 1941, 73 82.
6.56 Chajes, A. and Winter, G., Torsional-Flexural Buckling of Thin-Walled Members,
Journal of the Structural Division, ASCE Proceedings, 91, (ST4), 1965, 103 124.
6.57 Hone, C.P., Torsional-Flexural Buckling of Axially-Loaded, Thin-Walled, Elastic
Struts of Open Cross-Section, in Thin-Walled Structures, A.H. Chilver, ed., John
Wiley, New York, 1967, 103 135.
6.58 Pekoz, T.B. and Winter, G., Torsional-Flexural Buckling of Thin-Walled Sections
Under Eccentric Load, Journal of the Structural Division, Proceedings, ASCE, 95,
(ST5), May 1969, 941 963.
References 401

6.59 Chajes, A., Fang, P.J. and Winter, G., Torsional-Flexural Buckling, Elastic
and Inelastic, of Cold Formed Thin-Walled Columns, Cornell Engineering
Research Bulletin 66-1, School of Engineering, Cornell University, Ithaca, N.Y.,
August 1966.
6.60 Kennedy, J.B. and Murty, K.S.M., Buckling of Steel Angle and Tee Struts,
Journal of the Structural Division, Proceedings ASCE, 98, (ST11), Nov. 1972,
2507 2521.
6.61 Kennedy, J.B. and Madugula, K.S.M., Buckling of Angles: State of the Art,
Journal of the Structural Division, Proceedings ASCE, 108, (ST9), Sept. 1982,
1967 1980.
6.62 Loomis, R.S., Loomis, R.H., Loomis, Robert W. and Loomis Richard, W., Torsional
Buckling Study of Hartford Coliseum, Journal of the Structural Division, ASCE,
106, (ST1), Jan. 1980, 211 231.
6.63 Smith, E.A. and Epstein, H.I., Hartford Coliseum Roof Collapse: Structural
Collapse Sequence and Lessons Learned, Civil Engineering, ASCE, April 1980,
59 62.
6.64 Smith, E.A., Buckling of Four Equal-Leg Angle Cruciform Columns, Journal of
Structural Engineering, ASCE, 109, (2), Feb. 1983, 439 450.
6.65 Hancock, G.J., Distortional Buckling of Steel Storage Rack Columns, Journal of
Structural Engineering, ASCE, 111, (12), Dec. 1985, 2770 2783.
6.66 Lau, S.C.W. and Hancock, G.J., Distortional Buckling Tests of Cold-Formed Chan-
nel Sections, Ninth International Speciality Conference on Cold-Formed Steel Struc-
tures, St. Louis, Missouri, Nov. 8 9, 1988, 45 73.
6.67 Burr, W.H. and Elmore, G.H., Tests of Wrought Iron Eye Beams, Selected Papers
of the Rensselaer Society of Engineers, 1, (1), January 1884, 3 17.
6.68 Michell A.G.M., Elastic Stability of Long Beams Under Transverse Forces, Philo-
sophical Magazine, 48, 1899, 298 309.
6.69 Marburg, E., Tests of Standard I-Beams and Bethlehem Special I-Beams and
Girder Beams, Transactions of the American Society for Testing Materials, 9, 1909,
378 412.
6.70 Timoshenko, S.P., Sur la stabilité des systemes elastiques, Annales des Ponts et
Chaussees, Memoires et Documents, Paris, 1913, 496 566.
6.71 Moore, H.F., Strength of I-Beams in Flexure, University of Illinois Engineering
Experiment Station, Bulletin No. 68, Sept. 1913.
6.72 Prescott, J., The Buckling of Deep Beams, Philosophical Magazine, 36, (214),
Oct. 1918, 297 314.
6.73 Prescott, J., The Buckling of Deep Beams, (with appendix by H. Carrington), Philo-
sophical Magazine, 39, 1920, 194 219.
6.74 Bromley, S. and Robinson, W.H., The Lateral Failure of Spars, NACA TN 232,
March 1926.
6.75 Procter, A.N., Laterally Unsupported Beams, The Structural Engineer, 10, July
1932, 274 287.
6.76 Dumont, C. and Hill, H.N., The Lateral Instability of Deep Rectangular Beams,
NACA TN 601, 1937.
6.77 Dumont, C. and Hill, H.N., The Lateral Stability of Equal-Flanged Aluminum Alloy
I-Beams Subjected to Pure Bending, NACA TN 770, August 1940.
6.78 Flint, A.R., The Stability and Strength of Slender Beams, Engineering (England),
170, Dec. 22, 1950, 545 549.
402 Columns, Beams and Frameworks

6.79 Lee, G.C., A Survey of Literature on the Lateral Instability of Beams, Welding
Research Council Bulletin No. 63, Aug. 1960.
6.80 Galambos, T.V., Laterally Unsupported Beams, in Introductory Report 2nd Inter-
national Colloquium on Stability of Steel Structures, ECCS-ABSE, Liège, 1977,
365 373.
6.81 Trahair, N.S., Lateral Buckling of Beams and Beam Columns, Ch. 3 in Theory of
Beam Columns, 2, W.F. Chen and T. Atsuta, eds., McGraw-Hill, New York, 1977,
71 157.
6.82 Chen, W.F. and Lin, E.M., Structural Stability, Theory and Implementation, Else-
vier New York-Amsterdam, 1987, 317 380.
6.83 Nethercot, D.A. and Trahair, N.S., Design of Laterally Unsupported Beams, in
Beams and Beam Columns, R. Narayanan, ed., Applied Science Publishers, London
& New York, 1983, 71 94.
6.84 Fukumoto, Y. and Kubo, M., An Experimental Review of Lateral Buckling of
Beams and Girders, in International Colloquium on Stability of Structures Under
Static and Dynamic Loads, ASCE, New York, 1977, 541 562.
6.85 Fukumoto, Y., Itoh, Y. and Kubo, M., Strength Variation of Laterally Unsupported
Beams, Journal of the Structural Division, Proceedings ASCE, 106, (ST1),
Jan. 1980, 165 181.
6.86 Fukumoto, Y. and Itoh, Y., Statistical Study of Experiments on Welded Beams,
Journal of the Structural Div., ASCE, 107, (ST1), Jan. 1981, 89 103.
6.87 Fukumoto, Y., Itoh, Y. and Hattori, R., Lateral Buckling Tests on Welded Contin-
uous Beams, Journal of the Structural Div., Proceedings ASCE, 108, (ST10), Oct.
1982, 2245 2262.
6.88 Poowannachaikul, T. and Trahair, N.S., Inelastic Buckling of Continuous Steel I-
Beams, Civil Engineering Transactions, Institution of Engineers, Australia, CE18,
(2), 1976, 134 139.
6.89 Bansal, J.P., The Lateral Instability of Continuous Steel Beams, Ph.D. Thesis, Uni-
versity of Texas at Austin, August 1971.
6.90 Gedies, R.W., Beam Buckling Tests with Various Brace Stiffnesses, M.Sc. Thesis,
University of Texas at Austin, December 1983.
6.91 Cheng, J.J., Yura, J.A. and Johnson, C.P., Design and Behavior of Coped Beams,
PMFSL Report No. 84-1, Dept. of Civil Engineering, University of Texas at Austin,
July 1984.
6.92 Cheng, J.J. and Yura, J.A., Lateral Buckling Tests on Coped Steel Beams, Journal
of Structural Engineering ASCE, 114, (1), Jan. 1988, 16 29.
6.93 Cheng, J.J., and Yura, J.A., Lateral Buckling of Coped Beams, Journal of Structural
Engineering ASCE, 114, (1), Jan. 1988, 1 15.
6.94 du Plessis, D.P., Lateral-Torsional Buckling of End-Notched Steel Beams, Inter-
national Colloquium on Stability of Structures under Static and Dynamic Loads,
Structural Stability Research Council, May 1977, 563 572.
6.95 Meck, H.R., Experimental Evaluation of Lateral Buckling Loads, Journal of Engi-
neering Mechanics Div., ASCE, 103, (2), 1977, 331 337.
6.96 Ramirez, D.R., The Effect of Beam Yielding on the Stability of Columns an Ex-
perimental Study, CESRL Thesis No. 75-1, University of Texas at Austin, Jan. 1975.
6.97 Massey, P.C. Elastic and Inelastic Lateral Instability of I-Beams, The Engineer,
216, (5622), Oct. 1963, 672 674.
6.98 Trahair, N.S., Deformations of Geometrically Imperfect Beams, Journal of the
Structural Div., ASCE, 95, (ST7), July 1969, 1475 1496.
References 403

6.99 Yura, J.A., Discussion and Closure (by Trahair) of Deformation of Geometrically
Imperfect Beams, Journal of the Structural Division, ASCE, 96, (ST1), Jan. 1970,
162 163 and (ST11), Nov. 1970, 2523 2524.
6.100 Bijlaard, P.P. and Fisher, G.P., Interaction of Column and Local Buckling in Com-
pression Members, NACA TN 2640, March 1952.
6.101 Bijlaard, P.P. and Fisher, G.P., Column Strength of H-Sections and Square Tubes
in Postbuckling Range of Component Plates, NACA TN 2994, August 1953.
6.102 Thomson, J.M.T., An Engineering Approach to Interactive Buckling, International
Journal of Mechanical Sciences, 16, 1974, 335 336.
6.103 Pignataro, M. and Luongo, A., Multiple Interactive Buckling of Thin-Walled
Members in Compression, in Proceedings ECCS International Colloquium on
Stability of Plate and Shell Structures, Ghent University, April 1987, P. Dubas and
D. Vandepitte, eds., 1987, 235 240.
6.104 Rothwell, A., An Experimental Investigation of the Efficiency of a Range
of Channel Section Struts, Aeronautical Journal, Royal Aeronautical Society, 78,
Sept. 1974, 426 430.
6.105 Kalyanaraman, V., Pekoz, T. and Winter, G., Unstiffened Compression Elements,
Journal of the Structural Division, Proceedings ASCE, 103, (ST9), Sept. 1977,
1833 1848.
6.106 Hancock, G.J., Interaction Buckling in I-Section Columns, Journal of the Structural
Division, Proceedings ASCE, 107, (ST1), Jan. 1981, 165 179.
6.107 Usami, T. and Fukumoto, Y., Welded Box Compression Members, Journal of
Structural Engineering ASCE, 110, (10), Oct. 1984, 2457 2469.
6.108 Davids, A.J. and Hancock, G.J., Compression Tests of Short Welded I-Sections,
Journal of Structural Engineering, ASCE, 112, (5), May 1986, 960 976.
6.109 Lindström, G., Column Strength of Welded I-Sections in Postbuckling Range of
Component Plates, Bulletin No. 138, The Dept. of Structural Mechanics and Engi-
neering, Royal Institute of Technology (KTH), Stockholm, 1982.
6.110 Thomassen, P.O., Thin-Walled C-Shaped Panels in Axial Compression, Document
D-1978, Swedish Council for Building Research, Stockholm, 1978.
6.111 Mulligan, G.P. and Pekoz, T., Locally Buckled Thin-Walled Columns, Journal of
Structural Engineering, ASCE, 110, (11), 1984, 2635 2654.
6.112 Rasmussen, K.J.R., The Behaviour of Thin-Walled Channel Section Columns,
Ph.D. Thesis, University of Sydney, School of Civil and Mining Engineering,
December 1988.
6.113 Menken, C.M., Groot, W.J. and Stallenberg, G.A.J., Interactive Buckling of Beams
in Bending, Thin-Walled Structures, 12, 1991, 415 434.
6.114 Menken, C.M. and van Erp, G.M., Buckling of Thin-Walled Beams Under
Concentrated Transverse Loading, in Proceedings, IUTAM Symposium on Contact
Loading and Local Effects in Thin-Walled Plated and Shell Structures, V. Krupka,
and M. Drdacky, eds., Academia, Prague, 1992, 165 172.
6.115 Menken, C.M., Kouhia, R. and Groot, W.J., An Investigation into Non-Linear
Interaction Between Buckling Modes, Thin-Walled Structures, 19, (2 4), 1994,
129 145.
6.116 Narayanan, R., ed., Beams and Beam Columns, Applied Science Publishers, London
and New York, 1983.
6.117 Van Kuren, R.C. and Galambos, T.V., Beam Column Experiments, Journal of the
Structural Division, Proceedings ASCE, 90, (ST2), April 1964, 223 256.
404 Columns, Beams and Frameworks

6.118 Nakashima, M., Nakamura, T. and Wakabayashi, M., Post-Buckling Instability of


Steel Beam-Columns, Journal of Structural Engineering, ASCE, 109, (6), June
1983, 1414 1429.
6.119 Prion, H.G.L. and Birkemoe, P.C., Experimental Behavior of Unstiffened Fabri-
cated Tubular Steel Beam-Columns, University of Toronto, Department of Civil
Engineering Report No. 88-03 (ISBN 07 7727 7096 4), 1988.
6.120 Prion, H.G.L. and Birkemoe, P.C., Beam-Column Behavior of Fabricated Steel
Tubular Members, Journal of Structural Engineering, ASCE, 118, (5), May 1992,
1213 1232.
6.121 Yarmici, E., Yura, J.A. and Lu, L.W., Techniques for Testing Structures Permitted
to Sway, Experimental Mechanics, 7, (8), August 1967, 321 331.
6.122 Butterworth, J.W., Frame Instability in: Structural Instability, W.J. Supple, ed., IPC
Science and Technology Press, Guildford, Surrey, 1973.
6.123 Horne, M.R. and Merchant, W., The Stability of Frames, Pergamon Press,
Oxford, 1965.
6.124 Gregory, M., Elastic Instability: Analysis of Buckling Modes and Loads of Framed
Structures, E.V.&.N. Spon, London, 1967.
6.125 Chen, W.F. and Lui, E.M., Stability Design of Steel Frames, CRC Press, Boca
Raton, Boston, London, 1991.
6.126 Narayanan, R., ed., Steel Framed Structures, Stability and Strength, Elsevier
Applied Science Publishers, London and New York, 1985.
6.127 Lundquist, E.E., Principles of Moment Distribution Applied to Stability of Struc-
tural Members, Proceedings of the Fifth International Congress of Applied Mecha-
nics, John Wiley & Sons, New York, 1938, 145 148.
6.128 Hoff, N.J., Stable and Unstable Equilibrium of Plane Frameworks, Journal of the
Aeronautical Sciences, 8, (3), 1941, 115 119.
6.129 Hoff, N.J., Boley, B.A., Nardo, S.V. and Kaufman, S., Buckling of Rigid-Jointed
Plane Trusses, Transactions, American Society of Civil Engineers, 116, Paper
(2454), 1951, 958 986.
6.130 Brivtec, S.J., Elastic Buckling of Pin-Jointed Frames, International Journal of
Mechanical Sciences, 5, (6), 1963, 447 460.
6.131 Vaswani, H.P., Model Analysis Method for Determining Buckling Load of Rect-
angular Frames, Experimental Mechanics, 1, 1961, 55 64.
6.132 Chen, W.F. and Lui, E.M., Beam-to-Column Moment-Resisting Connections, in
Steel Framed Structures, Stability and Strength, R. Narayanan, ed., Elsevier Applied
Science Publishers, London and New York, 1985, 153 203.
6.133 Gerstle, K.H., Flexibly Connected Steel Frames, in Steel Framed Structures, Stabi-
lity and Strength, R. Narayanan, ed., Elsevier Applied Science Publishers, London
and New York, 1985, 205 239.
6.134 Narayanan, R., ed., Structural Connections, Stability and Strength, Elsevier Applied
Science, London and New York, 1989.
6.135 Rentschler, G.P., Chen, W.F. and Driscoll, G.C., Tests of Beam-to-Column Web
Moment Connection, Journal of the Structural Division, Proc. ASCE, 106, (ST5),
May 1980, 1005 1022.
6.136 Chen, W.F. and Kishi, N., Semirigid Steel Beam-to-Column Connections: Database
and Modeling, Journal of Structural Engineering, ASCE, 115, (1), Jan. 1989,
105 119.
6.137 Marley, M.J., Analysis and Tests of Flexibly-Connected Steel Frames, Report to
AISI, University of Colorado, Boulder, Colorado, 1982.
References 405

6.138 Kato, B., Beam-to-Column Connection Research in Japan, Journal of the Structural
Division, ASCE, 108, (ST2), Feb. 1982, 343 360.
6.139 Goverdhan, A.V., A Collection of Experimental Moment-Rotation Curves and
Evaluation of Prediction Equations for Semi-Rigid Connections, Master’s Thesis,
Vanderbilt University, Nashville, TN, 1983.
6.140 Nethercot, D.A., Steel Beam-to-Column Connections A Review of Test Data and
its Application to the Evaluation of Joint Behaviour in the Performance of Steel
Frames, CIRIA Project Record 338, London, 1985.
6.141 Kishi, N. and Chen, W.F., Data Base of Steel Beam-to-Column Connections, Struc-
tural Engineering Report No. CE-STR-86-26, School of Civil Engineering, Purdue
University, West Lafayette, IN, 1986.
6.142 Davison, J.B., Kirby, P.A. and Nethercot, D.A., Rotational Stiffness Characteristics
of Steel Beam-to-Column Connections, Journal of Constructional Steel Research,
8, 1987, 17 54.
6.143 Azizinamini, A., Bradburn, J.H. and Radziminski, J.B., Initial Stiffness of Semi-
Rigid Steel Beam-to-Column Connections, Journal of Constructional Steel Rese-
arch, 8, 1987, 71 90.
6.144 Nethercot, D.A. and Chen, W.F., Effects of Connections on Columns, Journal of
Constructural Steel Research, 10, 1988, 210 239.
6.145 Sandhu, J.S., Stevens, K.A. and Davies, G.A.O., Torsional Buckling and Postbuck-
ling of Composite Geodetic Cylinders with Special Reference to Joint Flexibility,
Composite Structures, 15, 1990, 301 322.
6.146 Goel, S.C., Seismic Stability of Bracing Members, Proceedings of the SSRC 4th
International Colloquium on Stability of Metal Structures, North American Session,
New York, April 1989, 451 455.
6.147 Foutch, D.A., Goel, S.C. and Roeder, C.W., Seismic Testing of Full-Scale Steel
Building Part I, Journal of Structural Engineering, ASCE, 113, (11), 1987,
2111 2129.
6.148 Roeder, C.W., Foutch, D.A. and Goel, S.C., Seismic Testing of Full-Scale Steel
Building Part II, Journal of Structural Engineering, ASCE, 113, (11), 1987,
2130 2145.
6.149 Yamamouchi, H., Midorikawa, M., Nishiyama, I. and Watabe, M., Experimental
Results on a K-Braced Steel Structure Under Seismic Loading Utilizing Full-Scale
Six Story Test Structure US/Japan Cooperative Research Program, Proceedings
of the Annual Technical Session of the Structural Stability Research Council (SSRC)
1984, San Francisco, California, 1984.
6.150 Yamamouchi, H., Midorikawa, M., Nishiyama, I. and Watabe, M., Seismic
Behavior of Full-Scale Concentrically Braced Steel Building Structure, Journal
of Structural Engineering, ASCE, 115, (8), Aug. 1989, 1917 1929.
6.151 Mahin, S.A. and Shing, P.-S.B., Pseudodynamic Method for Seismic Testing, Jour-
nal of Structural Engineering, ASCE, 111, (7), July 1985, 1482 1503.
6.152 Roeder, C.W., Seismic Behavior of Concentrically Braced Frame, Journal of Struc-
tural Engineering, ASCE, 115, (8), Aug. 1989, 1837 1856.
6.153 Foutch, D.A., Seismic Behavior of Eccentrically Braced Steel Building, Journal of
Structural Engineering, ASCE, 115, (8), Aug. 1989, 1857 1876.
6.154 Bertero, V.V., Uang, C.-M., Llopiz, C.R. and Igarashi, K., Earthquake Simulator
Testing of Concentric Braced Dual System, Journal of Structural Engineering,
ASCE, 115, (8), Aug. 1989, 1877 1894.
406 Columns, Beams and Frameworks

6.155 Lee, S.-J. and Lu, L.-W., Quasi-Static Tests of Scaled Model Building, Journal of
Structural Engineering, ASCE, 115, (8), Aug. 1989, 1895 1916.
6.156 Wallace, B.J. and Krawinkler, H., Small-Scale Model Tests of Structural Steel
Assemblies, Journal of Structural Engineering, ASCE, 115, (8), Aug. 1989,
1999 2015.
6.157 Fukuta, T., Nishiyama, I., Yamamouchi, H. and Kato, B., Seismic Performance of
Steel Frames with Inverted V Braces, Journal of Structural Engineering, ASCE,
115, (8), Aug. 1989, 2016 2028.
6.158 Mahin, S.A., Shing, P.-S.B., Thewalt, C.R., and Hanson, R.D., Pseudodynamic Test
Method Current Status and Future Directions, Journal of Structural Engineering,
ASCE, 115, (8), Aug. 1989, 2113 2128.
6.159 Whittaker, A.S., Uang, C.M. and Bertero, V.V., Experimental Behavior of Dual
Steel System, Journal of Structural Engineering, ASCE, 115, (1), Jan. 1989,
183 200.
6.160 Hjelmstad, K.D. and Popov, E.P., Characteristics of Eccentrically Braced Frames,
Journal of Structural Engineering, ASCE, 110, (2), Feb. 1984, 340 353.
6.161 Takanashi, K. and Nakashima, M., Japanese Activities on On-Line Testing, Journal
of Engineering Mechanics, ASCE, 113, (7), 1987, 1014 1032.
6.162 Elnashai, A.S., El-Ghazouli, A.Y. and Dowling, P.J., Verification of Pseudo-
Dynamic Testing of Steel Members, Journal of Constructional Steel Research, 16,
1990, 153 161.
6.163 Takanashi, K., Experimental Behaviour of Multi-Story Steel Frames, Journal of
Constructional Steel Research, 29, 1994, 175 189.
6.164 Cherry, S., Structural Testing in Earthquake Engineering Research, in The Future
of Structural Testing, Computational Mechanics Publications, Southampton, Boston
and McGraw-Hill Book Company, New-York, 1990, 93 118.
6.165 Takanashi, K. and Nakashima, M., Stability Considerations on Seismic Perfor-
mance of Steel Structures, in Proceedings of SSRC 50th Anniversary Conference,
Lehigh University, Bethlehem, Pennsylvania, June 1994, SSRC Fritz Engineering
Laboratory, 1994, 119 133.
6.166 Eiden, M., Brunner, O. and Stavrinidis, C., Deployment Analysis of the Olympus
Astromast and Comparison with Test Measurements, Journal of Spacecraft and
Rockets, 24, (1), 1987, 63 68.
6.167 [Anonymous] Astromasts for Space Applications, Report AAC-B-004, Astro
Aerospace Corporation, Carpinteria, California, July 1985.
6.168 Argyris, J.H. et al., LASTRAN 80 Users Manual, ISD Report No. 279, University
of Stuttgart, Germany, Jan. 1984.
6.169 Hedgepeth, J.M., Application of High-Fidelity Structural Deployment Analysis to
the Development of Large Deployable Trusses, Preprint IAF-89-339, 40th IAF
Congress Malaga, Spain, Oct. 1989.
6.170 Hanks, B.R. and Pinson, L.D., Large Space Structures Raise Testing Challenges,
Astronautics & Aeronautics, 21, Oct. 1983, 34 40,53.
6.171 Herr, R.W. and Horner, G.C., Deployment Tests of a 36-Element Tetrahedral Truss
Module, in Large Space Systems Technology 1980, Vol. II Base Technology, NASA
Conference Publication CP 2168, Nov. 1980, 59 69.
6.172 Balas, G.J. and Babcock, C.D., Identification of the Zero-g Shape of a Space Beam,
Journal of Spacecraft and Rockets, 25, (6), 1988, 405 412.
6.173 Pinson, L.D., Recent Advances in Structural Dynamics of Large Space Structures,
Acta Astronautica, 19, (2), 1989, 161 170.
References 407

6.174 Pyle, J.S., COFS II 3-D Dynamics and Controls Technology, in NASA/DOD
Control/Structures Interaction Technology 1986, NASA Conference Publication
CP 2447, Part 1, 1986, 327 345.
6.175 Crawley, E.F., Intelligent Structures for Aerospace: A Technology Overview and
Assessment, AIAA Journal, 32, (8), 1994, 1689 1699.
6.176 Kwon, Y.B. and Hancock, G.J., Tests of Cold-Formed Channels with Local and
Distortional Buckling, ASCE Journal of Structural Engineering, 117, (7), July 1992,
1786 1803.
6.177 Hancock, G.J., Kwon, Y.B. and Bernard, E.S., Strength Design Curves for Thin-
Walled Sections Undergoing Distortional Buckling, Journal of Constructional Steel
Research, 31, 1994, 169 186.
7
Arches and Rings

7.1 Background
The analysis of the stability behavior of transversely loaded rings and arches
is presented and discussed in Sections 2.1.9 and 2.1.10. It is emphasized in
Section 2.1.9 that buckling behavior of rings and high rise arches, which is assumed
to be “inextensional”, completely differs from that anticipated in Section 2.1.10 for
shallow arches. In the latter case the elastic instability is primarily affected by the
axial thrust induced in the arch, and thus the buckling strains are “extensional”.
Obviously, in this case buckling may occur at a smaller load than that for
extensionless buckling. It is shown in these Sections, as well as in [2.32] that,
depending on the arch rise parameter K, type of loading and boundary conditions
either symmetric or asymmetric buckling criteria may govern the behavior of the
arch. The arch may experience no buckling, limit load behavior or bifurcation
buckling. The distinction between high rise arches and shallow arches has already
been emphasized in [7.1].
Though considerable attention was given to investigations on the stability of
arches, particularly during the fifties and sixties, most of these studies focused
on the stability of shallow arches. Studies on the stability behavior of high rise
arches are quite scarce. In addition to the investigations of Bresse [2.26] and Lévy
[2.27] on buckling of thin circular rings and those of Hurlbrink [2.30] and Timo-
shenko [2.31] on the stability behavior of high rise arches, which are mentioned
in Section 2.1.9, “inextensional” buckling of arches was studied by E. Chwalla,
R. Mayer, E. Gaber and E.L. Nicolai (references to original papers are given in
[2.1]). More recent studies, which are concerned with steep arches, were presented
by Huddleston [7.2] and [7.3] and by Lo and Conway [7.4] and [7.5]. As a matter
of fact, formulation of the theory in these studies allows analysis of the buckling
behavior of arches with any height-to-span ratio. When discussing the behavior of
deep arches under concentrated loads it should not be forgotten that they may as
well be susceptible to sideways, in the plane of the arch, buckling. This problem
is treated in [7.6].
The shallow arch, though being a simple structure, has many of the peculiar
nonlinear characteristics of the more complicated shell structures. Most of the

Buckling Experiments: Experimental Methods in Buckling of Thin-Walled Structures: Basic Concepts, Columns, Beams and
Plates – Volume 1. J. Singer, J. Arbocz and T. Weller Copyright © 1998 John Wiley & Sons, Inc.
410 Arches and Rings

Figure 7.1 Unstable symmetric type behaviour corresponding to perfect and imperfect thin
shells and arches (a) and degrading effect of geometrical imperfections (b) (from
[4.15] and [7.13])

features of elastic instability theory can be inferred from it ([7.7], [7.8], [4.14],
[4.15], [2.32], [7.9] [7.13]). Therefore, rather than determining the buckling loads
of arches, most of the numerous studies on their buckling conducted during the
fifties and sixties were undertaken to explain the perplexing buckling and post-
buckling behavior of the more complex thin shell structures with the aid of the
relatively simple shallow arch structure.
Roorda ([4.14], [4.15] and [7.13]) emphasized that experimental verification of
the unstable symmetric type behavior of thin shell structures of Figures 7.1a and b
is extremely difficult to achieve, especially if imperfections are taken into consider-
ation. (Figure 7.1a represents the responses of a “perfect” and geometrically imper-
fect shell. Figure 7.1b depicts the influences of geometrical imperfections on the
buckling capacity.) However, experimental validation, qualitative and sometimes
even quantitative, of this type of behavior for a simple arch structure is feasible.
The next sections in this chapter will deal with experimental studies first on
“shallow” arches, then on rings and high rise arches and finally, on lateral buckling
of relatively deep cross-section arches.

7.2 Shallow Arches


7.2.1 Arches Under Concentrated Loads
Experimental studies on relatively shallow arches, for which the rise (23.2 cm) over
span length (180 cm) ratio equals 1/7.75 have already been reported by Gaber, in
1934 [7.14]. In this investigation the arch specimens were made of flat St 80 steel
Shallow Arches 411

strips (E D 2, 072, 500 Kg/cm2 ) with a rectangular cross-section 30 ð 4 mm2 . The


arch shape was formed by bending the steel strips in cold condition with the aid of
a wooden jig into a symmetric quadratic parabola. This shape was chosen since in
a parabolic arch subjected to uniform loading along its span only axial compression
prevails in the arch and thus it was assured that the axis of the arch coincided with
the line of the supports.
The arches were either simply supported or clamped. They were uniformly
loaded by seven equal single loads applied at equal distances. The loads were
introduced through tension members, which were attached to the arch by clamping
rings at their upper end and loaded at their lower end by suspended buckets filled
with steel sawdust (see Figures 7.2a and 7.2b). The sawdust was slowly and evenly
distributed into the buckets by means of a special distribution arrangement. The
vertical and horizontal displacements at the load application points were monitored
with the aid of a telescope.
Prior to starting a test, the weights of the unequal tension members were compen-
sated by introducing sawdust into the corresponding buckets. Then, five to eight
loading steps were applied to reach the critical load, by adding to the buckets exact
pre-measured amounts of sawdust at each load step.
It is interesting to note that the degrading influence of geometrical imperfections,
as well as loading irregularities were realized in this investigation. Hence, in order to
load the tested arch as close as possible by vertical compression only and to avoid any
bending at incipient buckling, and compensate for inevitable irregularities in arch
shape and loading, as well as visible deviations of the arch axis from a parabola,
small additional loads were added. It was recognized that had this not been done,
these relatively small irregular deformations would have increased so rapidly and
subsequently the bending moments would have grown, so that the arch would have
lost its load carrying capacity at a relatively low load level, not necessarily by
buckling but perhaps also as a consequence of excessive bending moments.

Figure 7.2 Pre- and postbuckled shapes of arches in the tests of Gaber (a) simply supported
shallow arch (b) clamped shallow arch (from [7.14])
412 Arches and Rings

Another observation worth noting in this study, is the identification of a vibration


that the arches experienced after every symmetric or asymmetric loading, as a
means to detect buckling and to determine buckling modes. It was observed that
the closer the load was to the critical one, the slower the arch oscillated around
its equilibrium position, till finally it buckled with an infinitely large vibration
period. Furthermore, in the beginning of loading, the “stiff” unloaded arch vibrated
vigorously around its equilibrium position after applying the load. With increase
in load, the arch became weaker and “tired”, so that the near exhaustion of the
arch was apparent. It was observed that the vibration pattern corresponding to this
situation presented the future buckling mode of the arch. Typical buckling modes
of simply supported and clamped arches are depicted in Figures 7.2a and 7.2b.
Timoshenko and Gere [2.1] indicate that the experimental results of [7.14] are
in good agreement with Eq. (7.27) in their book
EI
qcr D 4 7.1
3
where qcr represents the critical value of the intensity of a load uniformly distributed
along the span of the arch, values of 4 are given in Table 7.5 of [2.1], and E, I
and  are the elastic modulus, the moment of inertia of the arch cross-section and
the arch span, respectively.
An intensive study to investigate the buckling behavior of small rise arches for
which the buckling deformations are “extensional” was undertaken by Fung and
Kaplan [7.1]. The study was concerned with the buckling behavior only and aimed
at demonstrating that the critical load is significantly affected by the curvature
of the arch. Two buckling criteria were used in the analysis: the classical one
which is based on the stability with respect to infinitesimal displacements about
the equilibrium positions, and Tsien’s energy criterion [7.15] which is based on
finite displacements and yields critical loads much lower than those obtained from
classical theory. To resolve the issue which one of these criteria agrees with the
real practical situation, an experimental study was conducted.
Tests with pin-ended arches with rigid simple supports, centrally loaded by a
concentrated load were performed. The test apparatus is shown in Figures 7.3a
and 7.3b. To simulate ideal end conditions the arches were supported on knife
edges mounted in a heavy steel frame (Figure 7.4). The stiffness of the frame
was 100 times that of the specimen. Calculations showed that allowing a reduc-
tion as high as 20 percent of this stiffness, due to flexibility of the knife edges
and fittings, insignificantly affected the buckling load. Therefore, considering the
supports as perfectly rigid, introduces a negligible error. It is observed in Figure 7.4
that the knife-edge fittings had sockets which aligned the specimen ends with the
knife edges.
The authors emphasized that spacing of the supports was the most critical problem
in setting up the arch for testing. An appreciable error in the buckling load will stem
from any looseness or initial compression because they change the initial arch shape.
To eliminate the undesirable “play” between the arch and the knife edges the spacing
was adjusted by a wedge controlled by a screw (Figures 7.3a and b).
Shallow Arches 413

(a)

Figure 7.3 Fung and Kaplan’s test apparatus (a) schematic of test rig (b) with specimen in
place (from [7.1])
414 Arches and Rings

Figure 7.4 Fung and Kaplan’s knife-edge fitting for simulation of ideal end conditions (from
[7.1])

The specimens were rolled from strips cut from 2024-T3 and 7075-T6 aluminum
sheets and milled to 1/2-in. width. To reduce the roll eccentricity several passes
were made at each setting of the rolls. At the start of each pass the rolls were
indexed to a new position.
The curvature of each specimen was measured at 12 stations by a dial gage
placed between knife edges 2 in. apart. These curvatures were numerically inte-
grated to obtain the shape of the specimen for which a 12-term Fourier expansion
was made. The actual measured central rise of the arch was compared with that
predicted by the numerical integration. The rise predicted by numerical integra-
tion of each specimen deviated from the actual one by no more than 4 percent,
and by no more than 1 percent from that predicted by the Fourier coefficients.
In calculating the theoretical critical load the Fourier coefficients 1 , 2 and 3
were used.
As in [7.14], it was noticed that vibration of the specimen provided the clearest
indication of the onset of buckling as well as information about the buckling mode.
Even very careful application of load resulted in slight vibration in the fundamental
mode, and a rapid decrease in the frequency of vibration was experienced when
approaching the critical load within a few pounds. Further load applications were
made then in extremely small loading increments.
The ratio of the test results Rcr (critical load) to that yielded according to the
classical criterion, as well as that of Rcr given by the energy criterion to that given
by the classical criterion, as a function of 1 are shown in Figure 7.5 which was
Shallow Arches

Figure 7.5 Comparison of theoretical predictions, yielded by the analysis developed by Fung and Kaplan, with the experimental results observed in
their experimental studies (from [7.1])
415
416 Arches and Rings

reproduced from [7.1] (1 D 21 a1 L/r, the first coefficient in the 12-term Fourier
coefficient used to describe the initial shape of the arch, where L is the arch span
and r is the radius of gyration of the arch cross-section). It is apparent from this
figure that the test results agreed quite well with results based on the classical
theory for higher values of 1 , but dropped appreciably below them for the lower
values of 1 . Nevertheless, all test results were higher than those predicted by the
energy criterion. Furthermore, the test results indicate that buckling will take place
only for 1 > 1.38, whereas the analysis predicts that buckling should already
occur for 1 ½ 1.05.
A calculation of stresses in the arches at buckling was performed to determine
whether yielding occurred. It was found that the maximum stresses at buckling in
all of the specimens tested were well below yield. Yielding was however obtained
in post buckling for all the specimens, except those having the lowest values of 1 .
In [7.8] Gjelsvik and Bodner studied the energy criterion for systems that exhibit
snap buckling. It was shown that the equal-energy load under “dead weight” loading
is a lower bound on the snap buckling load for elastic systems experiencing transi-
tional unstable buckling, or influenced by a particular class of destabilizing initial
geometric imperfections. To illustrate the behavior of a nonlinear system that under-
goes snap buckling, a shallow clamped arch under a concentrated central load was
theoretically and experimentally analyzed. As indicated by the authors, the fully
clamped end conditions were chosen because from an experimental point of view
better control is obtained with these boundaries. The experimental part of the inves-
tigation was concerned with obtaining complete load-deflection curves for a wide
range of arch geometries. Special attention was given to obtaining the unstable
regions of the response and to the investigation of the mode shapes at various
stages of loading. A schematic drawing of the test apparatus, which was employed
in [7.8], is depicted in Figure 7.6.
In the test frame, two fixing blocks were used to clamp each end of the arch.
Because the tests covered a wide range of geometries, a series of blocks of different
angles was manufactured. The fixing blocks were bolted to a stiff auxiliary frame
resting on the main frame. The fixing blocks were also blocked against two other
blocks that were bolted to the auxiliary frame to assure complete rigidity of
the clamping. Several dial gages were employed to measure the displacements
of the arch. The extension rods attached to these dials were counterweighted
with springs in order to minimize the effect of their weight on the response of
the arch.
The load at the center was introduced through a knife edge fixture and an
extension rod connected to it. To achieve the goal of obtaining the complete load-
deflection curve, a system of a jack and a strain gage instrumented load cell in
series with the applied load was employed. The load, which was greater than the
upper buckling load PU (limit load) in the load deflection curve, rested on the
jack (see Figure 7.6). Lowering the jack in small increments and noting the load
experienced by the arch from the load cell and the corresponding deflection from
the dial gages, the entire load-deflection relationship was obtained (in effect this
presents a controlled deflection procedure).
Shallow Arches 417

Figure 7.6 Schematics of Gjelsvik and Bodner’s setup and concept for determination of
complete load-deflection curves of shallow arches (from [7.8])

The arch specimens were fabricated from 2024-T4 aluminum alloy. They were
1 in. wide, 3/16 in. thick and had a nominal span of 34 in. Their circular shape was
obtained by rolling through a three-roller sheet metal roll. Geometry of the arches
was determined after bolting them into the auxiliary frame and measuring their
central height. In some specimens, particularly at lower values of  ( D ˇ2 R/t,
where ˇ, R and t are half the included angle of the arch, radius of the center line
and thickness, respectively) the clamping resulted in some prestressing, because it
was difficult to have the angle of the fixing blocks exactly coincide with the base
angle of the specimens.
Fourteen specimens were tested for  ranging from 3.69 to 16.25. All the
tests were completely elastic. As indicated in [7.8] the experimental buckling
loads that were determined from the maximum points in the experimental load-
deflection curves were in good agreement with the two term anti-symmetrical mode
analytical solution presented in this investigation, except for the lowest values of .
418 Arches and Rings

The authors explained that the disagreement might have stemmed from clamping
prestresses. The comparison of the empirical and analytical results, represented
by the nondimensional buckling load PŁ (PŁ D PR/Et2 fˇ, where E represents
Young’s modulus and f the width of the arch beam), versus the parameter , is
shown in Figure 7.7, reproduced from [7.8]. It is worthwhile noting that the equal-
energy load is noticeably lower than the experimental buckling load. A typical
load-central deflection curve for  D 11.62, together with the theoretical curves for
the 2 terms symmetrical mode and 2 terms transitional mode, which were obtained
in [7.8], are depicted in Figure 7.8. It is observed in this figure that although the
buckling load values were in agreement with the tests, the load-deflection curves
were not conclusive, in so far as the buckling mode in the unstable region is
concerned. Theoretically, the anti-symmetrical mode should govern the buckling
process, but this was not reflected by the experimental results. As explained in
[7.8] this might have been due to the fact that the actual amplitude of the anti-
symmetrical mode is relatively small and thus could have readily been overlooked
in a superficial test.
Conway and Lo [7.5] compared the results predicted by the analysis developed
by them with the experimental data of Gjelsvik and Bodner. They demonstrated
quite good agreement between their predictions and the data of Gjelsvik and
Bodner. The limit load, as well as the loads in the unstable postbuckling region,
yielded by their analysis were however slightly higher than those observed in the
tests. Conway and Lo claimed that this discrepancy might have stemmed from the
nonsymmetrical deformations which were observed in the tests.

Figure 7.7 Comparison of theoretical predictions corresponding to various approximations,


yielded by the analysis developed by Gjelsvik and Bodner, with the empirical
results obtained in the tests conducted by them (from [7.8])
Shallow Arches 419

Figure 7.8 Theoretical and experimental load-deflection curves obtained by Gjelsvik and
Bodner (from [7.8])

Schreyer and Masur [2.32] have also compared the experimental results of
Gjelsvik and Bodner with their analytical predictions. They have shown that for
values of  < 11.5 the experimental buckling loads were bounded from below by
the equal energy critical loads and from above by the symmetric limit loads. For
 > 11.5 the critical loads were slightly higher than those corresponding to upper
asymmetric buckling loads.
The test apparatus of Gjelsvik and Bodner was in essence employed in two
additional test programs, that of Roorda, ([4.14], [4.15] and [7.13]) and that of
Cheung and Babcock, [7.9] and [7.10]. Modifications were, however, introduced
in each test program to comply with the objectives of the investigation.
The arch experiments of Roorda aimed at demonstrating the unstable post buck-
ling behavior described in Figure 7.1, bringing to light the effect of imperfections
on the buckling behavior and providing a check on the theory developed in [4.14]
and [4.15]. Two types of simply supported arches were tested in these experiments,
a shallow circular arch and a shallow prestressed sinusoidal arch. Simple supports
were achieved with the use of a knife edge filed on the end of the bar resting in a
Vee groove (Figure 4.23b), and as in [7.1] a rigid supporting base was employed
with one knife seat bearing against a wedge block, controlled by a screw to elim-
inate the undesirable “play” between the arch and knife edges. The arches were
420 Arches and Rings

fabricated from high strength steel strips to ensure elastic behavior, even at large
deflections.

a. Circular Arch

The circular arch was obtained by rolling the strip into a circular shape. It had a
1 in. by 1/32 in. nominal cross-section, a span of 24 in. and a rise of 1.55 in. at
the crown.
The vertical load at the arch crown was applied through a pair of heat-treated
knife edges resting in small Vee grooves to allow free rotation of the point of
loading (Figure 4.46). A mechanism was provided to adjust the lateral position of
the Vee grooves, thus introducing small eccentricities of loading by moving the
point of application of the load perpendicular to the line of action of the load.
The loading device (see Figure 7.9, from [7.13]) consisted of a spring balance
and a screw-jack arrangement (note that Gjelsvik and Bodner, Figure 7.6, used
an extension rod and a strain-gage instrumented load cell instead of the spring
balance). As indicated by Roorda, according to the analysis of Thompson, [7.16]

Figure 7.9 Roorda’s test setup for studying buckling and postbuckling behavior of arches
(from [7.13], courtesy of Professor J. Roorda)
Shallow Arches 421

this arrangement provides a semi-rigid loading device that permits tracing of at


least parts of the unstable postbuckling equilibrium curves, as well as the stable
curves. The spring balance, which is suspended from the load point, is attached
to a heavy weight resting on the screw jack. The load applied to the arch changes
by turning the screw-jack up or down, and the change in load is registered by the
balance. When the load applied to the arch reaches a maximum or a critical load
and becomes unstable, the spring balance automatically reduces the load sustained
by the arch so that the total system, arch plus loading device, reaches a stable
position. The weight of the spring balance exceeded the buckling load of the arch.
It was counterbalanced by applying a counterweight to the loading knife edges
through a string passing over a set of “frictionless” pulleys.
To comply with the aforementioned objectives of the experiments, which require
a more general deformation variable, that admits both positive and negative values,
the rotation  of a point on the arch, rather than the common practice movement of
the point of load application in the direction of the load, was chosen as a measure
of the buckling deformation. The rotation was measured optically by means of a
stationary beam of light reflecting off a small mirror, attached to the rotation point,
onto a scale.
In the first stage of the experiment the load versus the rotation  was traced
(Figure 7.10a). Since the aim of this stage of the experiment was to obtain an as
close as possible approximation to the ideal postbuckling behavior, it was necessary
to give the load a small initial eccentricity d0 to cancel the effect of the unknown
geometrical deviation from the perfect arch (see [7.17]). The experimental curves
in Figure 7.10a correspond to a given value of d (slightly larger than d0 ). It is
apparent from this figure that for a given value of d two experimental curves result.

Figure 7.10 Comparison of approximate theoretical predictions and test results obtained by
Roorda (a) load rotation curves, (b) influence of load eccentricity on critical load
(from [4.14], [4.15] and [7.13])
422 Arches and Rings

One is the natural equilibrium curve emanating from the origin. The other, the
complementary one, was obtained by suitably adjusting the load W and manually
forcing the arch to buckle in the direction opposite to the natural direction of
buckling. Once the system was at a point on the complementary equilibrium path,
additional points on this path were obtained just by changing the load.
The approximate theoretical curve which is superimposed in Figure 10a repre-
sents the first order approximation to the initial postbuckling path of an ideal arch.
Following [4.14] it is given by:

P 302 a4
D1C 2 7.2
Pcr š81 š 3
here Pcr is the buckling load corresponding to a “perfect” arch, a D /20 , 0
determines half the central angle of the arch and  measures the rotation of the
crown of the arch. (Note that the non-dimensional force W D P/EA0  rather
than the applied force P is used in this figure).
In the second stage of the experiment the influence of load eccentricity was
determined. Treating d/L as the non-dimensional eccentricity, the variation of
the critical load versus eccentricity presented in Figure 7.10b was obtained. The
theoretical approximation presented in this figure is given in [4.14] by
 1/3  
PN Ł 3 3L 2 a2 d 2/3
D1C 7.3
Pcr 2 šR2 1 Ý 3 L

where PN Ł is the critical load of the eccentrically loaded arch. This figure serves as
well to determine the value of d0 .

b. Sinusoidal Arch

The prestressed sinusoidal arch was made by buckling of an initially straight steel
strip, with a 1 in. by 1/16 in. cross-section, into a near sinusoidal arch form. The
strip length was such that the resulting arch had a span of 24 in. and a rise of 1.5 in.
at the crown. Due to the fabrication process, the arch was in effect preloaded by
a thrust equal to the Euler load corresponding to the 24 in. long simply supported
strut of the same cross-section, and a bending moment that varies approximately
as a half sinus wave along the arch. The maximum initial moment at the crown
equals the rise times the thrust. Unlike for the circular arch no counterweight was
necessary in the test with the sinusoidal arch to counteract the weight of the spring
balance, because the buckling load exceeded this initial dead weight.
The tests and procedures with the prestressed arch were identical to those
employed with the circular arch. The test results, more or less, duplicated qualita-
tively those experienced with the circular arch in Figures 7.10a and b. However, in
the tests with this type of an arch, an additional relationship between the vertical
deflection of the crown  and the nondimensional load W was studied. This rela-
tion is presented in Figure 7.11. As was theoretically shown in [4.14], this figure
Shallow Arches 423

Figure 7.11 Experimental results obtained by Roorda for a shallow prestressed sinusoidal arch
(a) load-vertical displacement of the crown and the arch, (b) load-rotation of the
crown of the arch, (c) influence of load eccentricity on the critical load (from
[4.14], [4.15] and [7.13])

reveals the existence of a linear variation of the applied load with  after the outset
of buckling. Thompson and Hunt [7.17] showed that the test results of Figure 7.11
were in good agreement with their theoretical predictions.
Cheung and Babcock, [7.9] and [7.10], chose the simple arch structure as a
means to demonstrate the complicated buckling and post-buckling behavior of
shell type structures. As indicated by them [7.9] their experimental set up was
based on Gjelsvik and Bodner [7.8]. Like in Roorda, the heavy loading weight
in their loading apparatus was attached through a load cell to a spring, which
was suspended from the loading knife edge, and provisions existed to introduce
loading eccentricities. However, unlike Gjelsvik and Bodner, or Roorda, the heavy
weight in their loading rig rested on a hydraulic jack. This allowed slow continuous
loading of the arch, once the valve of the jack was opened.
The arch specimens were cut from 1/16 in. thick 2024-T3 aluminum sheet,
trimmed by a milling machine to 3/4 in. wide and then the strips were rolled to
424 Arches and Rings

Figure 7.12 Apparatus for initial geometrical imperfection measurements employed in the tests
of Cheung and Babcock (from [7.9] and [7.10])

approximately 30 in. radius in a three roll roller. Following the rolling process, the
arches were heat treated for eight hours at 375° F. Then the arches were mounted
into a heavy steel frame and their initial geometrical imperfections were measured.
A pendulum like apparatus was used to measure the imperfections (Figure 7.12).
It consisted of a fixed center and a rotatable arm, which could be adjusted to the
appropriate nominal radius of the arch. A dial gage was installed at the tip of
the arm.
Measurements were made by first adjusting the arm to the appropriate nominal
length. Then, starting from one clamp of the arch, dial gage readings were taken at
1/2 in. intervals along the arch. This furnished the deviation of the arch shape from
the preset nominal radius. Applying the “Least-square-method” to the measured
data, the “best-fit radius” and “best-fit imperfections” were determined.
Following the imperfection measurements, the test was carried out. It should be
noted that the arch was reusable and was reloaded under different load eccentric-
ities. The test results yielded by the four arches of various span length that were
tested in the program (see geometric description in Table 1 of [7.10]) are presented
in Figures 7.13 and 7.14. The test results in Figure 7.13 reveal the degrading effect
of load offset for each arch and like Roorda (Figure 7.10) experience the cusp type
behavior. It is observed from this figure that this effect is especially sensitive in
the vicinity of the arch center due to the cusp which forms at ε D 0. Good agree-
ment of the test results, for arches with no load eccentricity (Figure 7.14), with the
classical theoretical results of Schreyer and Masur [2.32] was demonstrated. The
results were higher than those observed by Gjelsvik and Bodner in [7.8].
Shallow Arches 425

Figure 7.13 Effect of load eccentricity on nondimensional critical load observed in the tests
of Cheung and Babcock (from [7.9] and [7.10])

More recently Kawashima and Ito [7.18] used a specially designed testing
unit to experimentally investigate the asymmetric snap-through behavior of pin-
ended circular arches under controlled deflection. The test apparatus is depicted
in Figure 7.15 and it features a special forcing device of the arch ends which can
hold compression, as well as tension reaction during snap-through. To keep the
arch span constant and to measure the thrust at the ends of the arch, the testing
unit was installed in a MTS servohydraulic testing machine.
The arch specimen (1 in Figure 7.15, see also Figure 7.16) is held in the
supporting shafts (5 and 6 in Figure 7.15, see also Figure 7.17), which freely
rotate in bearing units attached to the crosshead and the actuator rod of the testing
machine. Before snap-through, the end of the arch is supported on the bottom face
of the groove in the supporting shaft 5. Since the shaft axis lies in the bottom
face, the arch ends rotate around the axis during snap-through. In the later stage
426 Arches and Rings

Figure 7.14 Comparison of experimental buckling loads obtained by Cheung and Babcock
with the analytical and test results of Gjelsvik and Bodner [7.8] and the theoretical
predictions of Schreyer and Masur [2.32] (from [7.9] and [7.10])

of snap-through, the tension reaction is maintained by the key slot in the shaft and
the key which is cemented to the arch (see Figure 7.17).
The transverse load, which was deviated by 10 mm from the arch center, was
introduced into the arch through the load bearing 2 as a reaction to the deflection
provided by the screw mechanism 4. The central deflection w0 , the transverse load
P and the thrust Q were measured by the transducer q and the load cells 3 and 8,
respectively and were plotted on an X Y recorder. Two displacement transducers
10 and 11 were used to delineate the shape of the deformed arch over 100 mm of
its central portion. This was done at several stages during the snap-through.
The arch specimens (Figure 7.16) used in the tests were machined from a 2017
aluminum alloy bar and were annealed in vacuum at 410° C for two hours to relieve
residual stresses. The widened end portion of the specimens were required to secure
enough adhesive strength between the key and the arch specimen.
Comparisons of the experimental results with the analytical predictions, which
were obtained by the numerical model developed in [7.18], are shown in
Figures 7.18a and 7.18b. As indicated by the authors and can be observed in
these figures, the experimental results agree with the numerical ones. Experimental
shapes of the deformed arch obtained at different stages of snap-through are
Shallow Arches 427

Figure 7.15 Kawashima and Ito’s test apparatus with a special forcing device of the arch ends
to hold compression as well as tension reaction during snap-through (from [7.18])

Figure 7.16 Kawashima and Ito’s arch specimen (from [7.18])

compared with numerical ones in Figure 7.19. Again, the experimental results
correlate well with the numerical predictions.

7.2.2 Arches Under Uniform Pressure Loading


The problem of the buckling and post-buckling behavior of an elastic orthotropic
radially reinforced shallow arch subjected to uniform pressure was studied by
428 Arches and Rings

Figure 7.17 Rotating supporting shafts employed in the tests of Kawashima and Ito (from
[7.18])

Figure 7.18 Comparison of theoretical predictions and experimental observations in the inves-
tigation of Kawashima and Ito (from [7.18])

Hsu and Nash in [7.11] and [7.12]. Experiments were performed to investigate the
validity of the various analytical models presented in these studies. The influence of
transverse shear, transverse normal stress, finite deflections, and initial geometrical
imperfections which may lead to bifurcation of the arch were included in the
investigations.
Shallow Arches 429

Figure 7.19 Comparison of the deformed shapes of the arch predicted by the analysis of
Kawashima and Ito with the deformed shapes experienced in their test program
(from [7.18])

The experiments were performed with arches composed of resin matrices in


which small diameter glass fibers, each oriented towards the center of curvature of
the arch were embedded. This type of construction, which was first manufactured
by Uniroyal, Inc. [7.19] made the arches well-suited to withstand loadings normal
to their convex surface, like those experienced in deep submergence pressure hulls.
The radially reinforced arches were fabricated from a glass impregnated type
of high strength S glass roving embedded in 1009-26-S epoxy resin (marketed by
Minnesota Mining and Manufacturing Co.). Curing and post-curing of the arches
is described in detail in [7.11] and [7.12].
The test apparatus consisted of a loading fixture which included end supports
for the arch, a system to provide lateral pressure on the convex surface of the arch
and three Linear Variable Differential Transformers (LVDTs Schaevitz type 300-
SSLT, with a range of š 0.3 in. and linearity larger than š 1 percent of the full
range output) which were epoxied to the concave side of the arch for measuring
the radial deflections of the arch.
In the tests, the ends of the arch were clamped with two metal blocks tightened
around the arch by means of two bolts, one of which went through the end of the
arch (Figure 7.20). A slot corresponding to the curvature of the arch was milled
into the blocks so as to provide good clamping. This approach for simulating a
clamped support was validated by testing a straight cantilever beam.
The pressure loading system (Figure 7.21) consisted of an oil jack, a pressure
chamber and a pressure gage. The pressure chamber was formed above the convex
side of the arch by bounding that side by a concave shaped piece of hard wood
together with two 0.5 in. thick plexiglass sheets on either side of the arch. These
sheets were bolted together through the hard wood piece, as well as through the
arch supports (Figure 7.21). The uniform pressure was applied by a bicycle-type
inner tube which was placed in the space between the convex side of the arch and
the concave surface of the hard wood piece. The tube was connected to the oil
jack (Hein Werner Push Master Hydraulic Pump Unit Model FP-4) and hydrostatic
pressure was applied by pumping oil into the flexible tube. Since it was found that
430 Arches and Rings

Figure 7.20 Arch with clamped supports used by Hsu and Nash (from [7.11] and [7.12],
courtesy of Professor W.A. Nash)

Figure 7.21 Hsu and Nash’s pressure loading system (from [7.11] and [7.12], courtesy of
Professor W.A. Nash)

ordinary jack oil quickly damages the inner tube, paraffin oil was used instead.
Such a problem was also encountered in the Technion annular plate test rig that
employed a somewhat similar loading scheme (see Section 8.2.8 of Chapter 8).
A calibrated pressure gage capable of monitoring pressures up to 200 psi was
connected in series between the oil jack and the tube.
Hsu and Nash compared the results obtained in the experiments with those yielded
by the various theories presented in [7.11] and [7.12]. It was found that collapse of
the arches was in good agreement with the theoretical predictions which were based
upon an asymmetric buckling mode. The experimentally obtained arch displace-
ments were, however, somewhat greater than those predicted by the theories.
Shallow Arches 431

Chini and Wolde-Tinase introduced a very peculiar method for applying uniform
pressure loading. This was accomplished by employing centrifuge model testing for
studying buckling and postbuckling behavior of prestressed shallow and high-rise
arches that make up the skeleton of large size clear-span prestressed domes under
self-weight [7.20]. Centrifuge model testing is based on the fact that inertial forces
and gravity forces are equal. So, the earth’s gravity, the main agent of instability
for large-size clean space enclosures, was scaled by means of centrifugal force. In
order to provide a small scale-model with the same state of stresses existing in the
actual structure at corresponding points, the weight of the model must be increased
N times, where N is the ratio of the prototype size to the model size.
A noteworthy characteristic of a centrifuge testing is that the model continues to
be loaded with the same forces after buckling, thus providing a means for studying
the postbuckling behavior of the model.
In the tests, the prestressed arch models were attached to the swinging platform
of the centrifuge and their self-weight was increased by accelerating the centrifuge.
Seven prestressed arch models fabricated from 0.8 mm thick 2024-T3 aluminum
sheets with a span length of 502 mm were tested. The span length was limited by
the dimensions of the swinging platform inside the centrifuge. The thickness was
chosen so as to ensure that the normal stresses in the extreme fibers of the arches,
after prestressing, will not exceed 50 percent of stress of the material yield, thus
ensuring elastic buckling. The centrifugal forces on the model reached their critical
values when the centrifuge was running with more than 100 rpm (at this speed the
normal gravitational acceleration of 1 g was negligible as compared with the 15 g
horizontal centrifuge acceleration).
The centrifuge used in the tests (Figure 7.22) was a Genisco model 1230-1
G-accelerator with arm and swinging platform assembly and capacity of 13,620 g-
Kg. A zoom camera was installed near the center of rotation on the centrifuge
arm for continuous visual monitoring of the arch. The camera was connected to
a VCR and video monitor so that video recording of the experiments could be
made. Downward deflection of the model during tests was observed with the aid
of an angle mirror assembly that was attached to the swinging platform. The mirror
provided maximum coverage of the platform area and permitted side viewing of
the model for downward displacement measurement.
The overall setup for the model in the centrifuge is shown in Figure 7.23. The
arch was supported by two hinges located at the edge of an aluminum plate, which
was used to attach the model to the swinging platform.
Four strain gages were bonded to the top and bottom surfaces of the model
to measure normal strains at selected points. Strain gage measurements were also
taken while the specimen was in a flat configuration prior to prestressing and
forming the desired arch shape by buckling. After buckling the specimen into
the arch shape it was pinned to the aluminum plates, and another strain reading
was taken. Then the centrifuge was started and slowly accelerated to minimize
tangential acceleration. Centrifuge speed and strain readings were simultaneously
recorded at various loading increments. Deformation and buckling processes of the
specimen were viewed and recorded through the television monitor and the VCR.
432 Arches and Rings

Figure 7.22 Centrifuge arm assembly employed by Chini and Wolde-Tinase for application of
uniform pressure loading (from [7.20])

The distributed applied load is given by

q D 10.97 ð 103 Arn2 7.4

where q(N/m) is the intensity of the load at a given point, A(m2 ) is the cross-
sectional area of the model at a prescribed point, (Kg/m3 ) is the density of the
model, n is the angular velocity and r(m) is the radial distance of a prescribed
point from axis of rotation.
The weight of the model was increased through centrifugal acceleration which
eventually resulted in buckling of the model.
The test results were compared with finite element calculations. It was found
that the finite element predictions were slightly lower than the experimental obser-
vations. This was attributed to the partial rotational resistance of the hinges used
in the experimental models, whereas they were assumed to be pinned in the finite
element system. A load deflection curve obtained in the tests is compared with the
predicted one in Figure 7.24.
Shallow Arches 433

Figure 7.23 Setup of arch model, tested by Chini and Wolde-Tinase, inside centrifuge swinging
arm (from [7.20])

Figure 7.24 Load-deflection curves comparison of finite element predictions performed by


Chini and Wolde-Tinase with their test results (from [7.20])
434 Arches and Rings

7.2.3 Additional Empirical Investigations

Additional noteworthy extensive experimental studies on buckling of shallow


arches were performed by Evan-Iwanowski and Mitchell [7.21], Dickie and
Broughton [7.22] and Saitoh and Kuroki [7.23].

7.3 Rings and High Rise Arches

7.3.1 Rings Contact Buckling

An elastic ring surrounded by a rigid circular confinement surface and subjected to


compression hoop stresses stemming from inward movement of the rigid confine-
ment, thermal expansion, etc. may buckle as depicted in Figure 7.25. Buckling
will take place when the compression hoop stress reaches a critical value under
which part of the ring snaps inwards and becomes free of pressure. This buckling
phenomenon is known as shrink or contact buckling. Because the deflection of the
ring is constrained in one direction it is also called one way buckling. Buckling
problems of this type may arise in mechanical designs which involve slewing the
inside of cylinders, pumps, or other types of pressure components and storage
containers, as well as in repairing components by installing a liner. One sided
buckling was studied by Lo et al. [7.24], Chan and McMinn [7.25], Sun et al.
[7.26] and [7.27].
Experimental studies on the instability of confined rings were carried out by
Sun et al. [7.26] and [7.27]. The objective of the experiments was twofold: to
investigate the effect of geometrical imperfections on the buckling behavior of the
ring and verify the relation between the critical load and initial center deflection

Figure 7.25 One way buckling of an elastic circular ring in a rigid confinement
Rings and High Rise Arches 435

Figure 7.26 Parameters of unbuckled ring

of the ring established in the analysis of [7.26]; to establish the relationship of the
buckling load with the dimensionless variables (R/t) and (υ/t) where R,t and υ are
the ring radius, thickness and initial deflection, respectively (see Figure 7.26).
In the experiments an initial imperfection in the rigid boundary was imposed
to introduce the center deflection of the ring, since it is much easier to control it
quantitatively and it is more reliable than one introduced by an external disturbance.
To simulate point imperfections on the rigid confinement υ, wires of different
diameter were inserted between the ring and rigid confinement, thus imposing
different center deflections on the ring (Figure 7.28).
The controlling parameter of the buckling load is the pressure between the ring
and outside confinement, which in reality is set up by a designed dimensional
mismatch between the ring and confinement. Practically it is impossible to obtain
a prescribed mismatch, because small uncontrolled tolerances are unavoidable in
machine shop practice. Furthermore, to obtain the critical pressure for a given
initial center deflection on the ring, continuous change of the mismatch is required.
This makes the experiments impractical and expensive. Sun et al. circumvented this
difficulty by employing a semi-circular boundary instead of a full one (Figure 7.27).
In the assembly of Figure 7.27 the load P is related to the pressure q between the
ring and confinement by the following formula

P D Rq. 7.5

By applying a vertical displacement at the ends of the semi-ring, the load P and
subsequently the pressure q change. In this manner, for a given center deflection
of the ring the pressure q can be continuously changed until the critical buckling
pressure is reached.
The arrangement of the experimental apparatus developed by Sun et al. is
depicted in Figure 7.28. The machined surface of the rigid confinement was
436 Arches and Rings

Figure 7.27 Half ring and rigid confinement assembly introduced by Sun et al. for continuous
changing of the pressure q between the ring and confinement (from [7.26] and
[7.27])

Figure 7.28 Sun et al.’s experimental apparatus for application of continuous increasing
uniform pressure loading (from [7.26] and [7.27])

polished to reduce friction between the ring specimen and the rigid support.
The apparatus was mounted in a servocontrolled hydraulic testing machine under
displacement control and the vertical load P was continuously measured in
response to the end displacement.
In the tests of [7.26], steel, aluminum and cardboard rings of different thicknesses
and width were tested, and nine wires of different diameter were inserted between
Rings and High Rise Arches 437

the ring and rigid support. In the tests of [7.27] three rigid supports with different
radii were used. The rings were fabricated from three different plastics of different
thickness and elastic modulus, and eight wires of different diameter were inserted
between the rigid support and the ring. As a result, information on buckling of rings
over a wide range of (R/t) and (υ/t) ratios and moduli was obtained in the tests.
The tests were conducted as follows: first, a specimen was elastically bent into
position and pressed against the surface of the rigid support by introducing displace-
ment on both ends. The specimen was unloaded, a wire was inserted between the
rigid boundary and the specimen (Figure 7.28) and the specimen was reloaded
until buckling occurred. The load displacement curve was recorded on an X Y
recorder during both loading and unloading processes.
A typical load-displacement curve obtained in the tests of Sun et al. is shown
in Figure 7.29. In this curve the compression load increases together with the end
displacement from B to C. At PC the specimen suddenly snaps to the low load

Figure 7.29 Loading and unloading load-displacement curves observed in the tests of Sun
et al. (from [7.26] and [7.27])
438 Arches and Rings

PD in a very short interval of time. Then the load slowly decreases with further
increase in displacement until load PE is reached. From E to F the load slowly
increases whereas the end displacement decreases. When the end displacement
decreases to a certain value, the specimen suddenly snaps back to the unbuckled
position and the load increases to PF . Sun et al. indicated that snap back for a
specific material always occurred at the same point, even if the buckling point
was different for the different initial deflections. Consequently, they concluded: if
the strain in the specimen is lower than that corresponding to snap back, only the
stable unbuckled state would be possible; if the strain exceeds this point the stable
original unbuckled state and the stable buckled large deflection state are possible.
Buckling of the specimen, when reaching this point, depends on the imperfection
or the external disturbance.
A typical comparison of the test results of [7.26] with the results yielded by
the approximate analysis developed in this study is shown in Figure 7.30. It was
indicated by Sun et al. that the smaller (υ/t), the larger was the scatter in the
measured data. Furthermore, from curve fitting of the test results in the form
cr D Cυ/tz the power of (υ/t) was found to be about 3/4 for all of the different
materials used in the test program, whereas the analysis yielded a power of 1
(see Eq. 21, [7.26]).
A comparison of the test results of [7.27] with the dimensional analysis buckling
function presented in this study is depicted in Figure 7.31. It is apparent from this
figure that the experimental results are in good agreement with the dimensional
analysis predictions.

Figure 7.30 Comparison of test results observed by Sun et al. with the analytical results
predicted by their simplified analysis (from [7.26])
Rings and High Rise Arches 439

Figure 7.31 Comparison of test results observed by Sun et al. with dimensional analysis buck-
ling function proposed by them (from [7.27])

Theoretical and experimental studies of an elastic arch in a rigid cavity were


also carried out much earlier by Zagustin and Herrmann [7.28].

7.3.2 High Rise Arches

Huddleston studied analytically and experimentally the behavior of steep


prestressed “non-rigid” arches, which were made from buckled struts before
attaching them to their supports [7.2] and [7.3]. These investigations focused on
establishing the relation between the buckling and snap-through behavior of such
prestressed arches and the type of loading acting on the arch and its height-to-
span ratio.
In the test program, experiments were conducted on a model dome comprised
of four prestressed arches (Figure 7.32). The test results indicated that the dome
type structure had a significant load carrying capacity and, to a limited extent, they
verified the results yielded by the numerical analysis presented in [7.2]. It was
indicated in the study that because a restraint against unsymmetric buckling was
added to each arch by the intersecting arches of the dome construction, the model
could have carried a load slightly higher than the bifurcation load before snapping
through.
Additional noteworthy experimental investigations on buckling of high rise
arches are those of Gaber [7.14], which were discussed in Section 7.2, and those of
440 Arches and Rings

Figure 7.32 Huddleston’s arch model dome (from [7.2])

Dickie and Broughton [7.22] and Saitoh and Kuroki [7.23] that deal with shallow
arches as well, and were also mentioned in Section 7.2.

7.4 Lateral Buckling of Arches


Lateral buckling of arches is of great concern for bars of relatively “deep” cross-
section, which are bent in the plane of the greatest rigidity of their cross-section.
Such arches are common in bridge construction. Contrary to the investigations on
in-plane buckling of arches discussed in Sections 7.1 and 7.2, which served as a
means for studying the peculiar nonlinear characteristics of the more complicated
shell structures, experimental studies on out-of-plane buckling of deep cross-section
arches are aimed at validating the various analytical tools developed for predicting
their buckling capacity. Furthermore, each test program was performed only to
verify a theory that was derived in it, usually for one particular loading case.

7.4.1 Theoretical Background

A considerable amount of theoretical research has been performed on the out-


of-plane flexural-torsional stability problem of arches. Timoshenko was amongst
the first to address this issue and his solutions for two buckling problems are
presented in Timoshenko and Gere [2.1]. These two problems deal with a bar of
Lateral Buckling of Arches 441

Figure 7.33 Narrow rectangular cross section with a circular axis subjected to (a) bending by
equal opposite couples M0 , (b) uniformly distributed and radially directed load q
along the center line (from [2.31])

narrow rectangular cross-section with a circular axis submitted to either bending by


two equal and opposite couples M0 acting in the plane of the arch (Figure 7.33a),
or to a continuous load of intensity q, uniformly distributed along the center line
and radially directed (Figure 7.33b). In both cases the ends of the arch were simply
supported, i.e. the ends could freely rotate with respect to their principal axis of
inertia, while rotation with respect to tangents to the center line of the bar at the
ends was prevented.
It was shown that in the first case the arch would buckle under:

 
EIx C C EIx C C 2 EIx C   2
Mcr D š C 7.6
2R 2R R2 ˛
where Ix is the area moment of inertia with respect to the x axis and C is the
torsional rigidity of the cross-section.
For small values of the subtended angle ˛ and initial curvature of the arch and
substituting ˛R D , the formula reduces to:
EIx C C    
Mcr D š EIx C. 7.7
2R 
442 Arches and Rings

In the second case, the critical compressive force in the arch, at which lateral
buckling occurs is given by:
EIx 2  ˛2 2
qcr R D . 7.8
R2 ˛2 [2 C ˛2 EIx /C]
This formula corresponds to the case where the directions of the loads q do not
change during buckling, i.e. they are only displaced laterally parallel to their initial
direction.
When ˛ is small, ˛R D  and the Euler buckling load for a beam of length 
2 EIx
qcr R D 7.9
2

is obtained. For ˛ D the critical load becomes:
2
EIx 9
qcr R D 2 . 7.10
R 4 C EIx /C
This corresponds to the critical compressive force for a complete ring that buckles
in four half-waves. When ˛ D  the critical load equals zero since the arch can
rotate freely around the diameter joining its ends.
If during buckling of the arch, the loads q change their direction so as to always
be directed toward the initial center of the arch, the critical compressive force is
given by:
2 EIx 2  ˛2
qcr R D 7.11
R2 ˛2 [2 C ˛2 EIx /C]
and for a complete ring buckling into four half waves by:
EIx 12
qcr R D . 7.12
R 4 C EIx /C
2

(This formula was obtained by H. Henkey, see [2.31].)


It should be noted that the slight changes in the directions of the loads, raise the
stability of the ring considerably. This has already been mentioned in Chapter 2,
Subsection 2.1.9 and is discussed in [2.29].
Built-in arches, for which the loads retain their direction during buckling, sustain
a critical load given by:
EIx
qcr D 5 3 7.13
R
where 5 is a numerical factor that depends on ˛ and is provided in Table 7.9
of [2.31].

For ˛ smaller than this equation can be approximated by:
2
EIx 42  ˛2 
qcr R D 7.14
R2 ˛2 [42 C ˛2 EIx /C]
which for very small values of ˛ yields the Euler buckling load.
Lateral Buckling of Arches 443

Additional theoretical studies can be found in [7.29] [7.43]. Papangelis and


Trahair [7.44] review most of these studies and indicate that the solutions provided
by [7.36] [7.38] are very different from those reported by Timoshenko [2.31] and
Vlasov [7.30], whereas their studies [7.39] tend to confirm them.

7.4.2 Experimental Studies

Papangelis and Trahair proposed experimental evaluation of the different theo-


retical predictions. This was undertaken by them in the late eighties (see [7.44],
[7.42] and [7.43]). It should be noted that the investigations reported in [7.29] and
[7.34] which preceded those of Papangelis and Trahair, also included extensive
complementary experimental studies to validate the analyses developed. Additional
extensive experimental studies on out-of-plane buckling of arches are reported in
[7.45] and [7.46].
Papangelis and Trahair based their experimental program on the equipment and
test techniques developed by Trahair [7.47]. The test setup, testing procedures, test
specimens and their fabrication are discussed in [7.42]. The loading rig for testing
the arches, which was employed in the program, is shown in Figure 7.34. The
arches were loaded with a central concentrated force applied at the upper surface
of the top flange. Dead weight lead shot was used in all tests to maintain the
loads vertical when the arches deflected laterally. The shot was carried in buckets
and hardened steel pins, which rested on sockets, transferred the weight to the top

Figure 7.34 Papangelis and Trahair’s test arrangement for loading of beams with deep
symmetric and asymmetric cross section (from [7.42], courtesy of Professor
N.S. Trahair)
444 Arches and Rings

Figure 7.35 Load introduction into the arch and lateral deflection measurement arch cross-
section in the tests of Papangelis and Trahair (from [7.42], courtesy of Professor
N.S. Trahair)

flange (see Figure 7.35). The lateral deflections of the arch were measured by a
micrometer that was attached to the load hanger (Figure 7.35). This was done in
order to avoid, or at least reduce, significant lateral forces that might be exerted
by the use of the other types of measuring instruments, e.g. dial gage springs. The
magnitude of the micrometer forces was minimized by connecting the arch and
micrometer electrically to a device shown in Figure 7.34, so that once the tip of
the micrometer touched the load hanger the device would light up.
Two types of arch specimens were tested, one with doubly symmetric cross-
sections and one with non-symmetric cross-sections. The doubly symmetric arches
were tested in the normal and inverted attitude (Figure 7.36a). The arches with
non-symmetric cross-sections were tested in the normal attitude (Figure 7.36b),
with the outer flange being either larger or smaller than the inner flange.
In the tests with the doubly symmetric arches, the arches were simply supported
in-plane, i.e. their ends were free to rotate and deflect horizontally, but prevented
from deflecting vertically (w D 0). The non-symmetric arches were pin-ended
in-plane, so that their ends were free to rotate, but prevented from deflecting
horizontally or vertically (v D w D 0). All of the arches were simply supported
out-of-plane, thus allowing free rotations of their ends around radial axes as well
as warping. Lateral deflection and rotation about tangential axes were prevented
(u D  D u00 D 00 D 0). Figure 7.37 depicts an arch end support.
The arch specimens employed in the program were fabricated from high strength
aluminum I sections beams, with a high ratio of major to minor axis flexural
Lateral Buckling of Arches 445

Figure 7.36 Loading arrangement in the tests of Papagelis and Trahair (a) doubly symmetric
test arches, (b) nonsymmetric arches (from [7.42])

Figure 7.37 Arch end support in the tests of Papangelis and Trahair (from [7.42], courtesy of
Professor N.S. Trahair)
446 Arches and Rings

rigidities. The beams were gradually curved by a rolling machine until the desired
radii of curvature were obtained, center line radii of 1000 mm and 500 mm for the
symmetric arches and shear center axis of 500 mm for the nonsymmetric arches.
In the later specimens either the lower or upper flange was machined to approx-
imately half its original width. It should be noted that residual stress calculations
were performed to ensure that the residual rolling stresses will have an insignifi-
cant effect, if any, on the stability tests. These calculations revealed that residual
stresses were present primarily near the neutral axis, whereas buckling is primarily
controlled by the flange stresses. Also, subsidiary bending and torsion tests were
carried out to determine the experimental flexural and torsional rigidities of the
arches.
After setting up the arch in the loading rig, loads were applied in increments
which were gradually reduced as the load approached the critical theoretical load,

Figure 7.38 Typical load-deflection curve obtained in the tests of Papangelis and Trahair and
corresponding Southwell plot (from [7.42])
Lateral Buckling of Arches 447

Figure 7.39 Comparison of Papangelis and Trahair’s test results with their analytical
predictions doubly symmetric beam (a) in normal attitude, (b) in an inverted
loading attitude (from [7.42])
448 Arches and Rings

Figure 7.40 Comparison of Papangelis and Trahair’s test results with their analytical
predictions nonsymmetric beam (a) wide upper flange, (b) narrow upper flange
(from [7.42])

which was predicted by the Southwell plot technique and the analyses derived in
[7.39] [7.43]. After each load increment, the testing frame was subjected to a few
minutes of small amplitude vibrations. These vibrations were induced by a hand-
held electric drill to reduce the effects of static friction in the ball bearings at the
supports. Lateral deflection of the arch was measured after each load increment.
Loading of the arches continued until the applied loads approached almost the
critical values (within 5 10 percent of the extrapolated Southwell critical load).
Lateral Buckling of Arches 449

At this stage the test had to be terminated, since the lateral deflections began
increasing rapidly (see Figure 7.38). This avoided permanent deformations of the
arch and allowed a series of tests to be made on the same arch for different values
of the subtended angle.
Papangelis and Trahair reported very good agreement between their experimental
results and analytical predictions, as well as the experimental results of Tokarz
[7.45] and the theoretical predictions of Tokarz and Sandhu [7.32]. On the other
hand significant discrepancies existed between their experimental and theoretical
results and the predictions of Yoo [7.36] [7.38]. These discrepancies are apparent
from Figures 7.39 and 7.40.
One of the problems common to all buckling tests is the need to significantly
decrease the loading increments when approaching the critical load, because of
the fast increase of the measured displacements. This becomes a very complicated
task in the case of multiple loading, which is used to simulate uniform loading
of structures. Tokarz [7.45] and Ojalvo, Tokarz and Nontanakorn [7.46] addressed
this issued by employing uniform loads by the concept which is schematically
described in Figure 7.41. In this concept, incremental loading is applied to the
arch by equally spaced hangers. The upper ends of these hangers are attached to
the arch (Figure 7.42), so that they impart only concentrated forces through the
centroid of the cross-section. To the lower ends of the hangers, canisters (which
are partially immersed in water to a predetermined level from their bottoms) are
attached, thus subjecting each hanger to a prescribed buoyancy force.
Employing the above loading system, increments of loads were applied to the
arch by first adding equal amounts of lead shot to the canisters. Once the deflections
started growing very fast, further fine loading of the arch was obtained by slowly
and evenly reducing the buoyancy force on the canisters. This was obtained by
pumping water out of the water tank, in which the canisters were submerged, with
a centrifugal water pump.

Figure 7.41 Schematics of Tokarz concept of uniform loading and loading arrangement by
employing controlled buoyancy forces (from [7.45])
450 Arches and Rings

Figure 7.42 Typical loading hanger to arch attachment used in the lateral buckling tests of
Tokarz and Ojalvo, Tokarz and Nontanakorn (from [7.45])

It is apparent from Figure 7.41 that the arches could be subjected to two types of
loading, vertical loading and tilt loading. For tests where tilt loading was applied,
the loading setup was modified by placing a pair of laterally rigid steel bars at
the elevation of the arch end supports. These immovable bars enforced the wire
hangers to pass through them, and since the arch was free to displace laterally with
increase in hanger loads, the hangers became tilted.
It is worth noting that Di Tommaso and Viola [7.34] also controlled the incre-
mental loading in their tests by means of buoyancy forces, in a similar manner to
Tokarz [7.45].

References

7.1 Fung, Y.C. and Kaplan, A., Buckling of Low Arches or Curved Beams, NASA TN
2840, Nov. 1952.
7.2 Huddleston, J.V., Behavior of a Steep Prestressed Arch Made from a Buckled Strut,
Journal of Applied Mechanics, Transactions ASME, Series E, 37, (4), Dec. 1970,
984 994.
7.3 Huddleston, J.V., Finite Deflections and Snap Through of High Circular Arches,
Journal of Applied Mechanics, Transactions ASME, Series E, 35, (4), Dec. 1968,
763 769.
7.4 Lo, C.F. and Conway, H.D., The Elastic Stability of Curved Beams, International
Journal of Mechanical Sciences, 9, (8), Aug. 1967, 527 538.
7.5 Conway, H.D. and Lo, C.F., Further Studies on the Elastic Stability of Curved
Beams, International Journal of Mechanical Sciences, 9, (9), Oct. 1967, 707 718.
7.6 DaDeppo, D.A. and Schmidt, R., Sideways Buckling of Deep Circular Arches Under
a Concentrated Load, Journal of Applied Mechanics, Transactions ASME, Series E,
2, June 1969, 325 327.
7.7 von Kármán, T., Dunn, L.G. and Tsien, H.S., The Influence of Curvature on the
Buckling Characteristics of Structures, Journal of the Aeronautical Sciences, 7, 1940,
276 289.
7.8 Gjelsvik, A. and Bodner, S.R., The Energy Criterion and Snap-Buckling of Arches,
ASCE Proceedings, Journal of the Engineering Mechanics Division, 88, (5), Oct.
1962, 87 134.
References 451

7.9 Cheung, M.C., The Static and Dynamic Stability of Clamped Shallow Circular
Arches, Ph.D. Thesis, California Institute of Technology, Pasadena, California, 1969.
7.10 Cheung, M.C. and Babcock, C.J., Jr., An Energy Approach to the Dynamic Stability
of Arches, Journal of Applied Mechanics, Transactions ASME, Series E, 37, (4), Dec.
1970, 1012 1018.
7.11 Hsu, M.Y.H., The Influence of Transverse Shear Deformation and Transverse Normal
Stress on a Nonlinear Orthotropic Shallow Circular Arch with a Consideration of
Initial Imperfections, Ph.D. Thesis, University of Massachusetts, Amherst, Massa-
chusetts, 1972.
7.12 Hsu, M.Y.H. and Nash, W.A., Influence of Transverse Shear and Transverse Normal
Stress on Large Deflections of Orthotropic Arches, Report No. UM-72-4, School of
Engineering, University of Massachusetts, Amherst, Massachusetts, April 1972.
7.13 Roorda, J., An Experience in Equilibrium and Stability, RILEM, International
Symposium, Experimental Analysis of Instability Problems on Reduced and Full Scale
Models, Instituto Nacional De Tecnologia Industrial, Buenos Aires, 1971.
7.14 Gaber, E., Über die Knicksicherheit vollwandiger Bogen, Bautechnik, Jahrgang 12,
Heft 49, Nov. 1934, 646 656.
7.15 Tsien, H.S., A Theory for the Buckling of Thin Shells, Journal of Aeronautical
Sciences, 9, (10), Aug. 1942, 373 384.
7.16 Thompson, J.M.T., Stability of Elastic Structures and Their Loading Devices,
Journal of Mechanical Engineering Science, 3, (2), 1961.
7.17 Thompson, J.M.T. and Hunt, G.W., On the Buckling and Imperfection Sensitivity of
Arches With and Without Prestress, International Journal of Solids and Structures,
19, (5), May 1983, 445 459.
7.18 Kawashima, K. and Ito, T., Snap Through Buckling of Arches in the Elasto-
Plastic Range, Proceedings AIAA/ASME/ASCE/AHS/ASC, 30th Structures, Structural
Dynamics and Material Conference, Mobile, Alabama, April 3 5, 1989, 1355 1363.
7.19 Uniroyal, Inc., Radial Filament Spheres for Deep Submergence Applications, Product
Data Sheet No. DD-109.
7.20 Chini, S.A. and Wolde-Tinsae, M.W., Buckling Test of Prestressed Arches in
Centrifuge, Journal of Engineering Mechanics, 114, (6), 1988, 1063 1075.
7.21 Evan-Iwanowski, R.M. and Mitchell, D.H., Effects of Asymmetries in Loads on
Stability of Shells and Arches, AORD Report No. 2465:12, Aug. 1965.
7.22 Dickie, J.F. and Broughton, P., Shallow Circular Vaults, ASCE Journal of Engi-
neering Mechanics, 99, (EM1), Feb. 1973, 1 12.
7.23 Saitoh, M. and Kuroki, F., Numerical and Experimental Research on Elastic Buck-
ling of the Grid-Arch Structure, World Congress on Shell and Spatial Structures,
Madrid, 1979, 1.103 1.119.
7.24 Lo, H., Bogdanoff, J.L., Goldberg, J.E. and Crawford, R.F., A Buckling Problem of
a Circular Ring, Proceedings 4th National Congress of Applied Mechanics, ASME,
1962, 612 695.
7.25 Chan, H.C. and McMinn, S.J., The Stability of a Uniformly Compressed Ring
Surrounded by a Rigid Circular Surface, International Journal of Mechanical
Sciences, 8, 1966, 433 442.
7.26 Sun, C., Shaw, W.J.D. and Vinogradov, A.M., One-Way Buckling of Circular Rings
Confined within a Rigid Boundary, Journal of Pressure Vessel Technology, Transac-
tions ASME, 117, (7), May 1995, 162 169.
7.27 Sun, C., Shaw, W.J.D. and Vinogradov, A.M., Instability of Constrained Rings: An
Experimental Approach, Experimental Mechanics, 35, (2), June 1995, 97 103.
452 Arches and Rings

7.28 Zagustin, E.A. and Herrman, J., Stability of an Elastic Arch in a Rigid Cavity, Exper-
imental Mechanics, 8, (12), Dec. 1968, 572 576.
7.29 Godden, W.G., The Lateral Inelastic Buckling of Tied Arches, Proceedings, Institu-
tion of Civil Engineers, 4, Part 3, (2), Aug. 1954, 496 514.
7.30 Vlasov, V.Z., Thin-Walled Elastic Beams, 2nd ed., Israel Program for Scientific
Translation, Jerusalem, Israel, 1961.
7.31 Ojalvo, M., Demuts, E. and Tokarz, F.J., Out-of-Plane Buckling of Curved
Members, Journal of the Structural Division, ASCE, 95, (ST10), Oct. 1969,
2305 2316.
7.32 Tokarz, F.J. and Sandhu, R.S., Lateral-Torsional Buckling of Parabolic Arches,
Journal of the Structural Division, ASCE, 98, (ST5), May 1972, 1161 1179.
7.33 Vacharajittiphan, P. and Trahair, N.S., Flexural-Torsional Buckling of Curved
Members, Journal of the Structural Division, ASCE, 101, (ST6), June 1975,
1223 1238.
7.34 Di Tommaso, A. and Viola, E., Ricerche Teorico-Sperimentali Su Problemi Di
Stataica E Stabilità Flesso-Torsionale Di Archi Sottili, Giornale del Gemio Civile,
114, (4 6), 1976, 181 204.
7.35 Wen, R.K. and Lange, J., Curved Beam Element for Arch Buckling Analysis, Journal
of the Structural Division, ASCE, 107, (ST11), Nov. 1981, 2053 2069.
7.36 Yoo, C.H., Flexural-Torsional Stability of Curved Beams, Journal of the Engineering
Mechanics Division, ASCE, 108, (EM6), Dec. 1982, 1351 1369.
7.37 Yoo, C.H. and Pfeiffer, P.A., Elastic Stability of Curved Members, Journal of the
Engineering Mechanics Division, ASCE, 109, (12), Dec. 1983, 2922 2940.
7.38 Yoo, C.H. and Pfeiffer, P.A., Buckling of Curved Beams with In-Plane Deforma-
tions, Journal of Structural Engineering, ASCE, 100, (2), Feb. 1984, 291 300.
7.39 Papangelis, J.P. and Trahair, N.S., Flexural-Torsional Buckling of Arches, Research
Report R492, School of Civil and Mining Engineering, University of Sydney,
Australia, Feb. 1985.
7.40 Papangelis, J.P. and Trahair, N.S., In-Plane Finite Element Analysis of Arches,
Research Report R523, School of Civil and Mining Engineering, University of
Sydney, Australia, April 1986.
7.41 Papangelis, J.P. and Trahair, N.S., Finite Element Analysis of Arch Lateral Buck-
ling, Research Report R524, School of Civil and Mining Engineering, University of
Sydney, Australia, April 1986.
7.42 Papangelis, J.P., Flexural-Torsional Buckling of Arches, Ph.D. thesis, School of Civil
Engineering and Mining, University of Sydney, Australia, 1987.
7.43 Papangelis, J.P. and Trahair, N.S., Buckling of Nonsymmetric Arches Under Point
Loads, Engineering Structures, 10, (4), Oct. 1988.
7.44 Papangelis, J.P. and Trahair, N.S., Flexural-Torsional Buckling Tests on Arches,
Journal of Structural Engineering, ASCE, 113, (7), July 1987, 1433 1443.
7.45 Tokarz, F.J., Experimental Study of Lateral Buckling of Arches, Journal of the Struc-
tural Division, ASCE, 97, (ST2), Feb. 1971, 545 559.
7.46 Ojalvo, M., Tokarz, F.J. and Nontanakorn, D., Out-of-Plane Buckling of Arches,
RILEM, International Symposium, Experimental Analysis of Instability Problems
On Reduced and Full Scale Models, Instituto Nacional De Technologia Industrial,
Buenos Aires, 1971, 535 549.
7.47 Trahair, N.S., Elastic Stability of Continuous Beams, Journal of the Structural Divi-
sion, ASCE, 95, (ST7), July 1969, 1475 1496.
8
Plate Buckling

8.1 Buckling and Postbuckling of Plates


8.1.1 Historical Background

The earliest practical case of plate buckling was the construction of the large span
suspension railway bridges projected by Robert Stephenson in 1845, the Britannia
and Conway Tubular Bridges [8.1] [8.4]. The structure was a long box beam,
through which the train would pass, made of small wrought iron sheets riveted
together to form larger plate structures. Since no theoretical methods were available
at the time to assess the buckling strength, Stephenson asked William Fairbairn
(who had gained extensive experience on malleable iron-plate ship structures) to
carry out tests to determine the strength of the box beam. Small scale models of
various cross-sectional forms were tested in the laboratories of University College
London, with the cooperation of Professor Hodgkinson (who with Fairbairn were
the best known British engineers in the middle of the 19th century, famous for
their experimental approach and results).
In Figure 8.1 the type of failure mode that occurred in some of the specimens is
shown, types of collapse that were only analyzed adequately nearly a century later.
Fairbairn observed: “Some curious and interesting phenomena presented them-
selves in the experiments many of them are anomalous to our preconceived
notions of the strength of materials, and totally different to anything yet exhibited
in any previous research. It has invariably been observed, that in almost every
experiment the tubes gave evidence of weakness in their powers of resistance on
the top side, to the forces tending to crush them.” These were the first experiments
with thin-walled structures which fail through plate or shell instability. Hodgkinson,
examining the results, stated “. . . that any conclusions deduced from received prin-
ciples, with respect to the strength of thin tubes, could only be approximations;
for these tubes usually give way by the top or compressed side becoming wrin-
kled, and unable to offer resistance, long before the parts subjected to tension are
strained to the utmost they would bear. To ascertain how far this defect, which had
not been contemplated in the theory, would affect the truth of computations on the

Buckling Experiments: Experimental Methods in Buckling of Thin-Walled Structures: Basic Concepts, Columns, Beams and
Plates – Volume 1. J. Singer, J. Arbocz and T. Weller Copyright © 1998 John Wiley & Sons, Inc.
454 Plate Buckling

(a)

Deflection.
Diameter.

No. of Length. Weight of Tube.


Sectional Weight Permanent
Experi- Date. Between Between Remarks.
Total. Total. Area. Applied. Sets
ment. Supports. Supports.

Feet. Feet. Inches. Tons. Tons. Inches. Tons. Tons. Inches.


1 July 6, 18.1 17. 12.18 .0455 .0429 1.56 .357 .06
1845 .857 .25 .02 Tube puckered and crushed at top,
1.357 .39 13 inches from the centre, before the
whole of the last weight was laid on.
Thickness of Plate,
.0408 inch,

(b)
Deflection.

Length. Diameter. Weight.


No. of
Permanent
Experi- Date. Sectional Weight Remarks.
Between Between Sets.
ment. Total. Major. Minor. Total. Area. Applied.
Supports. Area.

15 Sept. 18 26.25 24. 21.25 14.125 .1594 .1455 3.823 .379 .04

.772 .09
Barely perceptible.

1.152 .13
Failed at top, before the
1.534 .17 whole of the last weight came
on it. Breaking-weight probably
1.915 .22 3.246 tons.
Thickness of Plate, 2.297 .28
0.688 inch.
2.672 .34

3.050 .42 Minor diameter diminished


before failure by 1.9 inch
3.444 .45

(c)

Figure 8.1 (see p. 455)


Buckling and Postbuckling of Plates 455

strength of the tubes . . .,” and concluded that a series of fundamental tests had to
be made, which were later carried out by him.
Due to the urgency of the design decisions, Fairbairn, however, had to rely on
the results of his scale tests, using then a large, approximately 1/6 scale model. It is
of interest to note that the designers apparently accepted the validity of the results
from these scale tests, and that from the tests eventually a closely reinforced box
beam evolved to prevent different forms of plate buckling (see Figure 8.2).
In spite of this experimental start, however, much of the research effort for plates
in this century, following Bryan’s classical 1891 solution of the initial buckling load
of a square simply supported plate [2.12], has been devoted to theoretical develop-
ments and in recent decades also to numerical studies. Experimental research also
continued, but many tests were of empirical nature, not directly related to analytical
work. This was noted with regret by Walker in his 1984 review of plate buckling
research [8.4], where he also emphasized that the most important developments in
understanding the phenomenon of plate buckling and generating information for
designers have been made when theory and experiment proceeded hand-in-hand.
A statement that applies indeed also, to other structural elements.

8.1.2 Effective Width

As has been shown in the theoretical discussions in Chapter 2, plates have a stable
postbuckling behavior (see for example Figures 8.3, 8.6 or 8.7). This characteristic,
which was first pointed out by Schuman and Back in 1930 [8.6], who observed that
the failing load of a flat panel was materially higher than its buckling load (stating
“For the wide, thin plates, the Bryan load is as low as 1/30 of the maximum
load and in general varies from 1/10 to 1/20 of the maximum”), makes plates
very suitable structural-elements in lightweight structures. The interest of aeronau-
tical engineers in the thirties and forties in lightweight plate structures motivated
extensive theoretical studies as well as experimental investigations (for example
[8.7] [8.15]). Thin plate structures have also been used in marine and civil engi-
neering and hence many of the primary experimental studies on their buckling and
postbuckling behavior originate in these disciplines (see [8.16] [8.22]).
Since thin plates can carry loads considerably in excess of the buckling loads,
their postbuckling behavior and ultimate strength became the focus of experimental
and theoretical investigations. The variation of the stress distribution in the plate,

Figure 8.1 Typical Experiments of Hodgkinson and Fairbairn for the Britannia Tubular Bridge
in 1845 (from [8.2]): (a) “The apparatus used for the preliminary experiments
with the sheet iron tubes and beams. The models varied in length from 15 to 31
ft., and in diameter from 12 to 24 in., bearing at the centre from 1 to 6 tons.”
(b) One of the “Preliminary Experiments on the Transverse Strength of Cylindrical
Tubes of Riveted Boiler Plate,” Exp. 1. (c) One of the “Preliminary Experiments
on the Transverse Strength of Elliptical Tubes of Riveted Boiler-Plate,” Exp. 15.
(d) Experiments on “Crushing of Tubes of Wrought Iron.” “Resistance of Tubes,
Rectangular in Section to a Force of Compression in the Direction of their Length.”
Experiments 8 and 1, short rectangular tubes
456 Plate Buckling

Figure 8.2 Cross-section and longitudinal section of the Britannia Bridge in its ultimate
form 1849 (from [8.1])

as the load significantly exceeded the buckling load was studied and a concept of
“effective width” was conceived to facilitate assessment of the maximum load. The
concept was first employed by ship designers (see Faulkner’s review, [8.19]), then
indicated as the apparent behavior of wide plates by Schuman and Back [8.6] and
then defined and theoretically determined by von Kármán [8.7] and consequently
improved by Sechler, Cox, Maguerre and other aeronautical investigators (see
[8.8] [8.13]) with experiments and theory (see also Hoff’s review, [8.14], and
Sechler and Dunn [8.15]).
The concept is essentially as follows: Up to the buckling load, the stress distri-
bution in a uniformly compressed rectangular plate is uniform. With increase in
load, the central unconstrained portion of the plate will start to deflect laterally
and will therefore not support much additional load, whereas the portions close
Buckling and Postbuckling of Plates 457

Figure 8.3 Load-deflection curves for imperfect plates (from [8.5])

Figure 8.4 The concept of effective width

to the supported edges will be constrained to remain straight and will continue to
carry increasing stresses. The stress distribution is then like c in Figure 8.4 and
the total load carried by the plate of thickness t is
 b/2
Ppl D t c dx. (8.1)
b/2
458 Plate Buckling

In order to simplify design calculations, it is now assumed that the total load is
carried by two fictitious strips of width we (the “effective width”) directly adjacent
to the edges of the plate (see Figure 8.4), which carry a uniform stress st and the
remaining plate is entirely ineffective, such that
Ppl D 2we tst . (8.2)
The “effective width” can be obtained experimentally from measurement of the
stress distribution c in Eq. (8.1) and equating Eq. (8.1) with Eq. (8.2). The shape
of the stress distribution changes with load, but for design purposes the we at
failure is needed. By assuming the plate to be simply supported on the sides and
that the postbuckling deflection is such that horizontal tangents at the inner edges
of the two load-supporting strips are parallel to the original middle plane of the
flat plate, von Kármán [8.7] could disregard the center of the buckled plate and
handle the two strips as if they were together a simply supported plate of width
2we . Thus he derived a theoretical formula for the effective width
 
t E E
2we D p  1/2 D Ct 8.3
3 1 2 st st

and with  D 0.3 C D 1.9.


In view of the rather arbitrary assumptions in the derivation, the coefficient C was
later obtained by extensive experiments [8.8], [8.10] and [8.13]. As the theory
indicated that C is a function of the dimensionless parameter

E
 D t/b . (8.4)
st

C is usually presented as a function of . Failure of the plate will occur when st
reaches the yield stress 0 of the plate material, or if it is part of a stiffened flat
panel, when st reaches the stress at which the stringers (reinforced by the plate)
fail by crippling or overall buckling as a column st.ult . Then Eq. (8.3) becomes

2we D Ct E/0 , or st.ult instead of 0 , (8.5)
and the effective width is then presented as
we /b D C/2 8.6

E
where now  D t/b
0

(see for example [8.11] or [8.13]). Civil engineers then adopted the aeronautical
techniques to steel structures and contributed extensive experimental studies as well
as further analyses (see for example [8.16] [8.18]). In recent decades the marine
engineers added new theoretical and experimental studies (see for example [8.20]
or [8.21]), while civil engineers continued their studies (see for example [8.22]).
Buckling and Postbuckling of Plates 459

The “effective width” has become a universally used design formula for axially
loaded plate structures. In the thirties and forties the aeronautical engineers
perfected it for their thin plates in the elastic range. Then they and their colleagues
in civil and marine engineering turned to the plastic range (see for example [8.23] or
[8.24]). Many new semi-empirical formulae for we were developed (see Faulkner’s
review, [8.19]) and an alternative formulation the maximum average plate stress
m was also introduced for the same purpose. One should note that the two
approaches are essentially similar and
m /0 D 2we /b. (8.7)
One may also note that civil and marine engineers usually plot their effective width
or m /0 , instead of versus , versus the plate slenderness parameter

ˇ D b/t 0 /E D 1/. (8.8)
It may be added that for determination of the strength of a stiffened panel, the
effective width of the plate to be added to the stiffness of the stringer differs
from that used to assess the load carrying capacity of the plate. This “reduced
effective width” we0 is less than we since it allows for the decrease, with increasing
compressive strain, of the load carrying capacity of the supporting buckled plate
(see [8.10], [8.14] or [8.19]). As, however, the method is semi-empirical, this
difference is often not taken into account (see e.g., [8.15] or [2.10], 451 455,
497 500).
The usefulness of the effective width concept obviously depends on the avail-
ability of reliable experimental results for the constants in the semi-empirical design
formula, as was realized by von Kármán and his co-workers and by later investi-
gators. It was soon also recognized that the boundary conditions notably affect the
experimental determination of the effective width.

8.1.3 Postbuckling Behavior and “Secondary Buckling”

As was emphasized in Chapter 2, and can be clearly seen in Figure 8.3, the buck-
ling load of a plate strongly depends on its out-of-plane boundary condition, in the
case of a rectangular plate primarily on those of the unloaded long edges. When the
plate continues to deform in the postbuckling regime, it stretches and membrane
stresses come into action to resist deformation, in addition to the bending stresses
which resisted the initial buckling. The in-plane boundary conditions therefore join
the out-of-plane ones in strongly influencing the postbuckling behavior. Precise
definition and control of both out-of-plane and in-plane boundary conditions has
therefore been the primary concern in all plate buckling and postbuckling exper-
iments, in the latter particularly the possible change of boundary conditions with
increase in deformation. The boundary conditions will therefore feature predomi-
nantly in the discussion of experiments in the next section.
This change in boundary conditions is likely to be partly due to the change in
the buckled form as the plate proceeds deeper into the postbuckling region. One
460 Plate Buckling

can distinguish between minor and major changes in buckled form with increase
in load beyond buckling [8.25]. A minor change is a gradual smooth change like
the flattening of the transverse half-wave, whereas a major change consists of an
abrupt snap from one buckled form to another, like the increase in the number
of waves along the direction of compression in a rectangular plate. The major
changes in postbuckling pattern were already noted by Sechler in the thirties [8.8]
and [8.11], and were emphasized by Stein in the late fifties, when he compared his
perturbation analysis with experiments carried out at NASA Langley and studied
the change in wavelengths with the aid of a simplified mathematical model [8.26]
and [8.27]. Figure 8.5 shows the phenomenon very clearly, which is often called
“secondary buckling” (see for example [8.28] or [8.29]).
Secondary buckling and the accompanying increase in the number of waves
was observed in the last decades in different buckling experiments on rectangular
plates (see for example [8.29] or [8.30]) and was hence studied theoretically and
numerically by a number of investigators in the seventies (for example [8.28]
and [8.31] [8.34]). Uemura and Byon [8.34] also carried out a series of careful
experiments on clamped square plates, whose measured deflection patterns present
details of the secondary buckling behavior. The initial imperfection shapes along
the centerlines of the x and y axes were measured (by sliding dial gages, guided
by a cruciform frame) after the plates were placed in the experimental setup (see
Figure 8.8a, which presents the initial deflections and the deflections under load
of a Duralumin plate D-20-1, 2 mm thick with b/t D 150). Note in the figure, that
the scale of the initial imperfections is magnified five times compared to that of the
deflection under load. The primary buckling mode shown in Figure 8.8a has one
half wave vertically and horizontally. As the load increases, secondary buckling
occurs at P D 2, 420 kg and the load increases, the buckled pattern in the vertical
x direction changes abruptly from the one wave (marked by x) mode to two waves

Figure 8.5 Comparison of non-dimensional load-shortening curves of a plate in the postbuck-


ling region, as given by (elastic) theory and experiment (from [8.27])
Buckling and Postbuckling of Plates 461

Figure 8.6 Comparison of theoretical and experimental load-deflection relations for imperfect
rectangular plates (from [8.5])

(marked by ). Horizontally, obviously the large deflections in the y direction of


the previous pattern (marked by x) change to much smaller ones, but in the opposite
direction. In Figure 8.8b the deflections measured during the unloading process are
shown, starting with the two wave vertical deflection pattern at P D 3000 kg (the
solid line marked by O symbols, which also appears in Figure 8.8a) which snaps
462 Plate Buckling

Figure 8.7 Theoretical load-deflection curves for imperfect square plates, and comparison
of buckling stress ratios obtained by strain-reversal method and top-of-the-knee
method (from [2.15])

through to a deflection with one wave (marked by ) at P D 2150 kg. Note that
the deflection is in the opposite direction to that of the primary buckling pattern in
Figure 8.8a and to the initial imperfection pattern. The accompanying change in the
horizontal deflection is therefore also in the opposite direction to that of the primary
buckling pattern in Figure 8.8a. However, as unloading proceeds, at P D 1560 kg
the deflection pattern snaps again, both vertically and horizontally, and returns to
the original deflected pattern (marked by x) similar to that grown from the initial
imperfections during loading. The measured abrupt change in deflection pattern
demonstrates the secondary buckling phenomenon very clearly.
The major changes in the buckled form were studied in the eighties by Boncif
et al. [8.35] in a series of careful experiments on these brass alloy rectangular
plates with a/b D 9 and b/t D 200. They investigated also the influence on the
postbuckling behavior of the boundary conditions at the loaded short ends of the
plates. The abrupt snapping from one buckled form to another was studied here
by observing the behavior under a decreasing load in the postbuckling regime.
They applied an axial force, well above the critical load, then “forced” a given
wave number by applying temporarily a gentle pressure, perpendicular to the plate,
at locations corresponding to maxima of wave amplitude. Then they decreased the
Buckling and Postbuckling of Plates 463

Figure 8.8 Secondary buckling of a flat plate under uniaxial compression changes of
deflection patterns on the centerlines during loading and unloading (from [8.34]):
(a) during loading process, (b) during unloading process
464 Plate Buckling

axial force slowly, and at a lower value of the force, still above the critical, the
plate snapped to a pattern with a lower wave number. A characteristic “click” due
to the “absorption” of two of the bulges (one mode of the buckled wave form)
was heard.
This type of snap, or jump, occurred with simply supported loaded edges in the
central region of the plate, and was characterized by subtraction (or addition if the
force was increased) of one wave mode. For clamped loaded edges (see [8.36]) the
snap occurred at the ends of the plate and corresponded to a variation of only half
a wave mode (one bulge) at a time. In both series of tests the unloaded boundaries
were knife edges simulating simple supports with no in-plane restraints. It is of
interest to note that the difference in the boundary conditions of the loaded short
edges of rather long plates (aspect ratio 9), which would usually be negligible at
buckling, significantly affected the far postbuckling (secondary buckling) behavior
of these very thin plates b/t D 200.
A recent theoretical study of the mode jumping phenomenon [8.37], extended
the investigation to biaxially compressed plates and also showed the influence of
the loading conditions on the postbuckling behavior.

8.1.4 Influence of Geometric Imperfections

Initial geometric imperfections, or initial out-of-flatness, do not considerably affect


the buckling load and ultimate load of axially compressed plates, on account of their
stable postbuckling behavior. The effect is however most notable at stresses near
the theoretical perfect flat plate critical stress, where the imperfections mask the
buckling load and increase displacements and the resulting maximum membrane
stresses beyond those of perfect plates (see Figures 2.7, 8.3, 8.6 or 8.7). Beyond
buckling and well into the postbuckling region the effect of geometric imperfections
diminishes. As a matter of fact, well above or below the critical stress, plates with
initial out-of-flatness behave very much the same as perfectly flat ones.
This was shown for elastic behavior in a number of extensive theoretical inves-
tigations (like [2.15], [8.38], [8.40] and [8.5]) as well as experimental studies (like
[8.39], [8.40] and [8.41]). The early 1946 NACA study on the effect of small
deviations from flatness on buckling of simply supported square plates [2.15] is of
particular interest since its aim was to obtain simulated test data, and also study
(with this simulation) the suitability of experimental methods for definition of the
critical load, that will be discussed later. It was also shown there that the effective
width is less for a plate with initial geometric imperfections, but also this effect
diminishes when the buckling load has been considerably exceeded.
As the plate proceeds further into the postbuckling region and approaches its
ultimate load, or collapse load, some of the material is no longer elastic, and
plasticity determines the behavior of the plate. The effects of initial geometric
imperfections have therefore to be assessed also in the plastic range. This has been
done theoretically by elasto-plastic large deflection analyses (as for example [8.42],
[8.45] [8.50]) and the results have been correlated with experimental studies
Buckling and Postbuckling of Plates 465

[8.43] [8.46]. The effect of initial geometric imperfections are found to be rela-
tively small also in the plastic range. For example, Figure 8.9 (from [8.42]) shows
the predicted influence for square plates in uniaxial compression, the plate (a) with
the smaller ˇ D 1.037 being the thicker and narrower one, for which buckling
occurs beyond the elastic range. For typical initial imperfection amplitudes of
about υ0 D 0.2t, reductions in strength of about 5 percent are predicted. For plate
(b), with ˇ D 2.074, where buckling is entirely elastic, the effect is smaller and
notable only near the critical stress, as pointed out earlier. Figure 8.10 (from [8.45])
shows the predicted strength for long steel plates with initial imperfection magni-
tudes of υ0 D 0.0001b and υ0 D 0.001b. (Many investigators prefer to relate the
displacements and imperfections to the width of the plate instead of the usual thick-
ness, with an average b/t D 50 the amplitudes here are of the order of υ0 ³ 0.005t
and 0.05t respectively, typical for the relatively thick plates, 1/4 inch nominally).
Slightly larger effects, of the order of 10 percent are predicted here for 1.5 < ˇ < 2.
The good correlation between theory and experiment can be seen in Figure 8.11
(also from [8.45]), where the υ0 D 0.0001b imperfection predicted strength curve
fits the test results of the as rolled (R) plates well, though here the imperfection
magnitude was assumed to be representative of the tests, and is smaller than that
measured in practice, for example in bridge plating where the magnitude is about
0.001b to 0.007b (see [8.51]). More recent experiments (which will be discussed
later on account of their sophisticated test setup) deliberately introduced controlled
out-of-flatness in the center of the plate [8.43]. By holding down the plate edges
in a subsidiary rig while jacking up the center, and then releasing the jack to leave
a predetermined permanent set, a single dent was produced at the center of the
plate, and this central sub-panel was then expected to fail. Indeed these delib-
erate imperfections grew noticeably with load and usually initiated failure. Two
magnitudes of initial imperfection, υ0 D 0.001b and υ0 D 0.005b (for an average
b/t ³ 50 this corresponds to υ0 ³ 0.05t and 0.25t respectively) were induced. In
Figure 8.12 (from [8.43]) experimental results are compared with results of five
numerical elasto-plastic analyses (those of [8.42], [8.47] [8.50]). The curves and
test points for the unwelded plates, show indeed at worst a 10 percent effect when
the magnitude of initial geometric imperfection is increased five times. In the test
results the effect is actually obscured by experimental scatter. On the other hand,
the effect of residual stresses due to welding, which will now be discussed, appears
to be significant.

8.1.5 Influence of Residual Stresses

Residual stresses in plates are primarily welding stresses and hence their effects
on plate buckling and plate strength have been extensively studied by civil and
marine engineers, whose structures usually incorporate welded plating (see for
example [8.19], [8.21] or [8.42], [8.44], [8.45]). Welding introduces not only
residual stresses, but also causes initial distortions, which for example in ship
structures may be an equally significant component of the imperfection of the plate.
466
Plate Buckling

Figure 8.9 Effect of initial geometric imperfection (out-of-plane deformation) on the strength of square plates under uniaxial compression (from
[8.42]): (a) inelastic buckling, (b) elastic buckling
Buckling and Postbuckling of Plates 467

Figure 8.10 Theoretical strength curves for long steel plates with varying magnitudes of initial
imperfections (out of flatness) υ0 (from [8.45])

Figure 8.11 Comparison between test results and theoretical predictions for simply supported
steel plates, R D as rolled and W D welded (from [8.45)]

The influence of residual stresses on the buckling stress of long plates has been
shown by many experiments in the USA, Japan and England to be fairly well
represented by the simple reduction in the critical stress by cr D r where r
is the compressive residual stress in the plate (which balances the narrow regions
of tensile residual stresses near the stiffeners, that resulted from the weld heat).
Hence, with residual stress the critical stress becomes

cr r D cr  r (8.9)

provided cr  0.20 (see [8.19]); for cr > 0.20 the reduction in cr due
to residual stress becomes progressively less than that predicted by Eq. (8.9),
(see [8.45]).
The extent to which the residual stress reduces the ultimate load capacity of the
plate is, however, more important. The simple reduction formula, Eq. (8.9), applies
also to the reduction in plate strength within the elastic range, which corresponds
approximately to a plate slenderness ˇ > 2, and hence with low residual stress the
468 Plate Buckling

Figure 8.12 Comparison between test results and five numerical elasto-plastic analyses, for
simply supported steel plates, with two magnitudes of initial imperfections (out-
of-flatness), (from [8.43]): (a) υ0 /b D 0.001, (b) υ0 /b D 0.005

maximum average plate stress becomes

m r D m C cr . (8.10)

For smaller values of plate slenderness ˇ, inelastic effects dominate and the reduc-
tion in plate strength due to welding can be represented by a strength reduction
ratio Rr (see [8.19]), which is a function of ˇ and the tangent modulus ratio Et /E.
Buckling and Postbuckling of Plates

Figure 8.13 Load-shortening curves of welded and unwelded steel plates under axial compression (from [8.45])
469
470 Plate Buckling

The maximum plate stress is then


 
 Rr 0 for 0  ˇ < 1 

m r D 2 1 . (8.11)
 Rr 0  2 for ˇ ½ 1 
ˇ ˇ
Agreement between Eq. (8.11) and test data was shown in [8.19] to be fairly good.
Typical percentage values for 1  Rr  are 10 20 percent for ˇ > 1.5, and less than
5 percent for ˇ < 1. Some tests on welded steel plates (for example [8.45] or [8.46])
have shown larger reductions in strength due to weld-induced stress, occasionally
even above 25 percent. However, a fairly recent statistical assessment of 413 single
plate tests and 224 box column tests in Europe, North America and Japan [8.22]
determined the bounds of m as 5 15 percent over the entire practical range of
the slenderness parameter ˇ.
Load-shortening curves of welded and unwelded plates, like those of Dwight and
Ractliffe’s steel specimen in Figure 8.13 (from [8.45]), show the occasional large
reduction in strength due to welding, for example, for b/t D 54 in Figure 8.13,
m D 0.34m . It can also be seen that the curves for the welded plates in
Figure 8.13 are less peaky, the fall-off in load is less drastic for the welded plates
than for the “as rolled” (unwelded) ones, making the failure “softer.”
The combined effects of residual stress and initial geometric imperfections have
also been assessed by the elasto-plastic large deflection analyses (like [8.42],
[8.47] [8.50], [8.52]) and have been confirmed by experiments, see for example
Figure 8.12 (from [8.43] or [8.52]).
For the experimenter, the stable postbuckling behavior, as shown in Figures 8.3,
8.6 or 8.7, indicates the inherent difficulty in defining the buckling load for plates
with imperfections. This difficulty grows with the initial geometric imperfections
of the plate. It is often argued that the practical significance of the buckling load
becomes doubtful in very thin plates which carry, sometimes even elastically, loads
far in excess of the critical loads. The buckling load, however, though a theoretical
value in an imperfect plate, remains an important criterion defining the behavior of
the plate and assessing the actual boundary conditions in the test. Hence different
methods have been developed to define the critical load from the test results,
which will be discussed in the next section. The many methods in use indicate that
probably none of them is very precise in defining the critical load or stress.

8.2 Experiments on Axially Compressed Plates

8.2.1 The US Bureau of Standards Test Setup

A typical test setup for buckling and strength of rectangular flat plates under
uniaxial compression is that developed in 1930 by Schuman and Back at the US
Bureau of Standards [8.6]. These tests have become a classic on account of their
results, pointing out that thin plates can carry loads much above the buckling
load. Also the test setup, Figure 8.14 (from [8.6]), has become a classic, which is
Experiments on Axially Compressed Plates 471

Figure 8.14 US Bureau of Standards 1930 test setup for rectangular plates under uniaxial
compression (from [8.6])

still used today in some tests and appears in textbooks (for example [2.1]), and
the careful experimental procedure and evaluation of results justify a closer look
beyond the historical interest.
The test fixture of Figure 8.14 was designed and built after several forms of test
rig had been tried. It consists essentially of a base plate with two heavy vertical
channels, to which grooved supporting bars are attached. These bars have 45°
V-grooves, into which the test plate is set, its lower end resting on the base plate. A
horizontal bar, bolted to the top of the channels, limits their spreading at the top. By
472 Plate Buckling

Figure 8.15 Rotation of plates edges in V-grooves during buckling (from [8.6])

means of screws the grooved bars are adjusted so that the specimen can rotate about
its edges and slide vertically in the grooves. This was supposed to simulate simple
supports along the vertical unloaded edges. Schuman and Back realized, however,
that simple supports were not precisely achieved. For, as can be seen in Figure 8.15
(from [8.6]), the edges of the plate are not entirely free to rotate in the V-groove,
and any rotation during buckling is accompanied by an out-of-plane movement,
and therefore the edge support perpendicular to the plate is actually weaker than
presumed. They also noticed that some plates snapped out of the grooves near the
top, beyond buckling, and suspected insufficient rigidity of the supporting channel
and of the horizontal bar, that had to prevent spreading. They suggested as a partial
remedy the equalizing of the pressure on the two screws holding the V-grooves
against the plate, but the main problem is that shown in Figure 8.15, which is
inherent to the V-groove concept. Later investigators sometimes made the edges
semi-circular to reduce this effect and, in case of loaded edges, to ensure application
of the load in the middle plane of the plate. In the Schuman and Back test fixture
the loaded edges have no V-grooves but are “flat edges”, which implies partial
rotational restraint, and since their primary interest was in long plates, they were
probably relying on the relative weaker influence of the boundary conditions of
the shorter loaded edges in long plates, discussed in Chapter 2. Other investigators
at the time, like for example Sechler [8.8] used V-grooves for both unloaded and
loaded edges, with means for position adjustments to ensure accurate alignment
of the specimen, assuming that fairly close simulation of simple supports on all
edges is achieved.
It may be mentioned, by the way, that on account of their simplicity, V-grooves
are still used today for experiments on initial buckling of plates before significant
edge rotations occur, but for tests well into the postbuckling region they represent
incompletely defined boundary conditions, affected also by the changes in buckling
pattern that may occur in the deep postbuckling region. Furthermore, the friction
in the V-grooves may carry some indeterminate amount of the load applied to the
plate (load diffusion). This should be minimized by suitably lubricating the groove,
and measured to ascertain its magnitude.
To permit unrestrained loading, the specimen in the test fixture of Figure 8.14
extends 1/8 in. beyond the ends of the grooves. The load is applied at the middle
of the upper edge, through a 1 in. ð 4 in. bar, which can rotate about an axis
perpendicular to the plane of the plate at this location, in order to ensure a fairly
Experiments on Axially Compressed Plates 473

uniform distribution of load until buckling initiates. Then a redistribution of the


load occurs, since the vertical central portion, after buckling, exerts less force
on the loading bar and more load is carried by the side portions, as was earlier
explained in relation to the effective width concept. If the non-rigid loading bar
deflects under load, it counteracts this effect, tends to give a more uniform load
distribution and therefore somewhat “overloads” the central portion of the plate,
causing earlier plate failure. Hence Schuman and Back not only indicated the role
of the non-uniform stress distribution, but pointed out the possible distortion of
results that may be caused by insufficient rigidity of the supporting structure.
They also took great care to map out in detail the buckling patterns and their
growth at increasing load steps, with the aid of a dial micrometer that could be
moved in steps of 1 in. both vertically and horizontally, a scheme that later devel-
oped into modern displacement scans.

8.2.2 Needle and Roller Bearings and Knife Edges for Simple
Supports

The importance of the boundary conditions, however, motivated many efforts


to develop better simulation of simple supports. Two decades later, a significant
improvement in test boundary conditions was made by Hoff, Boley and Coan
[8.53] in a series of very careful tests on rectangular fiberglass plates (of aspect
ratios 1 to 2), in which the simple support on the loaded edges was provided
by an arrangement of split needle bearings. Essentially, the aim was to extend to
the loaded ends of plates the concept of roller bearings, or roller bearing blocks,
used for pin ends in columns (see Figures 6.6e and g in Chapter 6). The bearing
assemblies and the small bearings, which had a quadrant cut from their shells to
permit the insertion of the slotted rods which supported the test specimen, were
a fairly complicated set-up but ensured a close approximation to a continuous
simple support (see Figure 8.45). This approach has since been employed by many
investigators who aimed at precise simple supports. For example [8.41] and [8.40]
for loaded edges of rectangular plates of aspect ratio 2 and [8.54] for all edges of
square plates.
A typical example is Walker’s 1967 test rig for his experimental studies on the
strength of rectangular plates under eccentric in-plane loads [8.41]. Figure 8.16
shows the components of the plate testing rig, which is a good example of a small
laboratory size (for plates 10 in. ð 20 in.) setup, and will later be described in
more detail. The loaded edge at a corner of the plate is shown in Figure 8.17a,
the split needle bearings in Figure 8.17b and the details of unloaded edges in
Figure 8.18. The simple-support boundary conditions at the loaded edges were
obtained by mounting the plate in slotted rollers (see Figure 8.17b) which fitted
into needle bearings, which were mounted in groups of four in bearing holders
that fitted in the loading beams (see Figure 8.17a and part 8 in Figure 8.16). A
segment of 40° was cut from the bearings, and after removing two further needles
the ends of the bearings were resealed. To maintain loading right to the edge of the
plate, grooved blocks (make up pieces) were fitted in the corners in place of the
474
Plate Buckling

Figure 8.16 Walker’s 1967 plate testing rig (exploded view, from [8.41])
Experiments on Axially Compressed Plates 475

(b)

Figure 8.17 Simple support boundary conditions at loaded edges (from [8.41]): (a) needle
bearings and slotted roller, (b) loaded edges at the extremity of the plate

Figure 8.18 Details of unloaded-edge supports in Walker’s 1967 plate testing rig (from [8.41])
476 Plate Buckling

bearing and rollers (see Figure 8.17a). On the unloaded edges, rounded knife edges
simulated simple supports (Figure 8.18a) and restraining bars simulated clamped
edges (Figure 8.18b). The knife edges or restraining bars were bolted to the main
runners by 3/8 in. diameter bolts (see part 7 in Figure 8.16). With these bolts the
two types of edge supports could be positioned as desired. However, the conflict
of either insufficient out-of-plane support, or friction caused restrictions to in-plane
movements, which should be unrestricted, remains with this kind of supports.
The in-plane restraint caused by knife edges, or restraining bars, of the type
shown in Figure 8.18, has troubled experimentalists for decades, both on account
of the load dissipation or partial load carrying, and on account of in-plane restraint
perpendicular to the load direction. Hoff, Boley and Coan [8.53] also used rounded
knife edges (radius 1/64 in.) for their unloaded edges, but performed compar-
ison tests with the then frequently used slotted tube type side supports. The knife
edges were finally preferred, since they were more rigid and could be more conve-
niently aligned, but though load dissipation was assessed to be insignificant, doubts
remained and appeared in the conclusions of [8.53]. A decade later Yamaki [8.39]
used knife edges for both loaded and unloaded edges of his square plates, pairs of
single knife edges for simple supports and pairs of double knife edges for clamped
boundary conditions (Figure 8.43c). However, to diminish the effects of friction,
the edges of the plates were coated with Vaseline before mounting, a simple but
effective remedy. (Today experimenters use Teflon tape for the same purpose).
Comparing the results of his carefully executed experiments with his theoretical
predictions [8.38], Yamaki concluded that pairs of knife edges are satisfactory
for unloaded edges, both for simple supports and for clamped edge conditions.
For the simply supported loaded edges, however, the knife edges proved to be
unsatisfactory, since appreciable bending strains appear near the edge.
To ensure precise definition of the boundary conditions, Schlack [8.54] then
employed a series of independently acting needle bearing blocks (Figure 8.19) for
all the edges of his square plates. Each bearing block assembly, shown in the
figure, consists of two needle bearing assemblies and an edge support cylindrical
slotted insert. The 1 3/8 in. long bearing blocks are individually bolted to the rigid
frame and are placed 1/8 in. apart, to reduce load diffusion and permit alignment
and adjustments, carried out by monitoring the strain gages during preliminary
partial loading. His test fixture was designed to restrain perpendicular in-plane
displacements at the vertical edges of the plate and apply uniform displacements
at the loaded edges (the straight edge case). In order to “smooth out” the loading,
a thin strip of lead was placed along each plate edge, between the edge of the plate
and the grooves cut in the support inserts.
Later, Rhodes, Harvey and Fok [8.40] returned to the earlier pattern (of [8.53]
and [8.41]) of roller bearings for loaded edges and rounded knife edges at the
unloaded ones, to simulate simple supports on all sides of the plate.
A few years later, a similar pattern of roller bearings for loaded edges, and a
simpler scheme for the unloaded edges, was employed by Fischer and Harre for
their extensive larger scale experiments on short plates [8.55]. The purpose of this
investigation was to provide buckling strength test data for uniaxially compressed
Experiments on Axially Compressed Plates 477

Figure 8.19 Independently acting needle bearing blocks to ensure simple support boundary
conditions, employed by Schlack for his perforated square plates (from [8.54])

thin plates with aspect ratio ˛ D a/b  1, the tests covering ˛ D 1, 3/4, 1/2,
1/4 and 1/6. The experimental setup aimed at simulation of simple supports on
all four edges. A schematic cross-section of the setup (Figure 8.20a) shows the
approximately semicircular needle bearings at the loaded edges, which consist
each of 80 segmented circular plates, of 25 mm thick alloy steel, which sit on
80 precision bearing segments 22 mm thick (as can be seen in the details of
Figure 8.20b). The test plates fit into the central slots of these segmented circular
plates such that their edges are the center of rotation. Plates up to 2 meter length can
be tested in the test rig, which required very accurate machining of the 2 meter
long upper and lower bearing blocks (see Figure 8.20c), including finishing by
honing. The multi-segment loaded-edge fittings permitted practically continuously
varying rotation of the edges, with assured central load application, representing a
very close simulation of theoretical simple supports. Strain measurements during
the first tests indicated that the friction between the plate edges and the slots in
the segmented circular plates prevented free in-plane displacements. A 1 mm thick
Teflon strip was therefore placed in the slot, and the plate edges were covered with
a 0.15 mm thick Teflon film before being fitted into the slots.
The short vertical unloaded edges move in guide blocks (as can be seen in
cross-section BB in Figure 8.20c), which have T-slots into which the plate edges
478 Plate Buckling

Figure 8.20 Simple support boundary conditions in the larger scale Fischer and Harre exper-
iments on short plates (from [8.55]): (a) schematic cross section of the setup,
(b) details of needle bearings at the loaded edges, (c) guide blocks for unloaded
edges

protrude 3 mm and where they are supported on rounded edges permitting free
rotation (see detail A in Figure 8.20b). As can be seen in the figure, the plate
edges are also here covered with Teflon film to prevent friction, and 3 mm long
saw-cuts were also cut into the specimens, perpendicular to their unloaded edges
at intervals of a/5, the width b being measured from the ends of these saw-cuts.
Hence load diffusion was minimized. The extensive use of Teflon film and tape
Experiments on Axially Compressed Plates 479

in this set-up to eliminate friction should be noted, as well as the relatively stiff
loading blocks (see Figure 8.20a), aiming at constant edge displacement over the
entire width b of the loaded edges.
Initial geometrical imperfections, of one half wave in each direction, were intro-
duced into the test specimens, supported along all edges, with rubber mallet blows
in the central region of the plate, aiming at an amplitude of υ0 D 0.00004a2 /t.
This υ0 is similar to that of 0.001b representing the lower end of the measurements
on bridge box girders in [8.51], and selected as a representative out-of-flatness in
[8.43], since two-thirds of the bridge plating surveyed (in [8.60]) was found to have
an out-of-flatness below 0.001b, if say a/b D 0.5 and b/t D 100, which is in
the middle range of the specimens of [8.55]. These geometrical imperfections were
measured and marked as topographic curves on the specimen. They approximated
the desired curves, but probably changed somewhat as the plates were fitted into
the test-rig.
The deflections of the plates under load were measured both mechanically, by
dial gages with 1/100 mm divisions at a number of locations, and electrically by an
inductive displacement probe in the center of the plates, that provided a continuous
record on an X-Y plotter. The strains were also recorded by strain gages at many
locations.

8.2.3 The ETH Zürich and US Navy DTMB Plate Buckling Tests

Another technique for simple supports was developed at the Institut für Baustatik
of the ETH Zürich in the forties for an extensive series of buckling tests on 359
axially compressed aluminum alloy plates ([8.16]). These tests were to provide
design information for civil engineering structures usually constructed of mild
steel. But Kollbrunner, chose instead for all the specimens an aluminum alloy,
Avional M, which on account of its smaller modulus of elasticity (about 1/3 of
that of steel) would exhibit three times greater strain and buckling deformation,
that would permit a more accurate study of the buckling process, a rather advanced
“scientific” approach at the time. Also of interest is the ETH 7 ton levered universal
test machine (built in the thirties), a “dead load” machine, in which the compressive
load P can immediately follow the buckling of the plate, as happens in the buckling
of a loaded structure in practice. A decade later an improved larger 20 ton levered
compression test machine, based on the same principle, was built at ETH for
additional plate buckling tests [8.18]. Figure 8.23 shows the larger levered test
machine, which represents an excellent example of a precise civil engineering
buckling test rig of the fifties.
To return to the edge conditions, the unloaded edges of Kollbrunner’s plates
[8.16] were supported with accurately machined circular cylindrical steel riders,
that were fixed to them with set screws (see Figure 8.22). For simple supports
(Figure 8.22a), the cylindrical riders can rotate about the effective edge of the
plate (which is the axis of the rider cylinder), between two rigid steel guide beams
(see also Figure 8.21 which shows the plate and its riders assembled). Friction
may still provide some restraint here to rotation and in-plane movement, however
480 Plate Buckling

Figure 8.21 Plate supports in the loading frame of Kollbrunner’s ETH plate buckling tests
(from [8.16])

Figure 8.22 The circular cylindrical steel riders for the unloaded edges employed by Koll-
brunner (from [8.16]): (a) set screws arranged for simple supports, (b) set screws
arranged for clamped edges

less than in the case of the knife edges. To facilitate continuous curving of the
plate during buckling, prevent any stiffening of the plate and to minimize load
diffusion, the riders are short and separated by a small gap (see Figure 8.21). To
prevent seizing of the riders in the vertical guide beams, and assure their relatively
free sliding, the guide beams are tapped lightly during loading and unloading. The
guide beams were also checked, for possible picking up of some of the compressive
load, by tensometers (today electrical strain gages would serve this purpose), and
no load diffusion was observed in the tests. For clamped edges, the rotation of the
Experiments on Axially Compressed Plates 481

riders is prevented by additional positioning screws (Figure 8.22b). At the loaded


edges, the top and bottom compression beams apply uniform compression, with
their shallow V-grooves (see Figure 8.21) assuring load application at the middle
surface of the plate.
In the second series of the ETH plate buckling tests [8.18], the improved test
machine (Figure 8.23) was used, and improved edge supports were devised to
prevent any plate stiffening and load diffusion. Figure 8.24 shows two vertical
cross-sections and Figure 8.25 a horizontal cross-section of the buckling frame in
the test machine of Figure 8.22 (in the configuration for non-uniform compression,
as well as that for uniform compression) and shows the extensive use of ball bear-
ings to eliminate unwanted edge restraints. The left hand side of Figure 8.22 shows
a clamped unloaded end support, and the right hand side a corresponding simple
support. Figure 8.26 shows the details of this simple support. It consists of short
circular cylinders (of about 2 cm length) fitted onto the edge of the plate, with gaps
between them which rotate between pairs of short plates (also about 2 cm long),
which house the ball bearings, on which these plates glide between the rigid vertical
guide beams. Thus continuous curving of the buckling plate is facilitated and in-
plane restraints have been removed. For clamped ends, there are no cylinders and
instead the short plates have pairs of cylindrical rims which clamp the plate edges.
The plates are fastened between the rolling plates with wedges at each location (see
Figure 8.25). To further minimize any stiffening of the unloaded edges, slots of 1
to 1.5 mm width are sawed along the edges of the plate every 20 mm, with deeper
slots (27.5 mm) for the clamped edges and shallower ones (3.5 mm) for simple
supports. At the loaded edges, no attempt was made to attain simple supports and

Figure 8.23 Improved ETH test machine for plates subject to uniaxial compression: schematic
view of lever arrangement (from [8.18])
482 Plate Buckling

Figure 8.24 Improved ETH plate testing machine: buckling frame vertical cross-section and
view (from [8.18])

Figure 8.25 Improved ETH plate testing machine: buckling frame horizontal cross-section
(from [8.18])

“flat edges” are employed. However, in order to obtain continuous loading the load
application is through a sectioned load system, see Figure 8.25 (which shows the
configuration for triangular loading). Horizontal displacement of the loaded edge,
to permit Poisson contraction, is facilitated by a system of ball bearings permit-
ting sliding of the sectional load blocks, and the rollers between their rows. The
resulting test setup has well-defined boundary and loading conditions, and indeed
yielded reliable results, but is somewhat complicated.
Experiments on Axially Compressed Plates 483

Figure 8.26 Improved ETH plate testing machine: details of simple supported edges
(from [8.18])

Figure 8.27 US Navy DTMB plate buckling test fixture (from [8.20]): (a) the components
disassembled, (b) test fixture with test plate in position

At the David Taylor Model Basin of the US Navy, Duffy, Allnutt, Conley and
Becker tried, a decade later, to simulate simple supports in a simpler manner
([8.56] and [8.20]). The plates, 1/8 to 3/8 in. thick and made of various steels and
aluminum alloys, had all four edges machined round so as to be free to rotate
in the test jig (see Figures 8.27 and 8.28). After having been instrumented with
strain gages, on both sides of the plate to indicate the buckling loads, the plate
was placed in the radius insert pieces, which were in the full length milled slots in
the telescoping side columns and top and bottom supports. Note that the edges of
the plates were rounded to match the radius of the insert pieces (which required
a new set of inserts for each plate thickness) and that the surface of the insert
484 Plate Buckling

Figure 8.28 US Navy DTMB plate buckling test fixture: details of simple supported edges
(from [8.56])

pieces were lubricated with molybdenum disulfide. Thus rotational restraint was
minimized (though some small rotational restraint still remained, as indicated by
the small bending strains measured close to the edges of the plate), while out of
plane deflection was prevented, approximating the requirements of simple supports.
In-plane displacements were, however, not allowed for, which must have affected
the postbuckling behavior.
Load diffusion was not considered by these investigators, but was apparently
present, as indicated by the 10 percent lower membrane strains at the bottom of
the plate compared with those measured at the top. It was probably an important
cause of the higher experimental buckling loads.

8.2.4 The Cambridge University “Finger” Supports


The investigators at Cambridge University in the last decades have been worried
about the load diffusion and in-plane displacements of the unloaded edges in
their plate experiments. They therefore developed a system of discrete “finger”
Experiments on Axially Compressed Plates 485

supports, stiff against out-of-plane displacement of the plate, while flexible to in-
plane displacements, both in the direction of load and perpendicular to it (see for
example Figures 8.30 8.32).
The idea was first applied in the form of a system of stirrups by Cox at the
National Physical Laboratory in 1933 [8.9]. His stirrups were 1 in. wide duralu-
minum strips, 1/16 in. thick and 2 in. long, except for the attachment flanges to
which the plate was attached by four bolts each. They were flexible in the direction
of the applied load (by bending about their axis of minimum moment of inertia),
but stiff to out-of-plane rotation of the edge and to in-plane displacements perpen-
dicular to the load. Thus they simulated clamped edges along the unloaded edges
with lateral restraint, as are obtained in practice in a panel riveted to stiffeners (as,
for example, ribs in an aircraft wing). This eliminated most of the load diffusion,
except that required to bend the stirrups, which was computed and corrected for.
The loaded edges were clamped between pairs of angle plates, assumed to represent
fully clamped boundary conditions.
In the late sixties Dwight and Ractliffe [8.45] devised a system of “fingers”
which allow simple supported or clamped unloaded edges that are free to undergo
in-plane movement (to “pull in”), and do not hinder the shortening of the plate
under load. By making 2 1/2 in.-deep saw-cuts, at 1 3/4 in. intervals, 26 “ears” were
provided at the unloaded edges of the specimen (see Figure 8.29). An aluminum
bracket, cut from slotted tee extrusion was bolted to each ear (Figure 8.31). This
engaged with a loosely fitting triangular restraint arm made from 0.1 in. steel sheet
(Figures 8.30 8.32), which could be secured to the bracket by two pins A and B,
and to the rigid test frame by a third pin C. Pins A and C were exactly level with
the ends of the saw cuts (see Figure 8.30). In a simply supported test, pins A and
C only were used, thus making the specimen free to rotate about a line through
the ends of the saw-cuts, but at the same time preventing out-of-plane deflection
along this line. In a clamped test all three pins were inserted, thus producing full
rotational restraint. The loose fitting at pin A assured free rotation and at C freedom
to “pull-in”. Aluminum angles were secured to the other side of the ears by the
same bolts, to stiffen the ears (see Figure 8.32). The width b of the specimen
was measured to a line through the ends of the saw cuts which is the effective
edge (see Figure 8.29). Though it was assumed that these ears would only take
an insignificant part of the applied load, tests were carried out on 1/4 scale eared
models, loaded in tension, which showed that the ears carried an additional load
equivalent to increasing b by 2 5 percent of the length of the saw-cut (0.05 0.13
in. for the full scale test plates of width 8 20 in.). A small correction was made
accordingly. The 0.1 in. thick restraint arms are very flexible in the load direction
and therefore load diffusion is negligible.
This system satisfied the requirements of precisely defined unloaded boundary
conditions, but required considerable preparation of the edges of the plate. A later
system of fingers, developed by Moxham at Cambridge [8.57] could take plates
of uniform thickness without edge preparation, and more recently an improved
system built at Cambridge by Bradfield [8.43] and [8.44], also allowed plates of
varying thickness to be tested without edge preparation.
486 Plate Buckling

Figure 8.29 Cambridge University vertical test rig with plate specimen in position (stiffening
columns not shown), exhibiting saw cuts at the unloaded edges (from [8.45])

The more recent test rig, shown in Figure 8.33, is a horizontal setup of 100 ton
(1 MN) capacity, designed to be very stiff in relation to the specimen. A massive
motor driven wedge jack straining device (No. 1 in Figure 8.33) is therefore used,
as well as large cross-section tie bars (No. 2 in Figure 8.33). The support “fingers”,
shown in Figure 8.34a, are arranged in sets (one set is also shown in Figure 8.33).
Use of one roller and pin to grip the plate allows the plate to rotate through
moderate angles providing simple support. Use of two rollers and pins prevents
relative rotation, enforcing clamping against rotation. These “fingers” are hinged
to a stiff support beam (No. 3 in Figure 8.33) and their rotation about it permits
in-plane displacement perpendicular to the unloaded edge. The axial stiffness of
the finger restrains out-of-plane movements, whereas in the direction of the applied
load, the finger bends about its minor axis where its stiffness is very low, and there-
fore load diffusion is negligible. The fingers shown in Figure 8.34a are designed
to grip specimens of 4 to 6 mm thickness. One may note that their adjustment
Experiments on Axially Compressed Plates 487

Figure 8.30 Cambridge University vertical plate test rig: attachment of the triangular restraint
arms, the “fingers”, and position of the three aluminum stiffening columns (from
[8.45])

Figure 8.31 Cambridge University vertical plate test rig: details of simple support and clamped
arrangements of “fingers” (from [8.45])
488 Plate Buckling

Figure 8.32 Cambridge University vertical plate test rig: a triangular restraint arm “finger”
(from [8.45])

mechanism lies practically below the plate, leaving only small projections above,
which facilitates other tests.
It was found, however, during the tests that when large out-of-plane displace-
ments occurred near the maximum load, the accompanying larger slopes caused the
plate to force the rollers of the fingers apart, which damaged their ends. Bradfield
indicated in [8.44], that extension of the fingers by an extra 20 mm above the slot,
which takes the plate, would probably have prevented the damage. Hence in a later
test series, carried out in the same test rig ([8.58] and [8.59]), new fingers were
made to a stronger and simpler design, embodying this extension (Figure 8.34b).
These later fingers (with essentially the same stiffness about their minor axis) were
tested for their stiffness in the direction of the load, which yielded a possible load
diffusion of at most 1 percent at maximum load.
At the loaded ends, the plates were butted against the hard steel main T-piece
of the bearing platen (see Figure 8.35) and held by clamps which provided some
rotational restraint. Previous plate tests at Cambridge University had shown that
butting a machined plate end onto a machined platen face is not enough to prevent
failure near the ends, due to load irregularities which cannot be precisely predicted
and assessed, instead of near the center of the specimens. Hence some clamping
was required to slightly stiffen the plate near the loaded ends in order to assume
failure near the center. However, on account of the aspect ratio a/b D 4 of
the specimen, even complete clamping would have produced at most a 5 percent
increase in the buckling load and less in the ultimate strength. Since the clamping
Experiments on Axially Compressed Plates

Figure 8.33 Cambridge University large horizontal test setup for axially compressed plates with improved system of “fingers” (from [8.43])
489
490 Plate Buckling

Figure 8.34 Cambridge University large horizontal plate test setup: details of “fingers” (from
[8.59]): (a) original “finger”, (b) improved “finger”

Figure 8.35 Cambridge University large horizontal plate test setup: loaded edge section
through end clamps (from [8.59])

is incomplete, as can be seen by the typical buckling profile (Figure 6 of [8.59]),


the effect is significantly smaller.
Initial geometrical imperfections (out-of-flatness) was deliberately introduced
into the plates tested. This was achieved by holding down the plates edges in a
subsidiary rig, while jacking up the center and then releasing the jack to leave
the desired permanent set. This produced a single dimple in the center of the
plate, which would then bias initiation of buckling and failure in the central sub-
panel of the plate. Such a single dent could be produced to a desired amplitude and
Experiments on Axially Compressed Plates 491

Figure 8.36 Cambridge University large horizontal plate test setup: buckled aluminum plate
in test rig (from [8.58])

presented therefore a controllable initial geometric imperfection (or out-of-flatness).


Two values of amplitude υ0 D 0.001b and υ0 D 0.005b were chosen here for the
controlled initial imperfection, the lower value representing typical measurements
on bridge plating (see [8.51] or [8.60]) and the higher one an upper bound, proposed
for a new bridge code.

8.2.5 Examples of Simple and Clamped Supports

Recently another ingenious, but relatively simple, solution to the problem of simply
supported unloaded boundary conditions has been presented by Minguez [8.61].
In order to obtain along the unloaded (longitudinal) edges boundary conditions
which approach the idealized simply supported ones as closely as possible, he
supported them by two series of tensioned steel wires (see Figures 8.37 and 8.38).
Slots were machined in the unloaded edges, with a spacing d designed to ensure
a large number of support points per buckling half wave (here nine points), and
steel wires were attached, each using a brass collar with a set screw and a piece
of steel angle section. These high-strength piano wires were attached to the test
fixture frame with screws, that could tension the wires as required. Thus out-of-
plane displacement was restricted, whereas the wires allowed rotation and in-plane
displacements.
The loaded (transverse) edges were also intended to be simply supported. After
initially trying flat edges and then V slots in the loading platens, a set of 1/2 in.
492 Plate Buckling

Figure 8.37 Minguez plate buckling test fixture for simulation of simple supports, with typical
aluminum specimen in position, showing support wires at the bottom and rollers
on the left (from [8.61])

diameter steel rollers in a circular slot having the same radius (Figure 8.39) was
employed. Three arrangements were examined, one rigid roller per edge, and seven
and 13 independently rotating rollers per edge, in all cases rollers and slots were
lubricated to reduce frictional restraint to rotation.
To evaluate the experimental edge conditions, a thin aluminum plate
(a D 800 mm, b D 400 mm and t D 1.6 mm) was tested, with the different
arrangements, and experimental buckling loads and mode shapes were compared
with theoretical predictions for simple support boundary conditions. The out-of-
plane displacements at buckling, for mode shape comparisons, were measured at 20
selected points by a multipoint system of linear-motion potentiometer transducers
(LPTs). In Figure 8.40 sections of the measured buckling mode shapes (from
[8.61]) are superimposed for three different loaded edges: (a) triangular grooves,
(b) single rigid roller per edge, (c) 13 independent rollers per edge. The top of the
figure is the longitudinal mid-section (since a/b D 2, appropriately two half waves)
and the bottom the first quarter transverse section, all drawn to the same scale. The
influence of rotational restraint at the loaded edges on the buckling mode shapes,
and in particular their amplitudes, is evident, and as expected the arrangement with
13 rollers, which allows differential rotations (as permitted by theoretical simple
supports), has significantly larger deflections.
The experimental buckling loads obtained by Southwell plots (which will be
further discussed in the next section) also show the influence of the loaded
edge boundary conditions, for example a 6 percent difference between the
Experiments on Axially Compressed Plates 493

(a)

(b)

Figure 8.38 Minguez plate buckling test fixture unloaded (longitudinal) edge of specimen
(from [8.61]): (a) slots in specimen for support wires (before wires are attached),
(b) section AA showing attachment of a wire

stiffer triangular grooves and the more flexible 13 roller edges. In all cases the
experimental values of the buckling load are above the theoretical ones for simple
supports on all edges. Even with the most flexible loaded edge (13 rollers per
edge) the experimental buckling load is nearly 5 percent above the theoretical one.
Since, as has been pointed out, Minguez made considerable efforts to simulate
494 Plate Buckling

Figure 8.39 Minguez plate buckling test fixture roller support for loaded (transverse) edge
(from [8.61])

Figure 8.40 Minguez plate buckling tests sections of the buckling mode shape for different
loaded edge supports (from [8.61]): (a) loaded edges with triangular grooves,
(b) loaded edges with single rigid rollers, (c) loaded edges with 13 independent
rollers per edge

idealized simple supports (though even with lubrication, the friction of the rollers
may still represent some rotational stiffening, and also in-plane motion at the
loaded edges may not be entirely unrestrained), one can conclude that less carefully
constructed boundary conditions, as often employed, will represent considerably
stiffer boundary conditions than the nominal simple supports.
Experiments on Axially Compressed Plates 495

It may be of interest to mention here a 1968 test setup of Sharman and


Humpherson [8.62] with similar rollers for simulation of simply-supported
boundary conditions on loaded and unloaded edges of Perspex plates subjected
to combined lateral load and uniaxial compression. These rollers were short pieces
of slotted aluminum bonded at intervals to the edges of the plate, free to rotate
and slide in grooves machined in the edge support beams. The intervals between
the rollers were larger here, and only four rollers were used on each of the loaded
edges and eight on the unloaded ones. The friction was minimized by adjustment
of clearances, without recourse to lubrication. From a map of slope contours over
the whole plate obtained by the moiré method, the investigators were convinced
that neither the unsupported portions of the edges between the rollers, nor the
discontinuous changes in slope at the ends of the rollers, had any significant effect
on the deflections of the plate. However, when comparison between the deflections
predicted by theory and the experimental deflections, under uniaxial compression
only, showed the latter to be smaller and displaying a slight asymmetry, it was
realized that some frictional constraints do appear at high edge forces.
Another approach to simulation of the classical simple supports was that of
Stein at the NACA Langley Research Center in the late fifties [8.26] or [8.27]. His
52.32 in ð 25.36 in ð 0.072 in 2024-T3 aluminum alloy plate was supported by
a multiple-bay fixture (Figure 8.41), in which parts of knife edges formed 11 long
panels of aspect ratio 5. Spur gears were attached to the ends of the knife edges and
racks to the base plates, thus providing a positioning system for the edge supports.
The knife edges could rotate freely and thus allowed uniform in-plane movement
normal to the unloaded edges. A lubricant was applied to the plate under the knife
edges to minimize restraint to in-plane movement along the unloaded edges and
to leave them practically free of shear as required.
The loaded ends were “flat ends” compressed between the platens of a
1 200 000 lb. capacity hydraulic testing machine (load-control loading). “Flat ends”
for the loaded edges were deemed to be satisfactory here, as by virtue of the plates
being long, with an aspect ratio of 5, their influence would be insignificant.
A good example of a relatively simple simulation of clamped supports on all
edges is the experimental apparatus of Uemura and Byon for the study of secondary
buckling of a square plate, already mentioned earlier [8.28] and [8.34]. The exper-
imental setup is shown in Figure 8.42. Clamping is applied by rigid blocks on all
edges. At the loaded ends (x D 0, a), the plate specimens are compressed through
the rigid upper (2) and lower end blocks (3), which are firmly bolted to the spec-
imen to provide the desired clamped and in-plane boundary conditions, which
are here u D constant and v D 0. At the unloaded edges (y D 0, b), the desired in-
plane boundary conditions are Nxy D 0, Ny D 0 (or v D 0 with maximum friction).
These, and the clamping, are achieved by inserting spacers (7), which are made
from the same material with the same thickness as the specimens, between the
side blocks (1) and leaving a gap between them, thus allowing in-plane displace-
ment of the plate specimen. By adjusting the bolting forces, the in-plane boundary
conditions could be adjusted between the two extremes of v D 0 and Ny D 0. With
maximum friction, diffusion (load sharing) to side blocks was likely to occur,
496
Plate Buckling

(a)

Figure 8.41 Stein’s NACA multiple-bay test fixture showing the vertical steel knife edges to support the aluminum-alloy flat plates (from [8.27]
and courtesy of Dr. M. Stein): (a) multiple-bay fixture for eleven flat panels, (b) end view of vertical knife edges, which support the
unloaded edges of the eleven flat panels and are adjusted by spur gear sectors (continued on p. 497)
Experiments on Axially Compressed Plates 497

Figure 8.41 (continued )

but since the mean axial shortening was measured by 16 strain gages, mounted
on both sides of the plate close to the unloaded edges, the diffusion could be
assessed and the axial stress corrected accordingly. The same test rig was later
also employed for tests with simple support boundary conditions, by replacing the
clamping blocks with knife edges for the unloaded edges and with V-type blocks
for the loaded edges.
498 Plate Buckling

(a)

Figure 8.42 Test setup for clamped square plates used by Uemura and Byon in their study
of secondary buckling (from [8.34] and courtesy of Prof. M. Uemura): (a) test
setup, (b) the guiding cruciform frame for the sliding dial gages, measuring the
deflections

8.2.6 Loading Systems

In the discussion of the different types of boundary conditions, some of the rele-
vant test rigs have been shown and outlined, with emphasis on the edge fixtures.
The details of load application, of stiffness of the test setup and of displacement
measurements, however, warrant special discussion. One should note that the test
rigs can be roughly divided into two groups: (a) the small “table” models and
(b) the large “hall” test rigs, the latter being typical of civil engineering laborato-
ries. The discussion here will focus on test rigs for investigation of plate behavior,
while those primarily designed for study of stiffened plate panels will be considered
in Chapter 12, Volume 2.
Experiments on Axially Compressed Plates 499

The loading in the case of the small test rigs, is usually applied from a
universal testing machine, as for example in the classic test setup of Schuman
and Back (Figure 8.14) or in the test jigs of Hoff, Boley and Coan [8.53], of
Schlack [8.54] or of Rhodes, Harvey and Fok [8.40] or Uemura and Byon [8.34].
Usually this is displacement-control loading, applied by screw type machines or
by the deflection-control mode in modern servo-control testing machines; though
hydraulic load-control test machines have been used. For plates, on account of their
stable initial postbuckling behavior, the mode of loading control does not affect
buckling but only far postbuckling behavior near failure. Since with load control,
the load continues to act at the same magnitude as the postbuckling waves grow,
it will result in larger displacements and more violent failure.
Some small test rigs have their own loading system, usually some screw system,
as for example that of Walker ([8.41], shown in Figure 8.16) or that of Yamaki
([8.39], shown in Figure 8.43). In effect, rotation of the screw shaft in Figure 8.43,
or the thrust collars in Figure 8.16, moves the loading head, or the end loading
beams, respectively, yielding loading arrangements of high mechanical advantage
for easy manual operation and sensitivity. In Walker’s tests, eccentric loading was
required and it was obtained by differential rotation of the hand wheels (No. 6 in
Figure 8.16). Examination of the “exploded view” of the components of Walker’s
test rig, Figure 8.16, reveals the careful design of this compact and yet stiff setup
and shows the attachment of the loaded edge bearings (8) and unloaded edge knife
edges (7), discussed earlier. For accuracy of alignment, the inside faces of the
channels that carried the unloaded edge supports (7) were machined smoothly and
the faces of the bearing holders (9) were carefully ground to size. The magni-
tude and eccentricity of the applied load was measured by strain gages bonded
to the link bars (1), each bar assembly having been calibrated before in a tensile
testing machine and found to give a linear response in the test load range. In
order to obtain the distribution of the out-of-plane deflections, Walker developed
a photogrammatic system (see [8.41]), essentially photographic measurement of
the horizontal displacements of the shadows of a set of straight parallel wires,
due to vertical plate deflections; a forerunner of modern optical methods, like the
shadow-moiré technique, which are mentioned in Chapter 20, Volume 2. Agree-
ment between theory and experiment for Walker’s tests on plates subjected to
linearly-varying edge compressive loading in this test-rig was good [8.41].
Yamaki’s test setup (Figure 8.43) combines precision, accurately defined
boundary conditions and careful deflection measurements, a combination which
later typified his outstanding shell experiments, which will be discussed in
Chapter 9, Volume 2. Here the test-rig is also relatively simple, a set of a base,
two equal side support frames and an upper frame, bolted together, to which the
loading system (Figure 8.43b) is attached. The vertical movement of the loading
head is transmitted, through a dynamometer (a load measuring beam), to the upper
loading beam. The rectangular load measuring beam has strain gages, bonded on
both faces, which measure the applied load, and it is precalibrated to ensure that its
calibration curve remains linear for the entire loading range. It was found indeed to
remain linear even beyond 150 percent of the predicted failure load of the thickest
500
Plate Buckling

Figure 8.43 Yamaki’s test setup for postbuckling studies of rectangular plates (from [8.39]): (a) the test equipment, (b) loading apparatus, (c) details
of knife edges, one pair for simple supports and two pairs for clamped edges
Experiments on Axially Compressed Plates 501

square aluminum plate (300 ð 300 ð 1.25 mm) to be tested under the stiffest
boundary conditions all edges clamped. The loading beams are designed to be
much stiffer than the test plates, so as to closely simulate the loaded edge conditions
of constant displacement (which are assumed in the theoretical investigations that
these experiments aim to verify), that actually require completely rigid loading
beams. The test setup as shown in Figure 8.43 fits the square plates of the test
program of [8.39], but the bolted frame construction of the jig can easily be adapted
to plates of other dimensions.
Deflections are measured by two dial gages (see Figure 8.43a) at the position
of the theoretical maximum deflection. Four cases of boundary conditions are
treated in the experiments (simple supports and clamped edges being simulated
by one or two pairs of knife edges, as discussed earlier, see Figure 8.43c): I All
edges simply supported; II Loaded edges simply supported, other edges clamped;
III Loaded edges clamped, other edges simply supported; IV All edges clamped.
Hence the dial gages are positioned at the center of the plate in cases I, III and
IV (where buckling occurs in one half wave both horizontally and vertically), but
at a quarter length of the plate vertically in Case II, since in this case the plate
buckles in two half waves in the y direction. By coating the plates with vaseline
before fixing the knife edges, as mentioned earlier, the friction is diminished and
the unloaded edges simulate the condition of freedom from in-plane stresses, case
(b) in [8.38]. The measured maximum deflections are shown in Figure 8.44 for

Figure 8.44 Maximum deflection versus applied load, deep into the postbuckling region, for
Yamaki’s plates with the four cases of boundary condition (from [8.39])
502 Plate Buckling

loading and unloading. One may note small hysteresis loops, due to the remaining
small friction along the edges.
Most test fixtures for uniform compressive loading aim at applying constant
displacement at the loaded edges and use therefore relatively stiff loading beams
(for example those of Schuman and Back, Figure 8.14; Fischer and Harre,
Figure 8.20a; Stein, Figure 8.41; Yamaki, Figure 8.43; Schlack [8.54]; Rhodes,
Harvey and Fok [8.40], or Minguez [8.58]). Hoff, Boley and Coan, however,
suggested already in 1948 [8.53] an alternative approach, the use of flexible loading
strips, which could be bent and fitted to the edge of the plate by adjusting screws.
Using the strains measured on the plate, these screws were manipulated till a
satisfactory uniform strain distribution was obtained, hence maintaining uniform
strain instead of the usual uniform displacements. This was essentially an early
manual feed-back system, leading the way to modern feed-back loading systems
in more complex structures.
The flexible loading beams are shown in Figure 8.45 (a schematic drawing recon-
structed from the description in [8.53]). Slotted base blocks (1) ride on the loaded
edge of the 1/4 in. thick plate, separated by 1/8 in. gaps. Thin aluminum shims (2)
are used to fit the fiberglass panels into the slots of the base blocks. The base blocks
are cradled each into a three-quarter needle bearing with 1/16 diameter needles,
the cut-away bearings (3) are obtained by removing a quadrant, and one or two
additional needles for more freedom, and then sealing the open ends. The bearings
are press-fitted into steel blocks (4), which are attached to a 1/8 in. flexible steel
strip (5) with flush machine screws (6). In most cases, this flexible loading strip
is augmented by a 1/4 in. steel reinforcing bar (7) for better load distribution. The
bearing block assemblies fit into a deep slot in the loading beam (8), at intervals of

Figure 8.45 The flexible loading beam employed by Hoff, Boley and Coan in their 1948
Brooklyn Poly flat panel tests (from information in [8.53])
Experiments on Axially Compressed Plates 503

1 3/8 in. The loading beams, which are in contact with the loading platens of the
testing machine (9), were cut away to provide access to the 3/8 in. hexagon head
adjusting screws (10), spread at intervals of 2 3/4 in. which transmit the load to the
bearing assemblies, directly via the flexible strip (5) or via a reinforcing bar (7), as
shown in Figure 8.45. In addition to transmitting the load, the adjusting screws (10)
provide a means of sensitive adjustment of the applied strain distribution. Note that
since the spacing of the adjusting screws is twice that of the bearing blocks (4),
the screws bear directly on alternate bearing blocks. The remaining bearing blocks
are loaded indirectly through the flexible strip (5) and its reinforcing bar (7). The
bearing assembly (4,5,7) fits into the slot of the loading beam (8), with a maximum
clearance of 0.005 in. a fit sufficiently close to prevent the assembly from tilting,
while sufficiently free to allow the vertical displacements necessary for adjustment
of the strain distribution. The details of the edge loading distribution system have
been presented here as an example of good design for an innovative experimental
technique. These 1948 experiments stand out indeed as exemplary experiments,
and both the technique and the results have since been used and quoted by many
investigators.

8.2.7 Large Test Rigs

Earlier in the chapter, some typical examples of large “hall” test rigs have been
discussed (see Figures 8.23, 8.29 and 8.33). These large test rigs, and similar
ones, usually have their own loading system, but sometimes, as in Dwight and
Ractliffe’s test rig of Figure 8.29, the plate test-fixture is positioned between the
loading platens of a large testing machine, in this case a 500-ton Amsler hydraulic
machine. In these tests a steadily increasing strain was applied by the testing
machine and the force exerted by the specimen was measured. The investigators
wanted to strain the specimens beyond the point of the maximum load carried
and also obtain the falling part of the curve, which in the case of some of the
plates is a very steep load-shedding curve (see for example the non-welded R
specimens in Figure 8.13, especially the simply supported ones in the left half of
the figure). The hydraulic testing machine had too much give in it to prevent a
dynamic jump-through, and hence the test-fixture had to be stiffened by the three
massive T-section aluminum columns which acted in parallel with the specimen,
as shown in Figure 8.30. The combined area of the columns was 4 10 times that
of the specimen, and they were able to take a strain of 0.005 before becoming
inelastic. They were placed so that their combined centroid was the center of the
specimen. The columns were about 1/4 in. shorter than the plate and accurately
shimmed to ensure that they started to bear only when the specimen was already
taking considerable load. This stiff combination enabled the close following of the
steep unloading curves, like those of the “as rolled” steel plates in Figure 8.13.
This parallel load sharing is an example of special stiffening introduced to prevent
dynamic jump-through in unloading portions of load-shortening curves, a technique
sometimes employed also in other structures to permit accurate recording in cases
of strongly unstable initial post-buckling, or post-collapse behavior.
504 Plate Buckling

The large ETH Zürich test rigs [8.16] and [8.18] and the more recent Cambridge
University test rig [8.43], [8.44], [8.58] and [8.59] have loading systems specially
designed to the requirements of the planned test programs. As pointed out earlier,
the ETH test rigs have a “dead load” levered loading system, which for plate tests
is used with a large mechanical gain of 10:1. The principle of the levered machine
is clearly shown in a schematic diagram of the earlier 7-ton machine, Figure 8.46
(from [8.16]). The load G applies a 10 times larger tension P to the relatively stiff
loading frame that converts it to compressive loading P on the plates. The load P
immediately follows the shortening of the buckling plate, as does the load in a real
structure. On the other hand, with the aid of the crane, the load G can be arrested,
and lowered at a controlled slow rate for study of the post-buckling process. To
correct any errors arising from the inclination ˛ of the lever and other sources
the machine was precisely precalibrated. The later larger machine (Figure 8.23)
operates essentially on the same principle, except that the mechanical leverage
(gain) of 10:1 is obtained by a two lever system. The loading frame is independent
of the loading system and deformations of the latter will have no effect on the
test plate. The load is transmitted to the loading frame by a guided piston, via
a link that can move horizontally on ball bearings, thus eliminating any possible
horizontal forces on the piston which could cause friction. The loading is applied to
the loading frame as a concentrated force which is then transformed to continuous
uniform or triangular loading on the specimen, by a sophisticated load distribution
system shown in Figure 8.25.
As has already been mentioned, the 1979 horizontal large test rig at Cambridge
University (Figure 8.33) has a special wedge jack straining device, the 1 in 25 taper
providing additional mechanical gain to that obtained by the loading screw. One
face of this 80 mm thick wedge jack bears on a massive 400 ð 90 ð 1250 mm
reaction block (No. 4 in Figure 8.33), spanning between the tie bar link plates
(No. 5 there). The wedge and reaction block carry shallow lateral horizontal Vs,
of 170° included angle, which make the wedge self-centering under load. The
tapered face of the wedge bears on a 800 ð 80 tapered platen (No. 6 there), which
is guided in the load direction by needle rollers running between hardened steel

Figure 8.46 Schematic diagram of ETH seven ton levered testing machine (from [8.16])
Experiments on Axially Compressed Plates 505

male and female V-grooves (No. 7 in Figure 8.33). The wedge has a travel of
600 mm, which gives the platen a stroke of 24 mm. The bearing surfaces of the
reaction block and tapered platen carry strips of Glacier DX acetal bearing material
held in place by countersunk screws, whereas the corresponding faces of the wedge
have a ground finish. Note that at full load capacity, the pressure on the bearing
strip is 16 MN/m2 (³1.6 kg/mm2 ). This loading system has operated satisfactorily
for many years of extensive testing.
It is important here to stress the importance of the relative stiffness of loading
frames to specimens to ensure the correct boundary conditions and load distribution.
This has already been mentioned in connection with boundary conditions. Many
investigators also stress the relative stiffness of their frames in their reports (for
example Stüssi et al. in [8.18], or Bradfield in [8.43] and [8.44]). One way to
ensure the relative stiffness of the loading system is to use only a fraction of its
total load capacity, as is often planned (for example 1/5 in [8.6] and [8.39] or 1/15
in [8.27]). However, even an apparently stiff test rig has some flexibility which
may not be negligible, and hence this should be measured under test or simulated
test conditions. For example, Bradfield [8.44] measured the stiffness of the large
horizontal test rig at Cambridge University (Figure 8.33). He found that for a load
of 400 kN (³ 40 tons), a typical steel plate tested will shorten by 1.3 mm, whereas
the deflection of the rig is 1.03 mm, of which 0.73 mm is the shortening of the load
cell. Hence the stiffness of the rig is only slightly greater than the elastic stiffness
of the specimen, which indicates the difficulty of producing a large test rig which
has a stiffness greater than the elastic stiffness of the specimens. In this case, the
rig could be stiffened, but this would require a stiffer and therefore probably less
sensitive load cell. When the same test rig was later used by Mofflin and Dwight for
aluminum plates [8.59], the specimens would shorten by approximately 3.7 mm,
and then the rig was obviously 3 1/2 times as stiff as the tested aluminum plate, a
more satisfactory situation.
One may point out, that with modern multi-point strain gage measurements and
feed-back systems, the flexibility of loading beams can be compensated without
great difficulty, and eventually compensation for the flexibility of the whole test
system may be possible.

8.2.8 Special Loading Systems for Annular Plates

In the loading systems discussed, which represent the usual plate test-rigs, the
loading was essentially mechanical, by a screw shaft arrangement, hydraulic jacks
or a “dead weight” array of levers. Two test setups for buckling of annular plates
that employ other types of loading systems may therefore be of interest. One
by Rosen and Libai [8.63] employs 30 radially moving loading segments pushed
by a contained circular rubber tube filled with oil, that is pressurized by a hand
pump (see Figures 8.47 and 8.48a). As can be seen in the figures, the loading
segments have triangular grooves which were to simulate simple supports on the
outer boundary of the annular plate, whose inner boundary was free. The outer
edges of the plates were rounded (or given a wedge shape in the case of the
506 Plate Buckling

Figure 8.47 Schematic loading system of Technion test rig for buckling of annular plates (from
[8.63])

ANNULAR RUBBER
PLATE TUBE

LOADING
SEGMENTS

LVDT

ROTATING
RADIAL BRIDGE
POTENTIOMETER

Figure 8.48 Technion experimental setup for buckling of annular plates (from [8.63]):
(a) intermediate stage of assembly, (b) complete test system

thicker plates) to minimize partial constraining of edge rotation during loading.


The oil pressure in the rubber tube was measured to indicate the radial load on
the plate, but as there was friction between the segments and their housing not
all the compressive load was transferred to the plate. Pairs of strain gages were
placed circumferentially on the plate to measure the in-plane compression and
bending response of the plate, circumferential strains were preferred to radial ones,
Experiments on Axially Compressed Plates 507

as they are larger and hence yield more accurate readings. The measured in-plane
compressive strains were used to calibrate the loading system. A study of their
behavior and of the geometry of the oil tube and its containing surfaces yielded a
relation between the applied oil pressure pN and the resulting radial loading Nr (in
N/mm)
Nr D cpN  pN 0  8.12
where c is related to the slope of the line of the average experimental compressive
strains, and pN 0 accounts for friction and similar losses which are independent
N The experimentally found constant c is practically the same for all plates,
of p.
about 12 mm, whereas pN 0 varies slightly around an average of 0.07 MPa, or about
10 percent of the buckling pressure. With the constants determined, the load system
was easily employed.
This loading system was also extended to square plates by Zaal, under the
guidance of Libai [8.64]. The buckling rig for square plates aims at applying a
uniform line load to all four edges of the plate. To achieve this, 24 loading segments
are employed, which are similar in principle to those used by the circular plates,
except that instead of with oil, the contained rubber tube is pressurized with CO2 ,
which prevents the deterioration of the rubber tube which occurred when oil was
used, but reduces the safe permissible pressure. For square and rectangular plates,
the extreme corners of the plate are not loaded (see Figure 8.49), but this has a
negligible effect on the behavior of the plate under load. The friction between the
sliding segments and the mounting may cause some load diffusion but this is again
dealt with by calibration and testing with increasing pressure only. An LVDT, on
a slide mechanism, is used to measure deflections, including initial imperfections,
along a grid of parallel lines.

Figure 8.49 Technion test rig for square plates subjected to uniform biaxial line load (from
[8.64]): (a) cross-section: (1) segment, (2) tube, (3) plate; (b) details of segment;
(c) details of corner
508 Plate Buckling

The other type of loading is heating of a loading ring made of a material with
a smaller coefficient of thermal expansion than that of the test plate, which results
in a uniformly distributed radial compressive load on the annular plate, employed
by Majumdar in [8.65]. He used 0.041 in. thick 2024 aluminum alloy, circular
and annular plates, which were clamped by 12 bolts between two 0.5 in. thick
steel rings of 8 in. inner diameter and 10 in. outer diameter. The effects of the
elasticity of the steel rings on the resulting stress distribution in the plates, and on
the nominal clamped boundary conditions, were estimated to be very small and
hence approximately negligible.
Two different heating methods were used. In one test series, the assembly was
heated in a Missimers environment chamber, where the temperature of the speci-
mens could be controlled within š1° F (š0.6 ° C), and by allowing soaking periods
of an hour for each temperature increment, temperature gradients in the specimen
were minimized. Temperature-compensated radial and circumferential strain gauges
were attached in pairs on both sides of the plates. Due to the initial imperfections
some bending occurred from the beginning of loading, and a Southwell plot was
employed to predict the buckling temperature from the recorded strain-temperature
data (see Figure 8.63).
In another test series, the assembly was placed on a turntable and heated by
means of a 1000 W quartz iodine photographic lamp connected in series with a
rheostat, to permit direct measurement of deflections. Again at every increment
in temperature enough time was allowed for the assembly to reach equilibrium
temperature. Then the vertical displacements were measured with an inductance
pickup placed at a certain radius, by rotating the assembly, the output being plotted
directly on an X-Y plotter.
Some slip did, however, occur between the aluminum plates and the clamping
steel rings, and as a result the experimentally observed buckling temperatures were
slightly higher than the predicted ones, but the predicted dependence of buckling
mode on the ratio of radii was clearly confirmed.
It is interesting to note that Majumdar used as a loading mechanism the differ-
ence in the coefficients of thermal expansion of test specimen and the test rig, a
phenomenon that often can cause inaccuracies and has to be watched and compen-
sated for. For example, Lahde and Wagner in their postbuckling experiments on
rectangular plates [8.10] made a point of keeping the temperature in the room and
test equipment constant within š0.3° C, to prevent any significant thermal stresses
due to the different coefficients of thermal expansion of their brass specimens and
steel test rig.

8.2.9 Deflection Measurement

The measurement of displacements, or the determination of the buckling mode


shape, have already been briefly mentioned in relation to some of the test rigs, but
the importance of the topic justifies a separate discussion.
The simplest form of deflection measurement, a traversing dial gage, was already
used in 1930 by Schuman and Back to obtain their detailed buckling patterns [8.6],
Experiments on Axially Compressed Plates 509

as mentioned earlier, and has since been employed by many investigators, see for
example Yamaki’s 1961 test setup, Figure 8.43 [8.39], or Uemura and Byon’s
1979 test rig, Figure 8.42 [8.34]. It is extensively used today, however with the
dial gauges replaced by displacement transducers, for example LPTs in Minguez’s
1986 experiments [8.61] or LVDTs in Zaal’s 1988 ones [8.64].
Another simple, essentially qualitative, approach to the determination of the
buckling mode shape was employed by Kollbrunner in 1946 [8.16], when he
observed the mirror image of a straight steel bar in the test plate. The initiation of
buckling could thus be clearly observed and after buckling one could read the wave
number very well from the wavy image of that straight bar in the buckled plate.
A similar idea is used in undergraduate student laboratory buckling experiments
at the Technion in Haifa. In an experiment exploring the postbuckling behavior
of a 4-bay stringer-stiffened plate and expounding the concept of effective width,
the distortion of the image of an enlarged chess-board reflected in the test plate
indicates the initiation of buckling and shows the shape and growth of the buckling
waves with load (see Figure 8.50).
Another interesting, though somewhat cumbersome technique was employed by
Ramberg, McPherson and Levy in their 1939 NACA experiments on axially loaded
sheet-stringer panels. They recorded the shape of the buckles in the sheet between
the stringers for one of their specimens, by means of plaster of Paris casts. The
casts were made of the sheet side of the specimen, after it having been lightly
greased. The plaster was poured slowly into a “container”, formed with a cover
and scotch tape, and allowed to harden 5 10 minutes. The cast was then removed
from the specimen, fastened to the table of a milling machine and its contours
measured with a dial micrometer. Clear and precise contours were obtained at
different load levels.
Some investigators, as for example Lahde and Wagner in 1936 [8.10], measured
the gradients of the buckled shape rather than the displacements, using them to
calculate the bending stresses as well as the compressive stresses. (This was long
before electric strain gages were available.) The gradients were measured by a
movable mechanical gradient sensor which lightly touched the deformed plate and
amplified the measurement with a long glass pointer on a finely divided scale.
About 600 gradient measurements were taken at each loading, which yielded a
closely spaced gradient grid that conveniently lent itself to numerical integration.
Nowadays, buckling and postbuckling displacements are usually measured by
displacement transducers sliding in suitable guides, though mechanical dial gages
are still employed sometimes on account of their simplicity, or by a moiré shadow
technique whose fringe patterns yield the displacement field, as will be discussed
briefly in Chapter 20, Volume 2.
In large test rigs, as used in civil engineering laboratories or for testing ship
grillages, mechanical dial gages are used extensively also today for measurement
of initial geometrical imperfections, or initial out-of-flatness, and the out-of-plane
deflections (see for example the Cambridge University tests [8.44], [8.57], [8.59],
or [8.157], [12.72, Volume 2], and the UK Admiralty Research Establishment
full scale ship grillage tests [12.57, Volume 2]). Sometimes even simpler methods
510 Plate Buckling

Figure 8.50 Reflection of chessboard in a uniaxially compressed plate as indicator of buckling


shape, in a Technion student experiment

proved satisfactory in the large scale tests. In an earlier Cambridge University 1971
test series [8.57], the initial geometric imperfections were measured by placing a
straight edge gently against the plate and trying to push a feeler gage between
the straight edge and the plate, repeating this simple check with the straight edge
in different directions. Indeed this remarkably simple method worked very well
and the plates were found to be exceedingly flat. Also the out-of-plane deflections
were partly measured in this test series manually, with a calibrated scale read
with the unaided eye. Sometimes simplicity is really the answer. In more recent
experiments, however, the Cambridge University investigators also built special
scanning devices, like the one in Figure 8.51 (from [8.66]) used in compression
tests on plain flat outstands (plates with one free unloaded edge) and positioned in a
Experiments on Axially Compressed Plates

Figure 8.51 Cambridge University special scanning device, in which a potentiometer-type displacement transducer attached to an aluminum carriage
was tracked up and down with its probe in contact with the plate, and the output was fed into an X-Y plotter (from [8.66]): (a) the
511

vertical straining rig, (b) the scanning device


512 Plate Buckling

vertical straining rig (Figure 8.51a). In this scanner a potentiometer-type displace-


ment transducer was tracked up and down with its probe in contact with the plate,
and its output fed into an X-Y plotter. The transducer was mounted in an aluminum
carriage which ran on two parallel steel rods (see Figure 8.51b) which made up
another potentiometer-type circuit feeding the position of the probe to the plotter.
Hence an exaggerated plot of the deflection was obtained. Similar devices were later
named “ripple scanner” instruments and have been extensively employed in further
test series (for example [8.44], [8.59] and [12.72, Volume 2]) for measurement of
growth of out-of-plane displacements, while mechanical dial gages, mounted in a
simple “gage bridge”, with a gage length usually equal to the width of the plate
b (see Figure 8.52, from [8.44]), continue to be used for measurement of initial
out-of-plane deflections.
Another example of a scanning device is the plate scanner built by Fok in 1980
(see Figure 8.53, from [8.67]). Here vertical probes move up and down as they
follow the contour of the horizontally supported plate, bending thin cantilever leaf
springs to which pairs of strain gages are attached. The output from these strain
gages is then fed to a Compulog data logger, which is programmed to convert
the strains into vertical displacements. The bank of 15 probes is driven by a small
electric motor along the surface of the plate, scanning the contours of the plate, first
the initial deflections and then their growth with increasing load. The probes touch
the surface of the plate, but the contact forces are very small (the very light helical
spring on each probe roughly balances the weight of the probe) and therefore do
not introduce significant errors. In Fok’s device, shown in Figure 8.53, the motor
is stopped automatically at each of the 16 mm equispaced holes in the side plates
by a switch, and the interpreted deflection at the particular points are stored in
the data logger in matrix form. With the aid of a computer the results are then
plotted as contour lines of initial imperfections, deflection under load, stress etc.
on a Textronic display unit.

8.2.10 Controlled (Deliberate) Initial Deflections

In his experiments Fok [8.67] deliberately introduced initial geometric


imperfections by first clamping the specimen in a special rig (Figure 8.54) and then

Figure 8.52 Simple bow gage (“gage bridge”) for measurement of initial out-of-plane deflec-
tions (from [8.44])
Experiments on Axially Compressed Plates 513

Figure 8.53 Motor-driven plate scanning device for measurement of out-of-plane deflections,
developed at Monash University in Australia (from [8.67]): (a) side view of
scanner, (b) view across scanner showing 15 leaf transducers

gently heating the surface of the plate uniformly with an oxy-torch. The purpose
of the screws in the rig was to control the desired shape of the initial dishing.
During heating the specimens deflected up to 8 mm in a few cases, but on cooling
these initial deflections usually reduced to less than 3 mm, with some residual
stresses left in. The remaining initial deformations were then scanned and the stored
514 Plate Buckling

Figure 8.54 Monash University plate buckling experiments: procedure for introducing
initial imperfections into plate (from [8.67]): (a) specimen clamped in frame,
(b) specimen buckled by heating, (c) specimen held by clamps until cool

measurements used as data for computing the elastic response of the plate. It was
possible to introduce in this manner one, two or three initial half waves as desired,
which could significantly differ from the predicted wave number for the lowest
buckling stress of the corresponding perfect plate. Figure 8.55 shows an isometric
view and initial deflection contours for an initial single half-wave imperfection
obtained by scanner and plotter, compared with the deflection contours obtained
by the moiré fringe method.
As a matter of fact, Fok and Murray showed in [8.67], on one of their thus
initially dished specimens, that the plate buckled at a higher buckling load than that
predicted for the corresponding perfect plate. The measured single half-sinewave
initial imperfection was apparently of large enough amplitude to dominate the
resulting buckling mode, which had three half-waves, and prevent the develop-
ment of the inherent four half-wave buckling mode (of the corresponding perfect
flat plate), which would have yielded a 20 percent lower buckling load. Solution
of the Marguerre equations, with the measured initial single half-sinewave imper-
fection, by finite differences correlated with the experimental results, and also
predicted three half-waves buckling instead of the four half-waves of the perfect
flat plate. Hence it was shown that “by imposing large initial imperfections in a
mode dissimilar to the lowest natural buckling mode, it may be possible to arti-
ficially raise the elastic buckling load”. A very interesting and important finding!
However, one should be careful not to overestimate the practical usefulness of
this imposition of “beneficial initial imperfections” for increase of the buckling
Experiments on Axially Compressed Plates 515

(a)

(b)

Figure 8.55 Typical initial imperfections obtained in Monash University plate buckling experi-
ments (from [8.67]): (a) isometric view of initial imperfections of specimen No. 9,
(b) initial deflection contour of specimen No. 21 obtained by scanner, (c) the same
contour obtained by moiré fringe method

strength of a plate. Since the required large amplitudes could introduce significant
residual stresses and large strains that would enter the plastic region, both causing
a reduction in buckling strength, the beneficial effects may be outbalanced by these
secondary detrimental ones. Great care and extensive computations and experiment
should therefore precede any design application of this phenomena.
516 Plate Buckling

As was already mentioned earlier, many of the civil engineering investigators,


and in particular the Cambridge University team, introduced a controlled initial
out-of-plane deflection, or “bump”, in a special “bumping” rig prior to testing.
The rig constrains the longitudinal edges of the plate to remain straight while a
hydraulic jack with a load-spreading cap forms the bump. Many of the specimens
had longitudinal and transverse weld runs laid along the edges and across the center
line to simulate the residual stresses which appear in welded structural assem-
blies. The residual stress levels were measured from strain measurements, with
mechanical Demec extensometers, before and after the welding (see for example
[8.57], [8.59] or [8.157]). More recently the mechanical extensometers have been
replaced in the Cambridge University experiments by a Weldscan demountable
extensometer [8.68]. This extensometer employs a strain gaged cantilever as the
measuring element, whose electrical output is suitable for automatic data recording.
(Other techniques for measurement of residual stresses are discussed in Chapter 10,
Volume 2).
It may be pointed out that “bumping” also produces residual stresses. Hence the
preferred sequence of introduction of the geometrical imperfections and those due
to welding has to be considered. The Cambridge researchers concluded from their
studies (see for example [8.157]) that it was advisable to perform “bumping” always
after welding, for consistent state of plates prior to loading, since welding param-
eters and the initial out-of-flatness could be controlled when applied in that order.

8.3 Determination of Critical Load and Southwell’s


Method in Plates
8.3.1 Definition of the Buckling Load in Plates

The measurement of the critical stress in plate buckling is a more formidable


task. As mentioned earlier, the precise definition of the buckling stress or load
in plates is sometimes difficult, on account of their stable postbuckling behavior
(see for example Figure 8.3). Many methods have therefore been employed for its
definition. For example, Hu, Lundquist and Batdorf, in their 1946 NACA study of
simulated test data [2.15] discuss and evaluate the two main methods developed
at NACA in the mid-forties, the “strain reversal method” and the “top-of-the-knee
method.”
Observing again Figure 8.7, one notes that for a plate that has an initial deviation
from flatness, this begins to grow (or one can say that the buckles, that will
eventually be prominent, begin to grow) with the beginning of loading. Hence
there can be no buckling stress for an actual imperfect plate in the strict theoretical
sense. “However, just as a defined yield stress has been found useful for materials
that have no actual yield stress,” Hu, Lundquist and Batdorf point out, “so a
defined buckling or critical stress for a plate can convey much meaning to a
practicing engineer”. Furthermore the measured critical stress or load serves as
a calibration value for the experimental setup, and in particular for the actual
Determination of Critical Load and Southwell’s Method in Plates 517

boundary conditions, as has been pointed out by many investigators (for example
[4.20] or [8.81]). Hence much effort has been devoted by many researchers to
identify the most suitable experimental definition of the buckling stress.
The rapid increase of lateral deflection with load near the buckling load is the
basis of the experimentally defined buckling stress, but as strains are usually more
easily measured than deflections, and as the extreme-fiber strains at either side of
the buckle crest correspond to the lateral deflections, the measured strains are often
used for determination of the critical stress.
In the “strain reversal method”, the critical stress is defined as the stress at which
the extreme-fiber strain ε2 on the convex side of the buckle crest stops increasing
and starts decreasing. This can be seen in Figure 8.56, where circles show critical
stresses obtained by this method.
In the second method developed at the time at NACA, the “top-of-the-knee
method”, the critical stress is essentially the stress corresponding to the top-of-the-
knee of a curve of stress versus lateral deflection (see Figure 8.7). An alternative
abscissa would have been (ε1  ε2 ), the difference in strains in the direction of
loading at the two sides of the buckle crest, since it increases in substantially the
same manner as the lateral deflections. Both methods give experimental critical
stresses that are lower than the theoretical values for a perfect plate, the strain
reversal method generally yielding lower values than the top-of-the-knee method.
Both involve a certain degree of personal judgement and the resulting uncertainty
increases with the initial deviation from flatness.
In 1948 Hoff, Boley and Coan published their careful experiments on buckling of
rectangular fiberglass plates [8.53], which have already been mentioned in connec-
tion with boundary conditions. They also discussed the then prevalent methods of
definition of the buckling stress and suggested others. Three proposed methods of
defining Pcr (or cr ) from measured strains are shown in Figure 8.57: (a) a suffi-
ciently sharp break in the algebraic mean compressive strain εA D 12 ε1 C ε2 , (b) a
sufficiently sharp break in the transverse strain curve εT , and (c) the extrapolation
of the parts of the strain difference (ε1  ε2  curve below and above the buckling

Figure 8.56 Variation of extreme-fiber-strain ratios at plate center with average edge-
compressive-stress ratio for simply supported square plates, with slight initial
deviations from flatness, subjected to uniaxial compression (from [2.15])
518 Plate Buckling

Figure 8.57 Brooklyn Poly experiments on buckling of rectangular fiberglass plates strain
curves for determining Pcr (from [8.53])

load (note that the scale for the strain difference is shifted, and appears at the top
of the figure). All three methods yield here fairly close results, a buckling load
of about 6600 lb, but such good agreement between different methods of defini-
tion of buckling stress is not always ensured. Note that the strain reversal method
applied to ε2 (convex side) in Figure 8.57 would again yield a lower buckling
load, whereas the top-of-the-knee method applied to the (ε1  ε2 ) curve in the
figure would result in a buckling load which is close to, though still slightly less
than, that obtained by the other three methods. In Figure 8.58 (another plot of the
test data from [8.53], which was the routine plot in those tests) a fourth method
of definition is shown the inflection point method, which consists in locating
the point of inflection of the load-deflection curve or, in other words, locating the
least slope of the load deflection curve (the left hand curve in Figure 8.58). The
application of the first method, of Figure 8.57, the location of a sharp break in
εA , is also shown to a different scale (the right curve in Figure 8.58), as well as
a variation of the third method, extrapolation of the center deflection curve below
and above the buckling load (again the left curve in Figure 8.58).
Determination of Critical Load and Southwell’s Method in Plates 519

Figure 8.58 Brooklyn Poly plate buckling tests: typical load-deflection and load-mean
compressive strain curves (from [8.53])

In evaluating the inflection point method (of [8.53]), the location of the least
slope of the load-deflection curve, one should recall that in a perfectly flat plate
this slope is zero at the instant of buckling. Hence when the plate is not perfectly
flat, the location of this least slope should be in fair agreement with the theoretical
buckling load, which it seems to be. The inflection point method yields a buckling
load slightly above that obtained by the top-of-the-knee method, and the former is
less dependent on individual judgement than the latter.
The suitability of the inflection point method, for both the load-deflection curve
and the load versus strain-difference curve, was reconfirmed in Coan’s 1951 large
deflection study of plates with simulated initial deflections [2.14], in which he
reevaluated the commonly employed techniques for determination of the critical
load. Coan also indicated that the vertical tangent of the load versus average
midplate axial strain curve (used for example by the US Forest Product Labo-
ratory) is an adequate criterion, as can be seen in Figure 8.59, (reproduced from
[2.14] with some omissions) where on the right hand side of the figure the vertical
tangent has been drawn in for the computed curve and yields a value close to the
critical load. This criterion, according to Coan, is preferable to the technique of
extrapolating the postcritical slope, as is apparent in the figure. In Figure 8.59 also
experimental curves of load versus average strain for fiberglass plates 27a and 28a
520 Plate Buckling

Figure 8.59 Load ratio versus median-fiber strain ratios at center of a simply supported square
plate subjected to uniaxial compression (from [2.14]): test data from fiberglass
plates [8.53] computed, for stress-free supported unloaded edges and uniformly
displaced loaded edges, initial imperfection w0 has same shape as buckling mode

from [8.53] are shown. Extrapolation of the postcritical branch of the computed
and experimental curves could indeed yield slightly low buckling loads. On the left
hand of Figure 8.59 the midplate transverse strain εT is presented for a computed
plate and for specimen 27a of [8.53]. Coan points out that the transverse-strain
criterion may not be reliable, yielding here low values for the critical load. He
also cautioned against the then widely employed NACA top-of-the-knee method
(as for example in [6.19]) and the strain-reversal method, as being unreliable.

8.3.2 Southwell’s Method in Plates

One of the most prominent of the many methods employed for the definition of
the buckling stress of a plate, is an indirect one, the Southwell method, discussed
in detail in Chapter 4. There the extension of the method, which was originally
derived for columns, to various structures that upon buckling essentially deform
into developable surfaces has also been discussed. In the latter part of that discus-
sion in Section 4.5, following Roorda’s 1967 remarks [4.46], the importance of the
postbuckling behavior on the applicability of Southwell’s method to different types
of structures has been emphasized. For plates, their stable nonlinear postbuckling
behavior indicates the possibility of unreliable and overestimated buckling loads.
In the same vein, Donnell [4.26] had already in 1938 cautioned that for plates,
on account of the non-negligible extensional strains “the differential equations
Determination of Critical Load and Southwell’s Method in Plates 521

would be non-linear, and Southwell’s method cannot be expected to apply exactly”.


Applying an approximate energy method, he showed that the buckling load for an
imperfect plate with an initial curvature was
 
 
 
1 C 31    W C 2W0 W C W0 
2
W
PD Pc   8.13
W0 C W  8t
2
 
II
where Pc is the buckling load of the perfect plate (with no initial curvature), W0
is the amplitude of the initial deflection, W that of the additional one and t is
the thickness of the plate and  is Poisson’s ratio. Comparing Eq. (8.13) with the
corresponding expression for a column, Eq. (4.11), or for a column on an elastic
support, Eq. (4.13), indicates that Southwell’s method would give accurate results
for plates only as long as the second term in the square bracket of Eq. (8.13),
marked II, is small. This will be the case as long as W and W0 are small compared
to the thickness of the plate t.
Hence the Southwell method is applicable to plates provided: (a) W0 /t < 1 and
W/t < 1 (where < means here significantly smaller) or (b) the bent surface of
the plate is very nearly a developable surface (as in a plate hinged on three sides
and free on the fourth see Figure 4.15).
This applicability was studied extensively in the sixties and seventies with the
aim to quantify these limitations (see example [4.40], [8.69] [8.79]) and thus to
outline the range in which the useful Southwell method (as recommended for
example by Timoshenko and Gere, [2.1], p. 346) can be expected to yield fair
results.
Horton and his associates at Stanford University were among the leaders of these
efforts in the sixties. As has already been mentioned in Chapter 4, they extensively
studied the applicability of the Southwell method to plates (see [4.40] or [8.69])
as well as to shells, discussed in Chapter 9, Volume 2. For plates, they reviewed
Donnell’s analysis, examined test results of other investigators (like, for example,
[8.53], [8.41] or [8.70]) and carried out experiments on square 4-ply fiberglas
panels to demonstrate the validity of the method. Typical Southwell plots from the
Stanford University buckling tests on fiberglas panels are shown in Figure 8.60.
In these tests the loaded edges were clamped and the unloaded ones were simply
supported with various side rail clearances. As can be seen in the figure, the side
rail clearances are equivalent to initial imperfections, represented by the intercepts,
and practically do not affect the buckling load, represented by the slope.
Their experiments also included thermal buckling of plates. First a centrally
heated circular plate with free edges was studied, and a Southwell formulation
was derived [8.71]. From the experimental results, the Southwell plot, shown in
Figure 8.61 for two plates, was then obtained. In the figure, T0 and W0 are the
midpoint temperature and deflection. The critical midplate temperature determined
from the plot, was about 8 percent below the predicted one, which had been
obtained by an energy approximation.
522 Plate Buckling

Figure 8.60 Typical Southwell plots from Stanford University buckling tests on square
fiberglas plates with various side clearances which are equivalent to initial
imperfections (from [4.40])

The Southwell technique was also employed to obtain the buckling tempera-
ture of annular plates, supported at their outer edges and uniformly heated along
their inner edge [8.72]. Figure 8.62a shows a typical plot of the central edge
temperature T0 and central edge deflection W0 versus time for one of the plates.
The corresponding Southwell plot is presented in Figure 8.62b, yielding a critical
temperature about 6 percent above the theoretical predictions. Similar correlations
(about 8 percent) were obtained for the other specimens. The overestimation of
the critical temperatures may have been due here to the curving of the Southwell
plots, which can be noted in Figure 8.62b, as mentioned in Chapter 4.
One may recall that in another series of tests on annular plates at Caltech, discussed
in Section 8.2.8, Majumdar [8.65] employed heating of the outer ring as a means
to load the plate, and also used the Southwell plot to determine the critical temper-
atures. A typical plot from these experiments is presented in Figure 8.63. For this
specimen, the correlation between the experimental critical temperature obtained
by the Southwell method and the theoretical prediction was excellent, the differ-
ence between them being less than 1 percent. Some of the other seven specimens
exhibited similar excellent correlation, but for most of them the difference between
experimentally observed critical temperatures (with the aid of Southwell plots) and
predicted ones was 7 17 percent. This discrepancy was attributed to inaccuracies
in temperature measurements, slight temperature nonuniformity, and errors in strain
measurements. Better control of these factors would probably have resulted in a
more consistent correlation and given more credit to the Southwell method.
Southwell plots were also employed in the Technion experiments on annular
plates [8.63], discussed in Section 8.2.8, for two specimens. These plots were found
to be curved and therefore overestimated the buckling pressures, as expected. On
the other hand a recent application of the method to the data of another of the
specimens (taken from [8.74]) yielded good correlation between the Southwell
plot and the theoretical buckling pressure.
The applicability of the Southwell technique to plates with nonuniform prebuckle
stress state was examined by Carlson and his students at Georgia Institute of
Determination of Critical Load and Southwell’s Method in Plates 523

Figure 8.61 Southwell plot for results of Stanford University experiments on thermal buckling
of centrally heated circular plates (from [8.71])

Technology in the early seventies [8.75] [8.79]. They extended Donnell’s and
Horton’s earlier analyses and derived an expression that tended to the Southwell
form, when the term involving the first eigenfunction and eigenvalue became domi-
nant (see [8.75]). This dominance would manifest itself in an emerging linearity
of the Southwell plot at certain locations on the plate. A judicious choice of the
measurement station (for example at a node line of the second mode) could there-
fore be used to suppress higher modes and promote larger linear regions of the
Southwell plots.
The first problem they studied was a rectangular plate subjected to a partial
edge compression [8.75]. The plate was simply supported along the two opposing
524 Plate Buckling

Figure 8.62 Typical plot of central edge temperature and central edge deflection versus time for
Stanford University tests on heated annular plates and its corresponding Southwell
plot (from [8.72]): (a) temperature and deflection versus time, (b) corresponding
Southwell plot

loaded segments and free on the remainder of the edges (see Figure 8.64). In
the experiments, the compressive load was applied through roller bearings, and it
was found expedient to apply a small preload, which was taken as the reference
state. The Southwell procedure was therefore extended in the manner suggested
by Lundquist [4.25], discussed in Chapter 4 to yield
 
W  Wi
W  Wi  D Pcr  Pi   A1 8.14
P  Pi
where the subscript i refers to the reference state, W is the deflection ampli-
tude, A1 a constant representing the initial imperfection, and Pcr is the buckling
load obtained from the Southwell plot. Equation (8.14) is practically identical
with the Southwell Lundquist expression for a column presented in Chapter 4,
Eq. (4.12) there. The Lundquist plot for plates is therefore, as for columns, a
plot of [W  Wi /P  Pi ] versus W  Wi , whose inverse slope yields Pcr 
Pi . The Lundquist relation, Eq. (8.14), can also be written in terms of strain
Determination of Critical Load and Southwell’s Method in Plates 525

Figure 8.63 Typical Southwell plot for Caltech experiments on buckling of thin annular plates
under uniform compression (from [8.73]): (a) radial strain difference (measured
in mV) versus temperature, (b) corresponding Southwell plot yielding the critical
temperature Tc (compared with the predicted one
c )

Figure 8.64 Rectangular aluminum alloy plates subjected to partial compressible edge loads,
applied through 12 in. long roller bearings, tested at Georgia Tech (from [8.75]);
the aspect ratio of the plates (a/b) varied from 1/12 to 2.5
526 Plate Buckling

Figure 8.65 Typical Lundquist plots for some of the rectangular 7075-T6 aluminum alloy
plates tested at Georgia Tech with partial compressive edge loads (from [8.75])

differences ε and then the plot is [ε  εi /P  Pi ] versus ε  εi .
Typical Lundquist plots for the Georgia Tech rectangular 7075-T6 aluminum alloy
plates of different aspect ratio, subjected to the same type of partial edge load,
are shown in Figure 8.65. The plates were fairly thin, with 2b/t D 96, and yielded
very consistent straight line Lundquist plots.
Another similar case studied was that of a rectangular Alclad 2024 aluminum
alloy plate subjected to uniform compressive in-plane forces over a portion of the
edges (as Figure 8.64, only with the loaded portions being clamped and shifted to
two corners of the plate, see [8.79]). Consistent straight-line Southwell plots were
obtained, as well as good agreement with energy estimates.
Similar studies were carried out at Georgia Tech on the buckling of thin tensioned
2024-T3 aluminum alloy sheets with cracks, holes and slots (see Figure 8.66
and [8.76] [8.78]). In the experiments, the Southwell Lundquist technique was
applied, and very consistent Lundquist plots were obtained. Typical examples are
presented in Figure 8.67 for a plate with a simulated 2.2 in. long crack and in
Figure 8.68 for a plate with rounded slots. The Georgia Tech studies also included
similar thin tensioned plates with an elliptical hole [8.78]. Again very consis-
tent straight-line Southwell plots were obtained, yielding buckling loads which
correlated well with theoretical predictions.
Carlson and his co-workers emphasized that the condition for applicability of
the Southwell method to plates was that middle surface stretching due to bending
must be practically negligible. Since the term marked II in Eq. (8.13) represents
Determination of Critical Load and Southwell’s Method in Plates 527

Figure 8.66 Specimens for Georgia Tech experiments on tensioned sheets with simulated
cracks and rounded slots (from [8.76] and [8.77]): (a) side view, (b) specimen
with simulated cracks, (c) specimen with rounded slot

Figure 8.67 Typical Lundquist plot for a prestressed thin tensioned 2024-T3 aluminum alloy
sheet with a simulated crack of 2.2 in. length, tested at Georgia Tech (from [8.76])

this middle surface stretching, the condition promulgated at the beginning of the
subsection is the same as that stressed here. Carlson and his students pointed out
that satisfaction of the precondition for applicability of the Southwell technique
could be experimentally determined by examination of the mid-surface strain curve.
When this curve deviates significantly from linearity (for example, beyond point A
528 Plate Buckling

Figure 8.68 Typical Lundquist plot for a thin tensioned 2024-T3 aluminum alloy sheet with
a rounded slot, tested at Georgia Tech (from [8.77]): (a) load versus midsurface
strain curve, (b) load versus strain difference, (c) corresponding Lundquist plot

in Figure 8.68a) the range of Southwell applicability has been exceeded. One may,
however, note in Figures 8.68b and c, that slightly exceeding this limit A does not
yet endanger the validity of the method.
It should be pointed out here, that Horton et al. [4.40] summarized their studies
with the assessment that the Southwell method should be valid in plates for
displacements less than one-half the thickness.
As already mentioned in Chapter 4, Spencer and Walker [4.20] re-examined
the limitations of the Southwell method for plates and emphasized the effect of
nonlinearities at higher loads. They pointed out that in certain cases the nonlinearity
of the plot makes the identification of a defined Southwell slope very doubtful.
However, one should note that in the examples given in [4.20] the prevailing
nonlinearities occurred beyond the usual range of applicability of the method, where
the condition W0 /tW/t − 1, promulgated at the beginning of this section, was
no longer satisfied.

8.3.3 Pivotal Plots for Plates

Spencer and Walker proposed an alternative method, when the usual Southwell
plot exhibited significant nonlinearities at higher loads. This “pivotal plot” (see
Determination of Critical Load and Southwell’s Method in Plates 529

[4.20], [8.80] or [8.81]), sometimes referred to as the Spencer plot, was based
on the concept of pivotal points, which are experimental points, chosen by the
experimenter as suitable and sufficiently accurate to serve as reference points.
They developed an approximate relation,

H2 D Pcr F1 H1  W0 8.15

where

N 2  W2 /
H1 D  W 
H2 D PW N 2  PW
N 2 / 8.16
N
D [PW/W3W N  PW/
C W N N
W3 N C W]F2 
W

W0 D initial deflection of the plate,


W D additional deflection due to load,
and PN and W N are the load and additional deflection at pivotal points. F1 and F2
are correction factors:

F1 D 1 C [3W0 /WN C W]
8.17
N
F2 D 1 C 2W0 /W[1 N
C W0 /3W]
which tend to 1 if the initial imperfection W0 is small.
The proposed method was to assume first that F1 D F2 D 1 and plot H2 versus
H1 . If a straight line was obtained, its intercept with the H2 axis was W0 . The
correction factor F1 was then estimated and Pcr was calculated, according to
Eq. (8.15), from the slope of the straight line divided by the correction factor
F1 . Note that dimensionally the functions H2 and H1 of Eq. (8.16) correspond to
(W/P) and W of the Southwell plot.
The pivotal plot was applied in [4.20] to the examples presented there, whose
Southwell plots had exhibited significant nonlinearities at higher loads, and indeed
yielded straight line plots of the functions of load and deflection H2 versus H1 , as
well as reasonable estimates of Pcr .
An improved “modified pivotal plot” was later developed by Fok and Yuen [8.81]
in which the functions plotted were [H2 H3  1/H1 ] versus H2 /H1 . Here H1
and H2 are again those defined in Eq. (8.16),
N C W]
H3 D [3/W 8.18

and using again one set of P W readings as “pivotal points” one obtains

Pcr D W0 [H2 H3  1/H1 ]  H2 /H1 . 8.19

The modified pivotal plot yields a straight line whose intercept with the H2 /H1 
axis gives Pcr and whose slope represents W0 .
Application of the modified pivotal plot to the cases of [4.20] and others yielded
straight line plots and consistent Pcr and initial imperfection W0 . For example
Figure 8.69 presents the modified pivotal plot for the test data of Schlack’s
530 Plate Buckling

Figure 8.69 Modified pivotal plot for three of Schlack’s tests on square aluminum alloy plates,
without holes (from [8.80])

Table 8.1 Buckling loads and initial imperfections in Schlack’s tests on square plates with
D 0 (no hole)
Pcr W0
Test Point of Modi ed Pivotal Theoretical Modi ed Pivotal
no. in ection pivotal plot prediction pivotal plot plot
[8.54] [8.81] [2.40] [8.81] [2.40]
kN kN kN kN (mm) (mm)
1 12.89 12.64 0.66
2 15.65 12.67 0.69
3 13.55 12.64 0.71
Average 14.03 12.65 12.7 13.6 0.69 0.75

aluminum alloy, 3.18 mm thick, square plates of [8.54], for D 0 (no hole).
There appears to be no significant difference between three tests for D 0, though
Schlack obtained some differences, using an elaborate curve fitting routine to
calculate the point of inflection (his definition of the buckling load), as can be
seen in Table 8.1. One may note that the estimates of buckling load and initial
Determination of Critical Load and Southwell’s Method in Plates 531

imperfection with the modified pivotal plot and the original pivotal plot are
practically identical.
The modified pivotal plot is probably preferable to the original pivotal plot, since
it eliminated the somewhat ambiguous correction factor. Both methods depend,
however, on the judgement of the experimenter in the choice of pivotal points and
may also be prone to computational instability (see for example [8.80]). Hence,
though they represent a possible alternative to the much simpler Southwell method,
they do not supersede it.

8.3.4 More Recent Applications of Southwell Plots and


Recommendations

Indeed, the Southwell method continues to be widely used for plates, as for example
for duraluminum and laminated aluminum-boron plates in [8.82], or more recently
for aluminum and graphite-epoxy composite plates in [8.83]. In both these cases,
the Southwell plots were relied upon to provide static reference loads with which
the results of the dynamical methods (vibration correlation methods, discussed in
Chapter 15, Volume 2) were compared.
Chailleux et al. [8.82] conducted, at the Ecôle de Mines ENSTA in Paris,
careful experiments on the buckling of unidirectionally compressed aluminum
and composite plates, paying special attention to the edge conditions. Along the
unloaded sides they simulated hinged edges (simple supports) in the simple manner
shown in Figure 8.70. Two thin strips of steel lamina (0.03 mm thick) were put
on each side of the specimen. A 1.5 mm thick rubber tape was attached on one
side and slightly compressed by the vertical knife edge to hold the plate against
the other knife edge. This arrangement allowed free rotation and nearly free in-
plane displacements, though it probably did not completely eliminate out-of-plane
displacement, and was claimed to give reproducible results. Typical load-curvature
plots and the corresponding Southwell plots are presented in Figure 8.71 for two
loading sequences, each with its particular alignment adjustments. The experi-
mental buckling loads obtained from the Southwell plots in Figure 8.71b agree very
well (within 3 5 percent) with the calculated theoretical value Pth . One may note
that the experimental buckling load Pdyn , obtained from the dynamical curve (vibra-
tion frequency squared versus P, not shown here) was very close (within 1 percent)
to the values obtained from the Southwell plots. In Figure 8.71b, the three

Figure 8.70 Hinged unloaded edges in ENSTA-Paris tests


on aluminum and composite square plates
(from [8.82])
532 Plate Buckling

Figure 8.71 Southwell plots for ENSTA-Paris duralumin plates with four edges hinged (from
[8.82]): (a) load versus vertical curvature K1 at the center of the plate for two
loading sequences, (b) Southwell plot for the same loading sequence

zones of the Southwell plot are marked: Zone I that of low-load nonlinearities,
Zone II straight Southwell line, and Zone III that of high load nonlinearities.
It should be pointed out that at the transition between Zones II and III, the ratio of
maximum deflection to plate thickness is about 0.5, as found by other investigators
and promulgated as the limit of applicability in the beginning of this section.
It may be of interest to note that also in the more recent Stanford University
experiments [8.83] on aluminum and composite plates, well-defined consistent
Southwell plots were obtained and the resulting buckling loads agreed well (within
10 percent) with those found by the dynamical method developed there.
Summarizing the application of the Southwell plot to plates, the authors
feel that in spite of the criticisms, alternatives and limitations discussed the
Determination of Critical Load and Southwell’s Method in Plates 533

method is a very useful, reliable and simple tool for determination of the buckling
load of plates from experimental data. When the Southwell plot for a plate shows
significant curvature, the straight line should be fitted to the experimental points in
the neighborhood of the origin as proposed by Roorda (see Chapter 4, Section 4.5,
or [4.47]), provided local scatter or “clouding” is not excessive (when a Lundquist
plot may be used). The buckling load will then be only approximate, and will
probably overestimate the critical load, but will usually be fairly consistent. One
should recall that the Southwell method also performs well when utilized for
“smoothing” of data for parametric studies of compressed plates, as it is shown in
Chapter 4 to do for columns and in Chapter 9, Volume 2 for shells.

8.3.5 Summary of Direct Methods for Determination of Buckling


Loads in Plates

Having examined and evaluated the Southwell method and the other indirect
methods for determination of the buckling load of plates (the pivotal and modified
pivotal plots), it may be appropriate to summarize the direct methods discussed
earlier in Sub-section 8.3.1 and evaluate them. The many methods employed for the
definition of the buckling stress of plates, and the often different results obtained
by them, indicate that none of them is very precise or completely reliable. As a
matter of fact, even their evaluations carried out by different investigators (as for
example in [8.53], [8.64], [2.14] or [4.20]) differ.
Before carrying out the summary and evaluation, some of the more recent studies
on the determination of the experimental buckling load of plates [8.84] [8.86]
ought to be briefly discussed. Fok [8.84] proposed two numerical methods based
on fitting the experimental data points into a general approximate equation of
load-deflection relationship (into the post-buckling range):
 
W
P D Pcr C CW2 C 2WW0  8.20
W C W0
where W is the measured maximum deflection due to the applied load, W0 the
initial one (the initial imperfection), Pcr the critical load and the constant C 
Pcr /A2 , A being a constant depending on the boundary conditions and plate geom-
etry. Equation (8.20) is identical to that proposed by Donnell in 1938, Eq. (8.13),
if the constant A2 D [8t2 /31  2 ].
The first method, the three point technique is quick and simple. In it three sets of
readings of P and W from a test are employed to yield three equations of the form
 
Wi
Pi D Pcr C CW2i C 2Wi W0 
Wi C W0 8.21
i D 1, 2, 3

which are solved iteratively. It should, however, be remembered that if the data
points chosen are too close together erroneous results might be obtained.
534 Plate Buckling

In the second method, the least square fit technique, the square of the difference
R between the measured P and that calculated from Eq. (8.20) is

R2 D fPi  Pcr [W/W C W0 ]  CW2 C 2WW0 g2 . 8.22

Minimization of R2 with respect to Pcr , C for an assumed W0 , and then with


respect to W0 yields the conditions for least square fit. There are computer codes
commercially available, which perform the required least square curve fitting.
When the imperfections or the deflections become large, Eq. (8.20) has to be
replaced by a more accurate load-deflection relation, as indicated in [8.84].
At the Technion, a similar least square curve fitting to a Donnell type parabolic
load-deflection relation was employed for determination of static and dynamic
buckling loads of plates, [8.85] and [8.86]. The Donnell relation of Eq. (8.13) was
modified by replacing the constant [31  2 /8t2 ] with a value of 1/A2  D aN
obtained from the curve fitting of the data. This modification was adopted after
significantly different “constants” appeared in computer simulations with different
boundary conditions. The procedure proposed in [8.85] and [8.86] was therefore
least squares curve fitting of the postbuckling load-deflection data to a Donnell
type parabolic relation without a priori assumptions of aN (or 1/A2 ) and W0 . The
appropriate constants aN and W0 would then emerge from the curve fitting together
with the buckling load, as shown for example in Figure 8.72. It may be mentioned
here, that in addition to the application of the method to actual experimental results,

Figure 8.72 Application of a modified Donnell technique for the determination of the critical
load of a Technion quasi-isotropic graphite epoxy square plate (150 ð 150 mm,
0.91 mm thick from [8.86])
Determination of Critical Load and Southwell’s Method in Plates

Figure 8.73 Schematic diagrams of the commonly used direct methods for determination of the buckling loads in plates summarized in Table 8.2
535
Table 8.2 Summary of commonly used direct methods for determination of buckling loads in plates
536
No. Method Brief Description Employed extensively by See Figures Evaluation
1. Average strain method (P Sharp break in load versus Hoff et al. [8.53] Figures 8.73a, 8.58 Simple very effective method,
versus εA ) average strain Coan [2.14] which does not require any
εA D ε1 C ε2 /2 curve. Souza et al. [8.80] assumption on the initial
Fok [8.84] imperfection data or
Vann & Sehested [8.88] postbuckling behavior.
2. Inflection point Location of the point of least Hoff et al. [8.53] Figures 8.73b or c, Very satisfactory method, if
Plate Buckling

slope on the load-deflection Coan [2.14] 8.58, 8.59 initial imperfection is not
curve (the maximum rate of DTMB [8.20], [8.56] excessive.
lateral deflection with load). Zaal [8.64]
Is used on P versus Vann & Sehested [8.88]
deflection Wcenter or P Venkataramaiah &
versus strain difference Roorda [8.89]
(ε1  ε2 ) curves.
3. Vertical tangent to P versus Location of the point of the Coan [2.14] Figures 8.73d, 8.59 Very satisfactory method for
εA curve vertical tangent to this curve. Venktaramaiah & plates whose precritical and
Roorda [8.89] postcritical load-strain rates
are roughly comparable.
4. Extrapolation of P versus Extrapolation (fitting of a Hoff et al. [8.53] Figure 8.73c Often yields satisfactory
(ε1  ε2 ) curve parabola) of P versus strain Coan [2.14] estimate of critical load.
difference (ε1  ε2 ) curve
beyond buckling.
5. Fitting of parabola to P Fitting of parabola to initial, or Weller et al. [8.85] Figures 8.73e, 8.72 More reliable and better than
versus deflection Wcenter preferably deep, Abramovich et al. [8.86] method No. 4. Very
curve post-buckling curve of P Fok [8.84] satisfactory if sufficient test
versus Wcenter data well into the
postbuckling region is
available.
6. Load versus end-shortening When no average strain Zaal [8.64] Figure 8.73g Satisfactory, but less liable
curve readings are available, end than method No. 1.
shortening may be used. The
intersection of the extrapolated
postbuckling curve with the
prebuckling one determines the
buckling load.
7. Load versus transverse strain Plots of load versus the average Hoff et al. [8.53] Figures 8.73f, 8.7, Usually not reliable, because
curve transverse strain at the center Coan [2.14] 8.57 of the small magnitudes
of plate, which usually involved.
involve small magnitudes
and show no pronounced
change of slope.
8. Strain reversal Defines the buckling load as NACA [2.15],[8.90] Figures 8.73h, 8.7, Not reliable and significantly
that at which the 8.56 under-estimates the buckling
extreme-fiber compressive load.
strain ε2 on the convex side
of the buckle crest stops
increasing and begins to
decrease.
9. Top-of-the-knee The top-of-the-knee of the NACA [2.15], [6.18], Figures 8.73i or j, Of limited usefulness, because
curve of load P versus [8.91], [8.92] 8.7 the location of the
lateral deflection or versus Vann & Sehested [8.88] top-of-the-knee point is
strain difference (ε1  ε2 ). inadequately defined and
depends on the judgement
of the experimenter.
10. Load versus deflection The intercept of the tangent to Venkataramaiah Figure 8.73k Satisfactory method yielding
squared the load versus W2 curve & Roorda [8.89] fairly close lower bounds to
with the load axis defines a Pcr .
lower bound to the buckling
Determination of Critical Load and Southwell’s Method in Plates

load.
537
538 Plate Buckling

extensive numerical studies (with the ADINA finite element code [2.99]) were
carried out at the Technion on simulated plates with different boundary conditions,
confirming the suitability of the technique. One should, however, emphasize that
since this method relies on least squares fitting of postbuckling data, the availability
of test data well into the postbuckling region of the plate is a prerequisite.
Hence one is led to the conclusion (also arrived at by Fok in 1989 [8.87]), that
a simple technique, like the averaged strain method, is probably the best, provided
suitable strain gage readings were taken.
A detailed summary of the commonly used direct methods for determination of
the buckling loads in plates and their evaluation is therefore presented in Table 8.2.
The methods are also shown schematically in Figure 8.73. Note that Wcenter in the
figure and in the table represents the maximum deflection, which usually occurs in
the center of the plate for symmetrical buckling modes. The schematic diagrams of
Figure 8.73 sketch the manner in which the different methods are applied, whereas
the other figures referred to in the fourth column of Table 8.2 are examples of
actual uses in experiments. Note that in Table 8.2 some references ([8.89] [8.92])
are quoted that represent additional examples of the various methods, but have not
been discussed in the text.
In the last column of the table the methods are assessed in the light of the
experience of the authors and other investigators. Since the relative merits of the
various methods are only general trends, it is recommended to employ in practice
more than one method.
It should be stressed that the indirect methods (like the Southwell and pivotal
plots), discussed in Sub-sections 8.3.2 8.3.5 have not been included in Table 8.2
only because they have already been extensively discussed and evaluated. In partic-
ular, the recommendations on the usefulness and reliability of the Southwell method
presented at the end of Sub-section 8.3.4 should be reiterated here.

8.4 Experiments on Shear Panels


8.4.1 Buckling and Postbuckling of Shear Panels
When a rectangular plate is subjected to shear stresses along the edges (pure shear,
see Figure 8.74a), tension and compression stresses exist in the plate, equal in
magnitude to the shear stress and inclined at 45° . The destabilizing influence of the
compressive stresses is partly resisted by the tensile stresses in the perpendicular
direction. The buckling mode (Figure 8.74b) is composed of a combination of
several waveforms, which makes the analysis more difficult and laborious.
The theory of plates in shear is presented in most relevant textbooks (see for
example [2.1] [2.5], [2.8] or [6.46]). After his and Boobnoff’s pioneering work in
Russia in 1913 5, Timoshenko was the first to present a practical solution to the
buckling of rectangular plates subjected to shear by applying the energy method
[8.93]. In the thirties these calculations were extended by several investigators to
obtain more accurate buckling stresses, whereas Southwell and Skan presented in
1924 an exact solution for infinitely long plates [2.45]. In 1947 Stein and Neff
Experiments on Shear Panels 539

Figure 8.74 Rectangular (square) plate subjected to shear stresses: (a) pure shear buckling
pattern, (b) deep postbuckling pattern leading to tension field, (c) model for
complete tension field web replaced by a series of ribbons which carry tension
only

at NACA (in what is considered by many a classical paper [8.94]) improved


the critical shear stress obtained with the energy method for simply supported
plates by considering both symmetric and antisymmetric buckling configurations,
and in 1948 Budiansky and Connor [8.95] obtained accurate buckling stresses for
clamped rectangular plates. Extensive studies on buckling of plates under shear and
combined loading continued at NACA and elsewhere in the late forties and fifties
540 Plate Buckling

(see for example the review in Chapter 11 of [6.46]), while work on stiffened web
plates has been going on until today (see for example the reviews in Chapters 4
and 6 of [6.3]).
The buckling, and even more the postbuckling behavior and ultimate strength,
of a shear panel is strongly influenced by the stiffness of the edge supports and
the reinforcing stiffeners. Since the most common shear panels are the webs of
beams or plate girders (see for example Figure 8.75) much of the research focused
on them, but it should be remembered that the flat elements of aircraft wings and
fuselages also act primarily as shear panels (see for example Figure 8.76).
The postbuckling strength of stiffened plates was first explained by Wilson in
1886 [8.96]. He observed (by means of a paper model with a very thin flexible
web) “that when stiffeners were properly introduced, the web no longer resisted
by compression, but by tension, the stiffeners taking up the duty of compressive
resistance”. After extensive discussions on web stiffening just before the end of
the nineteenth century and an early mathematical formulation of a tension field by
Rode in 1916 (briefly discussed and referenced in Basler’s 1962 paper [8.104]),
Wagner developed in 1929 his “diagonal tension” theory [8.97] for very thin sheet
and rigid reinforcements. The theory of “pure diagonal tension” was later extended
to “incomplete diagonal tension” theory, and both were widely used in aircraft
design and summarized in a convenient manual [8.98] with test results [8.99] and
in Kuhn’s book [8.100] or Hertel’s book [8.101]. Whereas in the “pure diagonal
tension” theory the web is considered as inclined tension members in a frame,
the extended theory also takes into account the compressive stresses in the plate
and its reinforcing contribution to the stiffeners. The results of the “pure diagonal
tension” theory were thus found to be actually limited to loads exceeding the initial
buckling load of the panel by at least an order of magnitude.
During the fifties and sixties, utilization of the stable postbuckling behavior of
plates in shear was extended from the aeronautical uses to civil engineering applica-
tions, mainly for plate girders. Extensive studies were therefore carried out on steel
and aluminum alloy plate girders by many civil engineering investigators (see for
example [8.102] [8.107]). In a review of the ultimate load methods for prediction
of the failure loads of plate girders, Rockey [8.102] stated that “aircraft structures

Figure 8.75 Collapse of a large scale plate girder tested at University College, Cardiff
(from [8.103])
Experiments on Shear Panels 541

Figure 8.76 Postbuckled shear panels on a wing of a Boeing Stratocruiser during flight (prob-
ably in a steep turn) as viewed from a window of the airplane

normally fail when the web plate tears, whereas the steel plate girders used in
civil engineering have more flexible flanges and fail by either the development of
plastic mechanisms involving the web and the flanges or by lateral buckling of the
compression flanges”. Indeed, the diagonal tension theory, adequate for aircraft
structures, with their relatively rigid stiffeners, was found to be insufficient for
civil engineering applications, where the ratio of applied shear load to critical load
of panel buckling is often less than four.
542 Plate Buckling

In the sixties and seventies theories were developed for the failure of civil
engineering-type plate girders (see for example [4.19], [6.3] or [8.103]). Initially
they were based on the extreme assumption of Basler [8.104] that the flanges
were so flexible that they could not withstand lateral loading. This was modified
by Rockey and Skaloud [8.106] to a more realistic assumption that the flexural
rigidity of the flanges contributed to the strength of the girder and that plastic hinges
developed in the flanges when failure occurred. Their method was substantiated
by extensive test data. One should note that the flexural stiffness of the flanges
significantly affects the initial buckling behavior of the shear panel, but even more
so its postbuckling behavior.
In the last decade, the trend to optimize the design of shear panels and the
employment of composites and higher strength materials, led to similar required
relative stiffnesses in both civil and aerospace engineering. The civil engineers
employ stiffer flanges in order to improve the postbuckling strength of the web
and the aeronautical engineers decrease the relative flange cross-sectional area
in order to save weight. Improved analyses were developed (see for example
[8.107] [8.109]) in parallel with extensive experimental studies (see for example
[8.109] [8.111]).
It should be pointed out, that although this section focuses on pure shear loading,
in practical plate girders also the bending strength and the interaction between
bending and shear has to be considered.

8.4.2 Experiments on Plates Subjected to Shear Picture Frames

The experiments on buckling and postbuckling of flat shear panels can be divided
into two main groups: one, in which the behavior of the plate is studied, in well-
defined but unrealistic stiff boundary conditions, the so-called picture frames; and
the second, in which the behavior of the plate is investigated with more realistic
boundary conditions of webs in beams, usually called Wagner beams when the
webs are thin. In the second group, various types of stiffened plate girders, as
well as the important influence of the relative stiffness of flanges and vertical and
horizontal stiffeners, have been extensively studied.
A broader array of typical shear-test fixtures is shown in Figure 8.77, but the
picture frames (a) and (b) and the three point beam, or three point Wagner beam,
fixture (c) are the primary ones, which will be discussed here.
The picture frame is essentially an assembly of four pairs of rigid steel edge
members (sometimes called loading tabs) pin-jointed at their ends, to which the
plate specimen is bolted. This frame forms a mechanism which transforms an
applied tensile force (from a universal testing machine) to a shear loading on
the specimen. Figure 8.78 shows a typical picture frame employed at University
College, Cardiff, UK for quasi-static cycling loading tests on unstiffened aluminum
plate shear panels.
Ideally, the test section of a square shear panel (or a rectangular one in other
test fixtures) will deform into a parallelogram. In this case, the frame members,
or loading tabs, undergo no significant deformation, since their in-plane stiffness
Experiments on Shear Panels 543

Figure 8.77 Typical shear-test fixtures schematic (from [8.110])

is one to two orders of magnitude greater (usually 30 times or more) than that
of the test specimen. In practice, however, some problems arise in picture frames
that were discussed in detail in [8.110] and [8.112] and lead to the development of
improved versions of this type of test fixtures. First, rotational clearance of frame
members has to be ensured in all picture frames to permit unrestrained kinematics.
This is usually obtained in a manner similar to that shown in Figure 8.78. Then
the corners of the specimens have to be cut away in circular arcs, slits or tangent
circle cutout, to clear the pins and allow the shear deformation of the panel.
Two major problems which plagued the conventional picture frames were the
bending and extension of the frame members, which caused non-uniform shear
loading, and the corner-pin location, which could also result in shear non-uniformity
and appearance of significant local normal stresses.
To overcome the first problem, an improved version of picture frame, the so-
called “Modified Picture Frame”, was developed at NASA Langley in the late
seventies [8.112]. In it the frame was loaded biaxially: a tension load applied
with a universal testing machine in one direction, and simultaneously a compres-
sion load of equal magnitude applied in the transverse direction with a specially
designed system (consisting of an hydraulic cylinder, a load cell and tension bars,
see Figure 8.79). By this biaxial loading the bending and extension of the frame
members, that occur in a uniaxially loaded picture frame, is mutually cancelled (see
Figure 8.80). Also shear lag is eliminated, resulting in uniform shear load, and the
stresses in the heavy corner pins are halved. On the other hand, the loading system
is more complicated and expensive, on account of the precision required in its
manufacture. Once made, however, it presents much better defined shear loading
conditions, and is often used nowadays (see for example [8.113]).
In the development of their biaxial shear frame, Bush and Weller also carried
out some frame friction tests to detect any frictional effects. They measured
544 Plate Buckling

Figure 8.78 Conventional picture frame for shear tests on flat plates employed at University
College, Cardiff (from [8.111])

experimentally in their frame the shear modulus of two aluminum plates and
obtained the identical value given in the appropriate Mil handbook. Hence they
could conclude that frictional effects were negligible in their modified picture
frame.
The influence of the corner-pin location has been studied both by finite element
analyses and experiments (see for example [8.110]) with the conclusion that for
uniform shear stress distribution the corner pins have to be located at the corners of
the test section of the panel, which prescribes the correct picture-frame kinematics.
Experiments on Shear Panels 545

Figure 8.79 Modified biaxially loaded picture frame developed at NASA Langley
(from [8.112])

Further shear panel experiments with picture frames at NASA Langley and else-
where enforced this corner-pin location requirement, either with the pins extending
through the test panels as before (as for example in [8.113]), or by separating the
corner-pins in the middle, to avoid interference with the test panel. The latter
arrangement, where the test panel is not penetrated by the corner pins, is usually
preferred today, as for example in the test setup PApS (see Figure 8.81) used at
the Institut für Flugzeugbau und Leichtbau, Technical University Braunschweig,
for repeated buckling experiments on aluminum shear panels [8.114]. Note that
here (in Figure 8.81) the effective panel size is fairly large (500 ð 500 mm2 ) and
that the shear loading is not applied by tension along the diagonal, but directly as
shear, like in a rail shear fixture.
546 Plate Buckling

Figure 8.80 Schematic of shear frame deformations for uniaxial and biaxial loading
(from [8.112])

8.4.3 Strength Tests on Plate Girders Under Shear

The beam type shear panel test fixtures serve both for experiments on buckling and
postbuckling behavior of flat plates in more realistic, but still predetermined, labo-
ratory boundary conditions, as well as for shear strength tests of practical designs
of plate girders with different types of stiffeners (like for example [8.115]), that
will be discussed further in Chapter 12, Volume 2. Their shear collapse behavior
will, however, be dealt with here as it involves primarily shear panel failure in a
flexible frame.
For over two decades one of the major research centers for buckling and
collapse studies of steel and aluminum alloy plate girders under shear has been
the Department of Civil and Structural Engineering at University College, Cardiff,
UK (now University of Wales, College of Cardiff), where Rockey, Evans and their
co-workers and students developed methods of design and analysis and carried
out extensive test programs (see for example [8.102], [8.103], [8.106] [8.109],
[8.115]).
The tests at Cardiff, and in many other laboratories all over the world, were
usually on beams in a three-point loading system. The exceptions were the
NACA Langley tests in the forties and fifties (see [8.99] and [8.100]) which were
mostly on cantilever beams and only their last series employed three-point loading.
Some typical steel plate girders of the earlier Cardiff tests [8.106] and their loading
are shown in Figure 8.82. They included panels of different aspect ratios and flange
flexural stiffnesses, whose effect was studied. One may note that in all three types
of girders shown in the figures, there were “overhangs” to stiffen the outer reaction
supports, but only in series 1 was the central point of load application reinforced by
a double transverse stiffener. Thus the series 1 beams provided test data for single
Experiments on Shear Panels 547

(a)

Figure 8.81 The Technical University Braunschweig shear panel frame PApS (from [8.114]):
(a) general arrangement of test setup, (b) aluminum shear panel under test in the
PApS setup

bay girders, whereas the series 2 and 3 beams dealt with the behavior of “two-
bay” girders, allowing for the influence of continuity effects. It may be pointed
out that the geometrical characteristics of plate girders are usually defined by three
non-dimensional parameters: the web aspect ratio b/d, the web slenderness ratio
d/t and a flange strength parameter
548 Plate Buckling

Figure 8.82 Typical Cardiff steel plate girders tested in the seventies (from [8.106]): (a) details
of series 1 girders with aspect ratios ˛ D 0.79 2.0 and web slenderness d/t D ¾
230, (b) details of series 2 girders with square panels, aspect ratio ˛ D 1.0 and
web slendernesses d/t D 150 or 316, (c) details of series 3 girders with aspect
ratio ˛ D 1.0 and web slendernesses d/t D 183 320

MŁp D Mpf /d2 t0.2  8.23

where Mpf is the plastic moment of the flange plate, d is the depth of the web,
b its width, t its thickness and 0.2 is the measured 0.2 percent proof stress of the
flange material.
The test girders were usually simply supported at their ends on roller supports
and the test loads applied by an hydraulic jack at the center (see for example
Figure 8.83). In the more recent tests (for example, [8.109], [8.115] or [8.116])
Experiments on Shear Panels 549

Figure 8.83 Test setup for a typical plate girder subjected to shear, investigated at the Cardiff
laboratory (courtesy of Professor H.R. Evans)

the jack was servo-controlled, allowing the load to be applied so as to achieve and
maintain specified deflections. At the start of each test, the initial deformations of
the webs, flanges and stiffeners were measured. Then the displacements of webs
and stiffeners, as well as the strains developed, were measured and recorded at
selected load levels. Displacements were measured with accurate dial gages and
strains with strain gages, placed where possible at the same location on opposite
faces to yield bending and membrane stresses. A continuous record was also kept of
the load-deflection response during each test. In addition to the panel and stiffener
behavior, the distortion of the compression flanges was also monitored.
The Cardiff group developed the concept of the failure sway mechanism, a
collapse mechanism with four plastic hinges in the flanges that allows a shear sway
displacement to develop (see Figure 8.84b). As mentioned earlier (see Figure 8.74),
prior to buckling equal tensile and compressive stresses develop within a web
plate subjected to shear. After buckling, the plate cannot carry further compressive
stresses and a new load carrying mechanism develops, whereby any additional
shear loading is supported by an inclined tensile membrane stress field (“diagonal
tension”). As the applied loading further increases, the tensile membrane stress
grows until it (combined with the original buckling stress) reaches the yield stress
of the material. When the web has yielded, final collapse will occur when plastic
hinges have formed in the flanges (E,F,G,H in Figure 8.84a) that permit a shear
sway displacement c (Figure 8.84b), where c indicates the position of the plastic
hinge on the flange (assuming cc D ct D c) and is the virtual angular displace-
ment of the yield zone. By consideration of the virtual work done within the yield
zone during the sway, simple formulae were developed to predict the ultimate shear
load of the plate girder (see for example [8.103] or [8.107]). One may note that the
550 Plate Buckling

Figure 8.84 Shear sway collapse mechanism in steel plate girders loaded predominantly in
shear (from [8.107]): (a) yield zone leading to the sway collapse mechanism, (b)
sway involved in the mechanism and action of tensile membrane stresses within
the yield zone

yield zone EFGH in Figure 8.84a is the minimum region that must yield before the
collapse sway mechanism can develop, but the yield zone can also spread outside
this region.
For the case in which the girder web is subjected to a bending moment in addition
to the shear load, the interaction effects have to be taken into account, usually by
interaction diagrams (presented for example in [6.3], [8.103], or [8.107]). The shear
sway failure mechanism is modified by the effects of bending, as can be seen in
Figure 8.85, in which the sway mechanisms for pure shear and combined shear
and bending are compared.
The concept of the shear sway collapse mechanism was developed for steel
girders, based on the typical yield behavior of steel. Recent experimental and theo-
retical studies on aluminum plate girders ([8.109], [8.115] and [8.116]) indicated
that though a shear sway collapse mechanism similar to that occurring in steel
girders developed also in the aluminum girders, cracks appeared in the weld-heat-
affected zones of the web plates as the sway mechanism progressed. The theory
and design methods based on this collapse mechanism had therefore to be modified
and refined for application to aluminum girders. In some of the experiments the
web panel was extensively instrumented adjacent to the compression flange and to
one transverse stiffener of the beam, in order to investigate the shear stress distri-
bution there [8.109]. The measured shear stress distribution roughly verified that
predicted by the improved analysis for the incomplete tension field of aluminum
alloy girders developed by the Cardiff group.
It should be pointed out that concurrently with the work at Cardiff, extensive
research on the collapse of plate girders, including hundreds of tests, was carried
out at many other research centers, for example Lehigh University, Liége, Tokyo,
London, Göteborg and Budapest, as is also evident in reviews of the field (see
Chapter 6 of [6.3] or [8.117], or Chapter 4 of [8.119]).
Experiments on Shear Panels 551

Figure 8.85 Comparison of shear sway collapse mechanisms in plate girders (from [8.107]):
(a) symmetrical girder loaded in shear only, (b) same girder loaded in shear and
bending, note that cc is not equal anymore to ct

The predominant stiffening of plate girders subjected to shear is by trans-


verse (vertical) stiffeners. But longitudinal (horizontal) stiffeners have also been
employed and investigated (see for example [8.107], [8.115], [8.118] or [8.120]),
and all longitudinally stiffened girders have clearly exhibited shear sway collapse
mechanisms. The extensive studies carried out in Cardiff [8.120] on steel girders
grouped into no stiffeners, weak, intermediate strength and strong longitudinal
stiffeners, showed, however, the strong influence of the relative rigidity of the
stiffeners on the collapse mechanism of the girder. When the longitudinal stiff-
eners were inadequate, as for example in the panels with intermediate strength
stiffeners (like Figure 8.86), an unstable collapse mechanism appeared, when one
or more of the stiffeners deflected out-of-plane. Hence it appears that further study,
leading to better and less conservative design methods for longitudinally stiffened
girders is still warranted.
552 Plate Buckling

Figure 8.86 Cardiff tests on longitudinally stiffened steel plate girders Panel L58-B, with
intermediate strength longitudinal stiffeners, after failure. Note that only the upper
stiffener failed, permitting a nearly continuous buckle past it (from [8.120])

8.4.4 Technion Repeated Buckling Tests on Shear Panels

In the eighties an extensive investigation was carried out at the Technion Aircraft
Structures Laboratory on the capability of stiffened metal shear panels to with-
stand repeated buckling [8.121] [8.125]. These were motivated by early failures
of aircraft spar shear panels, that could be attributed to repeated deep buck-
ling. Though essentially a buckling-fatigue problem, the study involved also static
ultimate strength tests and its test setup represents typical modern shear panel
experiments. These investigations have since also been extended to composite shear
panels (see for example [8.125]) which are discussed in Chapter 14, Volume 2.
The test setup for the “Wagner beams” is shown in Figure 8.87 and a typical
specimen in Figure 8.88. The loading is a three point type, with the load being
introduced through a pin into the center of the beam by an hydraulic jack via
a load cell, which is connected to two heavy parallel vertical bars into which
the pin is inserted. The load is controlled through an MTS system that permits
precise control of the applied load and loading frequency (from zero to maximum
load in a constant amplitude mode). The reaction loads at the edges of the beam
are also applied by pins inserted into the specimen and into two heavy parallel
vertical bars at each end, which are simply connected via ball bearings to the
heavy middle cross beam of the loading frame, well below the specimen. The
free rotation of these pairs of vertical bars in the plane of the shear web provide
any longitudinal movement required by the ends of the Wagner beam during its
Experiments on Shear Panels 553

Figure 8.87 Technion repeated buckling experiments on shear panels the test setup (from
[8.122]): (1) test specimen, (2) cross members for prevention of lateral displace-
ment, (3) loading frame, (4) video recorder, (5) TV monitor, (6) moiré lamp, (7)
multichannel data logger, (8) MTS hydraulic jack, (9) cameras

postbuckling deformation. The applied loads are introduced into the shear web via
the vertical stiffeners (uprights).
To preclude premature failure of the beams by lateral instability, transverse
deflections of the flanges are prevented by means of rigid parallel horizontal cross
members attached to the loading frame along the top and bottom flanges of the
specimens.
The test beams shown in Figure 8.88 consist each of five shear webs symmet-
rically framed by identical L section flanges and uprights. Two of the webs serve
as the test section fields. Through the other three webs, the central one and the
edge ones, the loads are introduced into the beam, as mentioned earlier. These
webs are therefore heavily stiffened in a manner which eliminates any local failure
in the loading zone and assures almost uniform shear load diffusion from the
loading pins into the test panels. Such Wagner-beam-type structures, with uprights
and flanges acting as stiffeners and with the shear transmitted into the test panels
from the neighboring panels and from the flanges, represent common aeronautical
applications in so far as stiffening and load transmission are concerned. The load
is primarily shear, and the effects of additional moments which are common to
the three-point loading configuration, although here of no major importance, were
considered to contribute to the realistic test environment.
In the first phase of the program, dealing with metal shear panels, 19 Wagner
beams were tested. All the specimens were manufactured from 2024 T3 aluminum
alloy plates and L section type stiffeners. The beams were fabricated using a
bonding process developed by Israel Aircraft Industries. In this process, the
554
Plate Buckling

Figure 8.88 Technion repeated buckling experiments on shear panels typical bonded aluminum alloy Wagner beam specimen (from [8.122])
Experiments on Shear Panels 555

longitudinal flanges and intermediate uprights were bonded to the web in a specially
designed template. The controlled bonding procedures developed by Israel Aircraft
Industries complied with aircraft standards, so as to be able to sustain the estimated
load spectrums planned for the test program. Indeed there was no premature
debonding in the tests.
The response of the specimens was measured by strain gages bonded to the
shear web surfaces, flanges and uprights of the beam specimen and recorded by
a multichannel data logger. Two or three pairs of face to face strain gage rosettes
were bonded to the test shear webs, one for detection of incipient buckling, and one
or two in the critical corners (where the deep buckle along the diagonal interacted
with the relative stiff framing). In more recent experiments (on composite shear
panels, see [8.125]), more extensive strain gage coverage with a higher capacity
data logger, connected to a PC system, was employed.
For the repeated buckling tests, the strain gage records are limited to quite a low
number of cycles relative to the overall life of the specimens, because the strain
gages cannot sustain the very high cycling strains experienced by the beams and
have a very low endurance life under such circumstances.
Since the strain measurements are always confined to localized areas, the
shadow-moiré technique is employed for overall observation of the progressive
behavior of the buckled shear web, as well as for comparison of the deflected
shape of the beam with predictions (see for example Figure 8.89, showing the
moiré fringe pattern and the STAGS [2.53] prediction for the deflection pattern
of specimen WB-7).
For simplicity of calibration of the moiré fringe pattern, the grid was positioned
very close to the shear web (less than 1 cm) and the light source and camera
far away from the beam at a distance of L D 285 cm and D D 285 cm apart. The
camera was located at a right angle to the test web whereas the light source made a
45° angle with the test field. Two types of grids were employed: one with 20 lines
per centimeter, or wavelength of p D 0.05 cm and the other one with 40 lines per
centimeter, or wavelength of p D 0.025 cm. An approximate calibration formula
W L
³p 8.24
n D
yielded a deflection of 0.05 cm/fringe and 0.025 cm/fringe for the two grids
respectively.
It should be noted that the deflection measurements obtained by this technique
alone are limited in scope; they do not furnish in-plane data and the bending
strains obtained are inaccurate, due to the low sensitivity of the technique (approx-
imately š0.1 mm). Hence a combination of both moiré deflection and strain gage
measurements is required to obtain an adequate representation of the response of
the web.
The main purpose of these Technion experiments was the investigation of the
durability of efficient shear panels, utilizing the deep postbuckling region, and the
influence of the surrounding structure. The studies yielded simple design formulae
(see [8.124]) which relate the life of the metal shear panels to three geometrical-
physical parameters, or approximately to a single dominant physical parameter
556 Plate Buckling

(b)

Figure 8.89 Technion repeated buckling experiments on shear panels comparison of exper-
imental shadow-moiré fringe pattern (a) and deflection pattern predicted by the
STAGS code (b) for beam WB-7 (from [8.123])

Vy (the shear load at which local yielding first takes place). The associated static
ultimate strength tests and residual strength tests, however, represent fully fledged
collapse tests of plate girders.
One interesting point on test procedure may be noted. In the early stages of the
test program, the two test sections were tested simultaneously. This was aimed
at checking the repeatability of the test data points, as well as confirming the
symmetry of the test rig. But then, from the third beam on, in order to save on
specimen cost and preparation time, two different tests were carried out on each
beam: either cyclic and ultimate static or two cyclings at different load levels. This
was made possible by heavily stiffening one of the test sections externally, to avoid
Experiments on Shear Panels 557

Figure 8.90 Technion repeated buckling experiments on shear panels typical failure modes
(from [8.123]): (a) tearing (tension field) failure mode for relatively stiff flanges
and uprights, (b) failure by bending and buckling of flanges when they are rela-
tively weak
558 Plate Buckling

its damage as much as possible during the cyclic loading phase up to failure of
the other section. A stop was also inserted between the center of the beam and the
loading frame to prevent large deflections of the beam at failure.
After the first section had failed, it was removed and replaced by a relatively
stiff dummy one. The beam was then placed back in the loading frame in an
inverted position to the former repeated loading test and retested. Thus the previ-
ously highly stressed and possibly damaged tension field of the web, which did
not fail, would experience compression stresses, whereas the portion of the web
formerly in compression would be loaded in tension under either repeated or static
further loading of the beam. Though some asymmetry was introduced, the consis-
tency of the results indicated that the more economic test procedure could be
adopted.

8.4.5 Aerospace Industrial Test Setups

To illustrate the type of test setups used in the aerospace industry for shear panel
tests, two examples of such tests will be briefly discussed: one for the Boeing 757
airliner and one for the SAAB 340 regional turboprop.
The 757 shear panel test carried out by the Boeing Commercial Airplane
Company in the late eighties, shown in Figure 8.91, represents a section of the
stiffened shear web of a wing spar. The stiffeners are closely spaced and hence the
shear panels between them are narrow and relatively rigid. At 90 percent design
ultimate load buckling initiates across stiffeners as can be seen in the middle of
Figure 8.91a. Note that the external frame, through which the stiffened panel is
loaded is very rigid and therefore acts as a kind of “picture frame”. As buckling
progresses with increased loading, many stiffener fasteners (rivets) fail which
facilitates the crossing of the buckling waves (see Figure 8.91b, which shows the
other side of the panel, clearly exhibiting the buckling waves across the stiffeners
whose fasteners have failed).
The SAAB 340 pure shear test of a curved panel, representing a fuselage side
with window cut outs [8.126] was carried out by the SAAB-SCANIA Company,
Linköping, Sweden in the early eighties. The instrumented panel, fabricated as
close as possible to the actual full size side panel of the 340, was mounted in
a rail shear fixture and subjected to pure shear loading (see Figure 8.92a). The
purpose of the test was to verify the stability of the fuselage and window frames
and the buckling behavior in the adjacent areas. The window frames were therefore
installed and one of the windows was also fitted with window glass in the test.
The test specimen was instrumented with 44 strain gages, 24 located at its
center on both sides of the panel and 20 on one of the window frames (see Figure
8.92b and c). Shear displacements and deflections around the window frames were
measured by 28 dial indicators. The primary failure mode appeared to be buckling
collapse of fuselage frames (see Figure 8.92b), most of the distortion remaining
permanent (see Figure 8.92c). After failure the glue lines within the actual test
area were carefully checked and no debonding was noted. Prior to testing, the
sheet thicknesses of the specimen and the material properties were also measured.
Experiments on Shear Panels 559

Figure 8.91 Boeing shear panel test for the 757 airliner: (a) the panel in the loading frame
at 90 percent design ultimate load with onset of buckling across stiffener, (b) the
other side of the panel at a higher load showing failure of many stiffener rivets
(courtesy of the Boeing Commercial Airplane Company)
560 Plate Buckling

Figure 8.92 SAAB pure shear test of 340 regional turboprop fuselage side: (a) the panel
mounted in a rail shear fixture subjected to pure shear loading, (b) panel at
failure collapse of fuselage frames, (c) permanent distortion of failed panel (cour-
tesy of SAAB Aircraft AB)
Web Crippling 561

8.5 Web Crippling

8.5.1 Web Crippling Due to Concentrated or Patch Loads

When a thin-walled plate girder is subjected to a concentrated in-plane


load sometimes referred to as localized edge loading, or to partial in-plane edge
loading usually called patch loading, it can fail either by web squashing, web
crippling or web buckling. Web squashing means local yielding of the plate in the
immediate neighborhood of the concentrated or patch load and therefore relates
only to thick-walled girders. A typical case of such local yielding has been shown in
Figure 2.26 in Chapter 2, where patch loading has already been briefly discussed.
Web buckling occurs when the distribution length of patch loading is larger than
half the depth of the web, and then the buckling of the web involves its whole
depth, representing complete buckling of a whole section of the web. Web crippling
is the intermediate local instability failure of a thin web under a concentrated or
narrow patch load, restricted to a portion of the web depth only and can be elastic
or plastic depending on the thickness of the web. All three possible modes of
failure have to be considered by the designer and have recently been discussed
with design implications by Herzog [8.127]. The discussion here will, however,
focus on web crippling.
Localized in-plane compressive edge loads act frequently on plate girders, such
as for example wheel or roller loads on crane girders or bridge girders. The web
behavior under these concentrated or patch loads, and in particular web crippling,
has therefore been investigated since the thirties, and more actively as the girders
and webs became more thin-walled in later years. Most of these studies (as can
be seen in recent surveys like [8.127] [8.131] or Chapter 6 of [6.3]), were exper-
imental, and since the theoretical analysis of web crippling is rather complicated,
the design rules have been primarily empirical.
The theoretical studies began with the calculation of elastic buckling loads of
webs. The earlier analyses were for plates compressed by two equal and opposite
in-plane forces, and then, from the mid-thirties, following Girkman [8.132], also
for actual patch loadings (as shown in Figure 2.24). The failure loads obtained
in experiments however considerably exceeded the elastic predictions (even those
obtained in the seventies by finite element analysis, like the predictions of [2.20]
discussed in Sub-section 2.1.7), and it was observed that the collapse is primarily
a plastic buckling phenomenon. As a matter of fact, web crippling under patch
loading was sometimes defined as the failure occurring “due to the formation
of plastic hinges in the flanges accompanied by yield lines in the web” (see for
example [8.128] and Figure 2.26).
Many simple empirical formulae were proposed in the seventies based on these
observed failure loads, the earliest ones being those of Granholm and Bergfelt based
on the investigations carried out at Chalmers University of Technology, Göteborg
(see for example [8.133] or [8.134]).
Since many experiments indicated the plastic hinges in the flange and yield
lines in the web, mentioned above, a model of a collapse mechanism shown in
562 Plate Buckling

Figure 8.93 Web crippling a plastic collapse mechanism with plastic hinges in the flange and
yield lines in web allowing web bending (from [8.136])

Figure 8.93 was proposed by Roberts and Rockey (see [8.128], [8.135] or [8.136]).
In the model, ˛ and ˇ define the position of the assumed yield lines in the web
and that of the plastic hinges in the flanges and
defines the deformation of the
web prior to collapse. Notice that the assumed local bending displacement of the
web resembles somewhat the measured ones of Figure 2.25.
By equating the external and internal work (neglecting the relatively small energy
associated with stretching of the web), a simple lower bound solution was derived,
which was then modified to include the effects of simultaneous bending of the
girder, resulting in a predicted collapse load:
Pu D 0.5 tw2 [Ew tf /tw ]0.5 [1 C 3c/dtw /tf 1.5 ][1  b /w 2 ]0.5 8.25
where tw is the thickness of the web, tf that of the flange, c the loaded length of
the patch load, d the height of the web, w is the yield stress of the web (which
for simplicity is assumed to approximately equal that of the flange, w ¾ D f ) and
b is the bending stress due to the coexistent bending of the girder. The last square
bracket in Eq. (8.25) includes the bending effect and the factor 3c/d in the second
square brackets is a simplifying replacement of the complex function that would
more accurately represent the girder dimensions and material properties.
The values of collapse load Pu predicted by the mechanism for web crippling,
Eq. (8.25), is compared in Figure 8.94 with over one hundred test results Pexp from
various sources for girders with tw D 1  5 mm (with a limitation of tf /tw ½ 3).
The mean value of the ratio Pexp /Pu  in Figure 8.94 is 1.43 and the coefficient of
variation is 15.8 percent. Figure 8.94 indicates that the collapse mechanism model
and Eq. (8.25) yield fairly reasonable predictions. A similar comparison with elastic
critical loads would have shown much larger ratios Pexp /Ppredicted , in particular
for thinner webs. One may note that, except the bending reduction factor, for which
alternative expressions have been proposed (for example in [8.131]), Eq. (8.25) is
today accepted as an appropriate design formula (by AISC and others).
Web crippling is a frequent cause of failure in tee and cross joints of rectangular
hollow sections (RHS) and has therefore also been extensively investigated for such
structural elements (see for example [8.137] [8.139], where [8.138] also includes a
comprehensive review of previous studies). Typical web crippling in the chord side
wall of a rectangular hollow section full width cross joint is shown in Figure 8.95a,
Web Crippling 563

Figure 8.94 Web crippling comparison of predicted collapse loads, derived from a mechanism
for web bending with test results from various sources (from [8.128])

Figure 8.95 Typical web crippling in tee and cross joints of rectangular hollow sections (RHS):
(a) web crippling in the chord side of a rectangular hollow section full width
cross joint (from [8.139]), (b) further examples of web crippling in representative
structural elements (from [8.138])

and further examples of web crippling in representative structural elements appear


in Figure 8.95b.
Web crippling has long been widely recognized as an important collapse hazard
for plate girders employed in bridges, buildings or cranes. But with the increase
use of thin-walled, high-strength steel structural components in the automotive
industry in recent years, its applicability as a likely failure mode has been consider-
ably broadened. This motivated investigations of web crippling as part of a research
project “Design of Automotive Structural Components Using High Strength Steels”
at the University of Missouri-Rolla (see for example [8.140], which includes
extensive experimental studies of high-strength steel hat sections and I-beams).
564 Plate Buckling

It should be noted that in these studies, as well as in many others (like for
example [8.128], [8.141] or [8.142]), combinations of web crippling and bending
moments have also been considered, since combined loading by bending moments
and concentrated forces presents a common design case.

8.5.2 Web Crippling Tests

It has been pointed out above that due to the complexity of web crippling analyses,
the design formulae are primarily empirical. Comprehensive test programs have
therefore been carried out at research centers in many countries, like those at
Cardiff, Rolla, Prague, Warsaw, Göteborg, Paris, Toronto, Sydney, Nagano, Zürich,
and others (see for example [8.133] [8.146]) and improved empirical formulae
have been proposed. For example, Herzog in 1986 [8.147]) reviewed and analyzed
164 tests, and more recently in 1992 [8.127] presented further empirical formulae
based on 340 tests.
In typical concentrated or patch loading tests, a central load P usually acts on a
beam supported at two end locations (where it is often reinforced to eliminate any
patch loading effects at the reaction loads, see Figure 8.96). The reaction supports
are commonly rollers, to allow horizontal displacements, and the central load P is
applied by a servo-controlled hydraulic jack or in a universal testing machine.
This loading actually includes some bending, but usually the concentrated load
effects predominate. In one recent experimental investigation on rectangular hollow
sections, at the University of Sydney [8.142], the bending interaction was studied in
detail. Three types of loading and support were employed: (1) a pure concentrated
force test (Figure 8.97) in which the specimens were loaded by a loading ram
through a central branch member and were seated on the solid steel base plate of
the testing machine; (2) a pure bending test (Figure 8.98) in which the specimens
were loaded via a spreader beam and two widely apart half rounds and loading
plates, and were supported by two half rounds on Teflon pads (to permit horizontal
displacements as required by “simple supports”), and thus a constant pure bending
moment was applied to the central portion of the test specimen; (3) an interaction
test (Figure 8.99) in which the load was applied by a central loading ram via a
central branch member while the end supports were again two half rounds on
Teflon pads, resembling the common test configuration shown in Figure 8.96.

Figure 8.96 Typical patch loading tests, central load acting on a beam supported at
ends schematic (from [8.145])
Web Crippling 565

Figure 8.97 University of Sydney experiments on T-joints in rectangular hollow


sections pure-concentrated-load test (from [8.142]): (a) schematic view of test
setup, (b) transducer arrangement

Figure 8.98 University of Sydney experiments on T-joints in rectangular hollow sections pure
bending test (from [8.142])
566 Plate Buckling

Figure 8.99 University of Sydney experiments on T-joints in rectangular hollow


sections interaction test of concentrated load and bending (from [8.142]): (a)
schematic view of test setup, (b) transducer arrangement

Cold-formed rectangular hollow sections (RHS) and square hollow sections


(SHS) were tested. The specimens were made of steel with a nominal yield stress
of 350 MPa, and were produced by cold-forming and electric-resistance welding.
The branch members (whose width was either half, ˇ D 0.5, or equal that of the
chord members, ˇ D 1.0) were welded to the chord member by fillet welds (except
for ˇ D 1.0, when the longitudinal welds were butt welds). Also the support plates,
which transmitted the loads and reactions to the chord members in the bending
and interaction tests, were welded by fillet welds.
In the pure-concentrated force tests (Figure 8.97), the base plate of the Dartec
testing machine provided continuous support to the specimens along their entire
length. The deformations were measured by six transducers, with transducers 3 6
located on the specimen centerline, as shown in the figure, and recorded on an
automatic data-acquisition system. Also in the bending and interaction tests, the
deformations were measured in a similar manner (see Figures 8.98 and 8.99),
supplemented by strain gages whose readings were also recorded on the same
data-acquisition system.
Web Crippling 567

Web-crippling was the dominating failure mode when ˇ D 1.0 (branch members
width identical to that of the chord member) in the pure-concentrated-force and
interaction tests. The web-crippling capacity of the specimens in the interaction
tests was reduced by the bending moments, due to the longitudinal stresses they
produce in webs of the RHS chord members. The Sydney tests provided interaction
diagrams and indicated when the interaction effects become significant, but pointed
out the need for further study.
Another test program at the University of Toronto studied the effect of chord
wall (or web) slenderness h0 /t0  and bearing length h1  on the crippling load of
RHS members (see Figure 8.95 and [8.139]). A series of 15 joint specimens were
prepared from one continuous piece of 252.7 ð 102.4 ð 4.44 mm cold-formed,
stress-relieved steel RHS, thus minimizing the effect of variations in yield stress
and manufacturing tolerances on the parametric study. The resulting RHS chord
members provided two wall slenderness values, 56.9 and 23.1, depending on the
member orientation in the test. The branch members applying the patch load were
rigid blocks of varying bearing length h1 , some welded to the RHS chord and some
not welded for comparison. The resulting specimen family is shown in Figure 8.100
and represents an example of an experimental parametric study. A typical testing
arrangement is shown in Figure 8.101. The experimental results were compared
with a finite element simulation yielding good agreement, thus permitting augmen-
tation of the experimental parametric study with FEM calculations. A general trend
of reduction in the web crippling strength of the RHS sidewalls (or webs) with

Figure 8.100 University of Toronto experimental parametric study of web crippling in rectan-
gular hollow sections “family photograph” of tested specimens (from [8.139])
568 Plate Buckling

Figure 8.101 University of Toronto experiments on web crippling in rectangular hollow


sections testing arrangement (from [8.139])

increase in wall slenderness h0 /t0  was observed, as well as a linear increase in
web crippling strength with welded bearing length h1 of the branch member.
In web-crippling tests, one sometimes classifies the test arrangements into the
following four basic loading conditions (see Figure 8.102 and [8.140]):

1. Interior one-flange loading (IOF).


2. End one-flange loading (EOF).
3. Interior two-flange loading (ITF).
4. End two-flange loading (ETF).

The interior one-flange loading (IOF see Figure 8.102a), is the usual simply
supported flexural member subjected to a concentrated load, discussed earlier (for
example Figure 8.96 or Figure 8.99). In the tests on high strength steel hat sections
Web Crippling 569

Figure 8.102 Web crippling tests classification of loading conditions (from [8.140]): (a) (d)
basic loading conditions, (e) (g) transition ranges. The clear distance between
bearing plates is usually presented in terms of h, the depth of the web

and I-beams at the University of Missouri-Rolla [8.140], a 2-in. bearing plate was
placed under the load, which was either mid-span or unsymmetrically located, while
4-in. bearing plates were used at both ends. To prevent premature end failures,
wooden blocks were inserted at both ends of the specimens.
The end one-flange loading (EOF see Figure 8.102b), differs in the bearing
plates which are now shorter at the ends. In the Rolla tests they were 4-in. under
the load and only 2-in. long at the ends. In both IOF and EOF loading, only one
flange is subjected to concentrated loads.
570 Plate Buckling

In the two-flange loading cases, ITF and ETF (see Figures 8.102c and 8.102d),
both flanges are subjected to concentrated loads at the same spanwise location. The
Rolla tests used 2-in. bearing plates in both cases. For the ETF loading an elastic
support was provided at the unloaded end to keep the specimens in a horizontal
position.
Transition cases can occur, and Figures 8.102e 8.102g show such transition test
arrangements used in the Rolla tests.
The test specimens in the University of Missouri-Rolla experiments were hat
sections and I-beams fabricated from five different types of high-strength sheet
steels. The tests were performed in a 120,000 lb. universal testing machine. The
lateral deformations during the tests were measured by closely spaced LVDTs at the
location of expected failure. For some hat sections the vertical strain distribution
was also investigated by pairs of strain gages. Based on the 264 tests performed
with a large range of yield strengths of steels, new design recommendations were
proposed in [8.140].
Before closing this section, it may be of interest to note that whereas the behavior
of unstiffened webs subjected to concentrated or patch loads is well-understood,
the investigations of stiffened webs are still incomplete (some recent studies carried
out at the University of Maine are discussed in [8.131]) and warrant future study.

8.6 Biaxial Loading


8.6.1 Plates Under Multiple Loading
Section 8.2 dealt with the experiments developed for the basic standard problem
of a rectangular plate under uniaxial compression. In Section 8.4 another basic
loading condition, that of shear, was discussed. While these render fundamental
information and provide essential starting points, it is necessary to consider also
multiple loadings which may cause plates to buckle in different modes, though
experiments for these become more difficult.
Some reference to multiple loading has already been made in Section 8.4, for
the case of combined shear and bending. Some analytical and many numerical
studies have been carried out in the last decades on the behavior of plates
subjected to combined in-plane loading with shear, combined in-plane bending
and axial compression, combined in-plane loading with lateral loading and biaxial
in-plane compression (see for example Chapter 6 of [6.3], Chapter 4 of [6.14], or
[8.148] [8.152]). Experiments on multiple loadings have, however, been relatively
scarce. To remain within the scope of our discussion, only one type of load
combination, biaxial compression, will be dealt with in the following.

8.6.2 Biaxial In-Plane Compression Tests


Many structures such as box girders, double bottoms of ship hulls, ship deck
structures or dock gates have plate elements that are subjected to in-plane
Biaxial Loading 571

compressive biaxial loading. Usually the load in one direction predominates, with
the smaller load in the transverse direction producing an additional destabilizing
effect. While extensive theoretical solutions are available (see for example
[8.150] [8.152] or [8.153], where NSHELL and STAGS C codes were applied),
there is only very little experimental information.
The earliest experiments were apparently two series of tests on steel plates
carried out by Becker and his co-workers in the USA in the seventies (see
[8.154] [8.156]). The specimens in these tests were small square tubes of 0.64 mm
thick AISI 1010 mild steel and width 19.2 57.6 mm, or of 0.76 mm 4130 stainless
steel and width 22.8 or 53.2 mm, electron beam welded, and in some cases stress-
relieved. The lengths of the specimens varied between 57.6 mm and 173 mm to
keep the aspect ratio of all plates as 3. The specimens were loaded longitudinally
by applying load to the ends of the tube, and transversely by applying four equal
inward loads along the corners of the tube (see Figure 8.103). Each specimen
therefore comprised four similar plates (of a/b D 3 and b/t D 30 90) under equal
loadings.
In such square plates the rotational out-of-plane boundary conditions of the four
plates along the longitudinal edges are elastic restraints, which however are very
weak (and actually approach rotational freedom) since the similar plates will buckle
in a mode that permits rotation of the corners without affecting the right angle at
the corners of the square tube.
In the first series of the Becker tests, the longitudinal loading was applied
through stiff platens. Friction between the platens and the tube ends restrained

Figure 8.103 Becker’s experiments on steel plates under biaxial compression: (a) schematic
loading arrangement, (b) one of the four equal biaxially loaded plates tested in
each specimen
572 Plate Buckling

the out-of-plane displacements of each face at the short edges, when longitudinal
loads were applied.
The transverse loading fixture applied loads to the two adjacent corners of the
tube, and reacted to them against the other two corners. The loading and reacting
members were relatively stiff to distribute the loads, which were applied pneumat-
ically, along the length of the tube. The loads were applied to the corners of the
square tube via pieces of steel or beryllium-copper shims.
As pointed out by the Cambridge researchers, who analyzed the Becker tests in
preparation for their own experiments [8.157], the use of stiff members to apply the
longitudinal loadings was likely to cause problems in load measurements. In biaxial
loading, part of the applied longitudinal load would be transferred by friction into
the transverse load spreader, and the load actually carried in the plates would
be less than that measured at the platen. Similarly, part of the applied transverse
loading would be transferred to the stiff platens.
In the second series of the Becker tests, therefore, the corner loads were applied
through a whiffletree of articulated links to exert eight equal loads on the short
ears welded to the corners of the square tube. The whiffletree could impose strong
restraints on the out-of-plane rotations of the plates, but they were extenuated by
the flexibility of the ears that transmit the loads. Though the whiffletree system was
claimed to permit transverse loading without constraining the longitudinal straining
of the specimen, this was doubted by the Cambridge team, since it would have
required considerable clearances at the pin joints.
Dwight and his students at Cambridge University aimed at expanding the limited
available experimental data, and overcome some of the difficulties encountered
in the earlier US Ship Structure Committee tests. The primary loading in their
experiments was longitudinal compression with smaller additional destabilizing
loadings. In the tests discussed here [8.157] the additional loading was a second
(transverse) in-plane compression as in Becker’s experiments. Also other studies
on the influence of additional shear loadings and that of in-plane and rotational
restraints of the unloaded edges were initiated at Cambridge.
The precise Cambridge plate test rig, employed in previous uniaxial compression
tests and discussed in Sub-section 8.2.7 (see Figures 8.33 8.35), was used as the
longitudinal loading system for the biaxial compression experiments after suitable
modifications. The stiff wedge-jack for application of longitudinal loading and the
systems of fingers for out-of-plane support of the longitudinal edges remained the
same. The tests were also carried out on 6 mm thick structural high-yield steel
plates, similar to those used in the earlier uniaxial compression tests.
The tests were not only carefully designed and performed, but also precisely
reported in [8.157]. Since they are of considerable interest to future investigators,
they are discussed and quoted here in detail.
The specimens were long, rectangular plates of aspect ratio (a/b) of 4 and 6.
The dominant longitudinal loading generated the higher stresses and was applied
to the short edges, while the smaller transverse loading was applied over a portion
of the longer edges, over length of 2b and 4b. Contrary to the Becker tests, no
lateral pressure loading was applied.
Biaxial Loading 573

“The longitudinal edges were simply supported against out-of-plane deflections.


The failure mode was therefore expected to be a limited modification of that
for uniaxial loading, with out-of-plane deflections forming approximately square
buckles. It was expressly intended that failure would occur in a buckle at the center
of the plate length, as the plates contained an initial dent in this position. The longi-
tudinal loading was applied and measured at the plate ends, and to ensure that this
load was maintained throughout the plate length, all attachments to the longitu-
dinal edges were designed to be flexible in the x-direction. This was possible as the
loadings on the longitudinal edge remained well below the plate yield loading.”
As can be seen in Figure 8.33, the short ends of the plate were loaded through
massive plates. Hence, if the transverse loading were also applied through a stiff
plate, the total load measured would be the sum of the plate transverse strength plus
the loads carried through the longitudinal loading platens, transferred by friction
between plate and platens, which also depend on clamping forces and longitudinal
loads. The tests would then yield too optimistic estimates of the transverse load
capacity.
Instead, the transverse loading was therefore applied as a constant force per
unit length, but only over the central section of the longitudinal edges, maintaining
uniform loading in this area, but not right up to the ends of the plate. If the loading
had been extended into end areas, friction effects would cause this load to be partly
transmitted via the platen. Uniform stresses in the end regions would still not be
achieved. The absence of the transverse loading in the end areas also avoided any
premature plate failure due to the unavoidable loading irregularities at the platens.
The transverse in-plane loads on the longitudinal edges were applied “through
a set of prongs, as shown in Figure 8.104 and in the photograph, Figure 8.105. In
plane, the prongs were inserted in two out of every three gaps between adjacent
fingers. Locally, the loading is irregular, but the variations in loading would be
restricted to a narrow strip, of say 3t width, adjacent to the edge. Prong center to
center spacings were therefore 20 mm and 40 mm alternately.
In elevation, the height of the prongs was reduced, terminating in a 90° included-
angle Vee, of 1.0 mm tip radius. This engaged in a Vee groove 2.0 mm wide
ð1.0 mm deep machined in the plate-edge. The prongs were machined from
12 mm diameter steel bar.” . . . “Tests showed this end detail to be capable of
carrying loads up to 1/3 of the plate yield stress without damage to the prong, or
excessive indentation of the plate.”
Pairs of prongs were attached to a mild steel cap. On one edge, each cap was
loaded through an Enerpac short stroke hydraulic jack. . . . “On the other edge,
dummies were used in place of the jacks. Each pair of jacks or dummies bore on
a short reaction member machined from 127 ð 63 channel, to which was welded a
vertical flat bar. Corresponding assemblies on each side of the plate were connected
by 60 ð 10 mm tie bars above and below the plate, the lower tie bar passing through
the gap between a pair of fingers. The tie bar bolt hole locations allowed the plate
width to be increased in steps of 5t.”
. . . “The deadweight of each transverse loading assembly was suspended from a
frame supported off the main longitudinal tie bars. Each jack and tie bar assembly
574 Plate Buckling

Figure 8.104 Cambridge University tests on plates under biaxial compression transverse
loading system, shown for a part of one longitudinal edge (from [8.157])

was therefore free to float in the x-direction, and would move freely as the plate
shortened under load. Within each assembly, the prongs provided little restraint to
movements of the plate edge in the x-direction. The measured stiffness between
the two prongs on each cap was 210 N/mm.
Each pair of jacks shared their hydraulic supply through a permanently connected
splitter. All jacks in use for a test were connected to a common manifold. Pressure
in the hydraulic system was maintained through an Enerpac jack (rated at 428 kN
at 690 bar), which was placed in an Avery 1 MN universal testing machine.
Biaxial Loading 575

Figure 8.105 Cambridge University tests on plates under biaxial compression biaxial loading
system, view from above (from [8.157])

Figure 8.106 Cambridge University tests on plates under biaxial compression schematic
arrangement of hydraulics of the transverse loading system and its control via
the Avery testing machine (from 8.157)

Once the hydraulic system had been filled, load control was exercised through
the Avery machine, taking advantage of its motor-driven pump, long scale for load
measurement, and sensitive valves.” The arrangement is shown schematically in
Figure 8.106. Note that use of an independent loading system instead would have
required duplication of several of the testing machine features.
576 Plate Buckling

As for the previous uniaxially loaded tests [8.44] discussed in Sub-sections 8.2.4
and 8.2.7, “the tests aimed to study the overall behaviour of the plates, by measuring
the longitudinal load-shortening curve for each plate. Sufficient out-of-plane deflec-
tion measurements were made to identify the mode in which the plate failed. Major
attention was concentrated on the maximum load carried, and the shape of the load-
deflection curve beyond maximum load. Most tests were performed under constant
transverse loading, but sufficient tests were performed under proportional loading
to allow assessment of the effects of different loading paths.”
Some modifications to the instrumentation employed in the uniaxial tests were
required to fit the transverse loading system, the transverse load itself being
measured through the Avery testing machine which controlled it.
Specimens were made of the same material and were manufactured in the same
manner as in the uniaxial experiments. The “tests were performed on a uniform
batch of material, and are directly comparable with previous uniaxially loaded tests.
Test conditions were chosen to represent practical structures containing substan-
tial initial out-of-plane displacements, (deliberately introduced), and covered both
unwelded and realistically welded plates”, as in the uniaxial case. “Magnitudes
of transverse compressive loading were chosen to study as much as practicable
of the interaction curve between longitudinal and transverse load, consistent with
the capacity of the testing rig and with remaining within buckle shapes indicating
dominance of longitudinal loading.”
As noted earlier, “the transverse loading was defined to be everywhere equal
along the loaded portion of the longitudinal edge. This effectively precluded oper-
ation in a post-buckling regime where transverse load is largely carried in two
zones adjacent to each loaded end, with considerably smaller transverse stresses
in the central part of the plate. The present tests therefore represent circumstances
where this post-buckled capacity is not significant, and therefore are representative
of very long panels. The distinction between these boundary conditions and those
of ‘longitudinal edges remaining straight’ should be borne in mind.”
. . . “The majority of tests were conducted at constant transverse load Ny . In these
cases the preliminary steps in the loading were to firstly apply a longitudinal load
of 40 kN to bed in the ends of the plate. The transverse load was then increased
to its intended value, and was maintained at this level. Longitudinal straining was
then begun.
Three tests were performed under proportional loading. In each case the test
followed a corresponding test at constant Ny . The loading sequence was chosen
to increase Ny , normally in 10 steps, at predetermined values of the longitudinal
load, such that the intended maximum value of Ny was reached just before the
longitudinal load reached the peak value obtained from the test at constant Ny .”
. . . “All plates failed by the formation of a single, large buckle at the center,
corresponding in position to the initial dent.” The buckle lengths were roughly
similar to the plate width, b to 1.3b. The plates were therefore within the regime
of fairly short buckles.
. . . “The uniformity of stresses in the central, loaded, fraction of the plate
was tested by comparing results for plates with aspect ratios a/b D 4 and 6,
Guidelines to Modern Plate Buckling Experiments 577

maintaining other plate parameters equal.” The comparison showed that the longi-
tudinal strengths were indeed very similar, within 2 percent.
The test results were presented in [8.157] as interaction diagrams, together with
those of both series of Becker’s tests. Whereas for uniaxial loading the Cambridge
results were consistent with those of the two Becker test series, there were differ-
ences for biaxial loading.
As mentioned earlier in this section in the discussion of Becker’s tests, the
transference of some of the longitudinal loading by friction into the transverse load
spreader and vice versa, should result in higher apparent capacities there. Indeed
the Cambridge tests, where load redistribution was prevented, yielded as expected
lower transverse capacities ym and less favorable interaction curves, compared to
those of Becker’s experiments. The interaction curves between longitudinal and
transverse strength yielded by the careful Cambridge tests also differed slightly
from the theoretical tests, and suggested that a more conservative interaction curve
should be used in design.
Before leaving this topic, it should be pointed out that biaxial loading is of
special significance in composite plates, on account of the anisotropic properties
of the material. Hence biaxial buckling of laminated composite plates has been
studied in recent years, and attention has been paid to experimental techniques. For
example, means to minimize the interference of biaxial compression at the corners
of rectangular plates have recently been developed. One particular and apparently
successful technique was to modify the specimens by cutting their corners resulting
in so called “modified specimens” (see [8.158]). The problem of multiaxial loading
of composite structural elements is further discussed in Chapter 14, Volume 2.

8.7 Guidelines to Modern Plate Buckling Experiments


8.7.1 Guidelines or Ideas for Future Tests
The extensive efforts of many investigators in different countries have been
reviewed, and in selected cases in detail, as their plate buckling test rigs and
measurement systems present possible solutions for future experimental studies.
Summarizing the lessons learned from these efforts, one may propose guidelines
or ideas for the preparation of a new series of plate buckling experiments.
The first basic decision is whether the experiments will be on “isolated” plates,
with intentionally unrealistic but well defined boundary conditions that can be
precisely reproduced by theoretical or numerical analyses, or will they be on plates
that are part of a stiffener-reinforced panel, where the plate boundary conditions
are as in a real structure dependent on the behavior of bordering stiffeners and
adjoining plates. It is assumed here that isolated plate experiments are chosen,
while the other avenue of tests on complete stiffened panels will be discussed in
Chapter 12, Volume 2. The edge conditions represent therefore the primary exper-
imental problem to be addressed.
There are now two basic approaches: (a) long plates with aspect ratio a/b ½ 4,
and (b) short plates with a/b  2.
578 Plate Buckling

The long plate approach aims at removing the loaded edges far away to eliminate
or diminish their influence. It can then be argued (as did the Cambridge researchers,
[8.39], [8.40] or [8.44]) that elastic buckling of a long plate occurs in a set of iden-
tical buckles in nearly square sub-panels, and that failure will finally initiate in only
one such sub-panel. Provided that the loaded ends do not fail, the end sub-panels
may differ slightly and may even be slightly stiffer, but they offer little restraint
to the deformation of adjacent sub-panels, which then represent idealized plates,
with simply supported loaded edges, and the type of unloaded edges defined by the
long edge support system (finger, knife edges, grooves etc.). The loaded edges in
this approach can then be rigidly clamped to prevent premature failure of the end
sub-panel and ensure efficient load transfer. However, for very thin long plates their
influence may not be entirely negligible in the far postbuckling range [8.29], [8.30].
The short plate approach, on the other hand, requires special attention to the
loaded edges to assure precise definition of the boundary conditions as well as
uniform and effective load transfer, and the same care for the unloaded edges as
before. The short plates, however, require a much smaller test rig.
This brings up the question of scale. At first it would seem that the smaller the
scale of the model the lower the cost of the experiments. Full scale tests are expen-
sive, primarily because of the large loading and reacting systems required. But they
present the most convenient means for simulation of actual fabrication methods
and boundary conditions and allow accurate measurements of local stresses and
displacements. On the other hand, the loading equipment for small models is
usually available in most laboratories and their test rigs are much less expensive.
But their fabrication, instrumentation and simulation of realistic boundary condi-
tions becomes more difficult and therefore also expensive, as the scale is reduced.
The factors affecting the choice of model scale have therefore to be considered
carefully. They have been discussed in Chapter 5 here, as well as in the literature.
For example, in [8.159] they were examined both from the technical and cost-
benefit aspects, with the conclusions:
“1. Due to the very wide range of model testing, it is not practicable to prescribe
a set of definite rules for the selection of scale factors.
2. In the experience of the members of the panel, steel products less than 3 mm in
thickness are likely to possess a wide scatter of material properties, sometimes
very significantly different from structural steel. Additionally, conventional
manufacturing and fabrication processes are likely to produce much higher
levels of residual stresses in models made of products less than 3 mm thick
than are likely to be present in prototype structures. It is recommended that
special attention should be given to this aspect in choosing the model scale.
3. When a series of small scale tests is undertaken it is recommended that there
should be a limited number of larger scale tests to confirm the results.”
The panel mentioned above consisted of the most eminent British civil
engineering researchers, and their recommendations represent the consensus at the
end of the seventies.
In other sections of their report [8.159] detailed suggestions were presented for
many aspects of testing, which could assist future experimenters and are worth
Guidelines to Modern Plate Buckling Experiments 579

looking into. As an example, the “recommendations for good practice” in the case
of static loading tests (which include buckling) on structural models are quoted
here (with some minor omissions and additions):
“1. The loads should be applied by hydraulic or screw jacks. A system of control
should be provided.
2. The overall accuracy and repeatability of the load-measuring system should
be equivalent to that of a grade ‘A’ testing machine.
3. Loading jacks and load cells should be provided with spherical bearings or
suitable alternatives.
4. The reaction bearings should be designed so that the position and direction
of the load reactions relative to the specimen remain within defined limits
during loading.
5. Where additional loading or reaction points are needed to provide moment
and shear in required portions, local stiffening of the model may be needed
to form a satisfactory load path into the specimen.
6. The stability of the test specimen (in the test rig) when under load (in the
event of catastrophic collapse) should be carefully considered.
7. The loading jacks should be operated under deflection control during the
inelastic stages of the test, but load control during the predominantly elastic
stages is optional.
8. The load should be applied in increments interspersed with pauses for the
making of measurements and observations.
9. The rate of application of load should be such that in the most highly
stressed part of the specimen, the rate of change of strain should not exceed
300 microstrain per minute.
10. After each increment of loading, sufficient time should be allowed for the
specimen to reach a stable condition before measurements are made.
11. If after any increment of loading the need for a close examination of the
specimen delays the application of the next increment, then an additional
set of load and gauge measurements should be made.
12. The entire loading history of the model should be recorded and this should
include the loads sustained at each increment of loading. The first application
of load is particularly important.
13. When loading to collapse has been commenced and substantial plastic flow
has occurred then the programme of the incremental loading should not be
interrupted until the test is complete.
14. Time related recordings of the behaviour of the model should be made at
selected stations as the collapse condition is approached. The collapse load
should be defined as the maximum load sustained.”
In some of the precise plate buckling experiments described in this chapter, it
has been pointed out how these factors were weighed carefully and what ingenious
solutions were employed to make and measure the small scale models.
In all plate buckling tests the importance of the boundary conditions has been
stressed, and in Chapter 11, Volume 2 the influence of the boundary conditions is
again reviewed and discussed for experiments on all types of structural elements.
580 Plate Buckling

The importance of the boundary conditions was also pointed out in a compre-
hensive comparative evaluation of previous plate buckling and postbuckling tests
performed in the mid-sixties by Davidson at Lehigh University [8.160]. The report
is also of interest because it discussed critically many experimental details in the
studies reviewed. For example, in the discussion of Ojalvo and Hull’s 1958 tests
on 24 aluminum alloy plates [8.30], it was noted that the possibility of partial load
transfer was considered by the authors. It was then pointed out that:
“An experimental determination of the load transfer from the plate to the jig was
carried out. Strain gages were attached to the longitudinal edges in order to measure
the vertical strain. The strain at the bottom of the plate was found to be approxi-
mately equal to that at the top which would not be the case if some of the load was
being transferred to the jig. The maximum load for which the strain was recorded
was approximately twice the buckling load and half of the ultimate load. It would
then appear that for this range of loading the load transfer was negligible.
It was noted by the authors, however, that failure of each specimen was caused
by tearing along a longitudinal line between the loading bar and the groove. This
would indicate a large shearing stress undoubtedly caused by vertical restraint at
the longitudinal edges. Thus it would appear that a fairly substantial portion of the
load was transferred to the jig by the time the ultimate load was reached.”
Another example of the details discussed by Davison were the experiments of
Botman and Besseling on aluminum plates in the early fifties [8.23] and [8.24]:
“The purpose of these tests was to approximate as closely as possible the behavior
of a plate as it exists in a stiffened plate panel. Instead of a stiffened panel, multi-
bay panels having from one to five bays and supported by knife edges at the
stiffener points were tested (see Figure 8.107). The load carried by the plate could
then be measured directly, which would not be the case if a stiffened panel were
tested since part of the load would be carried by the stiffeners. The knife edges
provided the same out-of-plane restraint as would the stiffeners.
. . . Careful consideration was given to the design of the test set-up. The most
important feature was the knife edges, which were placed on each side of the
plate. The knife edges were designed to provide a minimum amount of vertical
frictional restraint and a minimum amount of rotational restraint. A detail of a knife
edge is shown in Figure 8.107. A brass wire having a 2 mm radius was inserted
into a machined slot which was filled with graphite grease. The wire was placed
in 5 mm strips separated by 2 mm gaps.
A small amount of play was allowed between the knife edges and the plates as the
knife edges were not butted up against the plate.
As in the tests conducted by Ojalvo and Hull (8.30), an important unknown factor
was the amount of load being transferred to the jig. Botman and Besseling stated that
at the higher loads, transfer of large loads to the knife edges could not be completely
prevented. No indication was given as to what percentage of the total load this might be.
The loading edges were fitted in small slots approximately 2 mm in depth. The
slots were filled with graphite grease. Strain gages used in one of the preliminary
tests indicated that very little horizontal frictional restraint was being exerted on
the loading edge. This was the condition desired by the investigators.”
Guidelines to Modern Plate Buckling Experiments 581

Figure 8.107 Multi-bay panel with special knife edges used by Botman and Besseling (from
[8.23], [8.24] and [8.163])

The many important points on experimental techniques of plate buckling high-


lighted in Davison’s report make it valuable and interesting reading even today.
Also his recommendations for future tests are of interest:
“1. A wide range of b/t ratios should be included. Of special interest is the effect
of initial deviations on the behavior of plates having a high b/t ratio.
2. The initial out-of-flatness should be measured.
3. A complete record of the load-shortening behavior should be kept up to and
well beyond the ultimate load.
4. Out-of-plane deflections should be measured for the entire loading range.
5. Special attention should be given to the boundary conditions at the longitu-
dinal edges. A multi-bay test arrangement appears to be the only acceptable
approach.”
Most of these recommendations are incorporated in modern plate experiments,
except the multi-bay test arrangement demanded in (5), which though
desirable is very rarely employed.
582 Plate Buckling

8.7.2 Noteworthy Details in Some Modern Plate Tests

Two examples of recent plate buckling experiments will now be considered in the
spirit of the guidelines and ideas outlined above.
One example is a simple test rig built at Monash University, Melbourne,
Australia, in the early eighties for buckling tests of stiffened steel plates [8.161], but
used later also for unstiffened plates with deliberately imposed initial imperfections
[8.67]. The test rig (shown in Figure 8.108 with a one-stiffener panel) is essentially
similar in concept to the classical US Bureau of Standards test apparatus used in the
thirties, discussed in Sub-section 8.2.1 (see Figure 8.14 or [8.6]). It consists mainly
of two U-channel cantilevers, each welded to a bottom plate, both being bolted to
a base plate. To each of the U-channels two specially machined profiles (shown
in Figure 8.109) are bolted, which provide the “knife edges” for the longitudinal
unloaded ends of the plate. As can be seen, the test setup is for plates of large aspect
ratio, and panels of effective length 950 mm and width 350 mm, with one, two
or three welded outstand stiffeners were tested. Hence the “long plate approach”,
discussed in the beginning of this section, which assumes a negligible influence
of the loaded ends, applies and clamped loaded ends that prevent premature edge
failure are appropriate. Here the loaded edges of the specimens were clamped into
U-shape supports with tightly fitting steel blocks that were forced in place.
All tests were carried out in the deflection-control mode of a 50 ton Baldwin
testing machine, which also traced the load-shortening curves. In addition, a
moiré fringe technique was employed to study the development of the buckling
deformations. The later test series (of [8.67]) on imperfect unstiffened plates also
employed moiré patterns for most of the tests, and strain gages for the remainder.
The later tests also included some plates with clamped unloaded edges.
Simplicity was the main attribute of this test rig. The noteworthy details are
therefore the simple knife edges of the unloaded ends (Figure 8.109), which
represent a simulation of “simply supported” boundaries. They resemble the
rounded knife edges of Walker (Figure 8.18a or [8.41]) and those of Hoff, Boley
and Coan [8.53], or the knife edges of Stein (Figure 8.41 or [8.27]) and those
of Yamaki (Figure 8.43c or [8.39]), as well as others. These vertical knife edge
clamps are supposed to provide free rotation and zero in-plane restraint, which
they, however, only approximate. For very thin plates, free rotation is closely
approached, but in-plane lateral motion of the specimen as still partially restrained
by friction. These restraints can be alleviated to some extent by a lubricant or by
Teflon contact surfaces. But the conflict of either insufficient out-of-plane support
or some restraint caused by friction (already pointed out earlier in this chapter)
remains. Furthermore, load dissipation via friction may not be entirely negligible
and its measurement is therefore essential for accurate results.
It is of interest to notice, that though many more sophisticated and accurate
plate buckling test setups have been employed for decades, simple cost effective
test rigs with knife edges (like Figure 8.108) can and still are used extensively.
With appropriate sensors measuring the effect of the friction restraints on the
results, these results could today be corrected on-line, perhaps also by computer
Guidelines to Modern Plate Buckling Experiments 583

Figure 8.108 Monash University, Melbourne, test rig for buckling of stiffened plates
(from [8.161])

Figure 8.109 Monash University, Melbourne, test rig for stiffened plates detail of knife
edges, not to scale (courtesy of Prof. N.W. Murray)
584 Plate Buckling

controlled adjustment of the loading. This would transform the simple test rig into
an accurate test setup. Such on-line uses of measured data for automatic correction
of results and for control of the loading and the response of the boundary conditions
have recently been predicted as the trend of future buckling experiments, also for
more complicated structural elements (see for example [8.162]), a view shared by
the authors.
The second example is a more sophisticated modern test rig built at the Institut
für Leichtbau, RWTH Aachen, Germany, in the early nineties for buckling of
composite plates ([8.163] and [8.164]). Here the aim was to design a small scale
test setup, with well defined boundary conditions, that would remain so during the
whole test. The solutions to the problems of friction, insufficient adjustability of
different boundary conditions and non-uniform loading, arrived at in many earlier
tests in the literature, were carefully studied and appropriately applied. This resulted
in a somewhat elaborate but accurate test rig (Figure 8.110).
The test rig, for plates up to 400 ð 200 ð 2 mm, is set in a rigid square 1000 ð
1000 ð 100 mm loading frame and is of modular design, providing different
boundary conditions by change of appropriate edge elements. A bolted construction
was used, rather than a welded one, since this eliminated welding distortions
and therefore assured greater precision. In view of the relatively low loads, the
accompanying lower rigidity of the bolted rig could be disregarded.
As shown in Figure 8.110, two vertical cantilevers (1), which accommodate the
inserts for longitudinal support of the plate, are bolted to a base plate (2). The lower
cross beam (3), in which the lower transverse supports of the plate are placed, is
attached to the base plate, whereas the upper cross beam (4) is fixed to an adaptor
beam (5) whose vertical movement is guided by two axial bearings (6). A load cell
(7) is screwed into the adaptor and the load is applied to it by a 12 mm diameter
threaded rod (8) via a ball end (9). The loading system consists of the threaded
rod (8), which is loaded by tightening a nut (10) that presses against a lever (11)
connected to the square loading frame (12). Alternatively a wing nut (13) can
apply the load via the lever. However, since it was found that the accompanying
rotation of the beam induced a swaying motion of the threaded rod that resulted in
an inclination of the load, this alternative loading with the wing nut was not used
in the tests.
The loaded edges (Figure 8.111) represent a noteworthy detail. They represent
a variation on the theme of ball or roller bearings, as employed in many earlier
test setups discussed in this chapter (see for example Figures 8.17, 8.19, 8.20
or 8.45), into which the concept of a pressurized loading tube (developed in
[8.63] and [8.64], see for example Figure 8.47) has been incorporated. As seen
in Figure 8.111b, the cross beam of the loaded edges consists of a bearing block
(1) that holds half a ball bearing (2). A half shaft (3) rests on the ball bearing, and
it can be locked by two half-shaft rotation locks (8) and their screws (9). The test
plate (15) is supported by a guide rail, with either a V-slot (12) or a rectangular
thin slot (14), which rides between a pair of jaws (11) and roller bearings (10),
that are fixed in position by a guide pin (7) and a set screw (6).
Guidelines to Modern Plate Buckling Experiments 585

(b)

Figure 8.110 IfL, RWTH Aachen test rig for buckling of composite plates (from [8.163]):
(a) view of test rig with plate in position, (b) schematic diagram: (1) vertical
cantilevers, (2) base plate, (3) lower cross beam, (4) upper cross beam, (5)
adaptor beam, (6) axial bearings, (7) load cell, (8) threaded rod, (9) ball end,
(10) loading nut, (11) loading lever, (12) loading frame, (13) wing nut

Three types of boundary conditions of the loaded transverse edges are therefore
possible:

(A) Simple supports with constant rotation along the entire width of the plate,
here the plate is bonded or clamped in the rectangular slot of the guide rail
(14) and the half shaft (3) is free to rotate.
586 Plate Buckling

(b)

Figure 8.111 IfL, RWTH Aachen test rig for buckling of plates loaded edge (from [8.164]):
(a) view of bearing block with loading tube, (b) details of bearing: (1) bearing
block, (2) ball bearing, (3) half-shaft, (4) loading tube, (5) pressurizing fluid, (6)
set screw, (7) guide pin, (8) half-shaft rotation lock, (9) lock screw, (10) roller
bearing, (11) bearing jaw, (12) guide rail with V-slot, (13) pivot segment, (14)
guide rail with slot (for clamped boundary), (15) test plate
Guidelines to Modern Plate Buckling Experiments 587

(B) Simple supports with variable rotation along the width of the plate. Here the
half-shaft is fixed with the rotation locks (8) and the plate is supported by 10
pivot segments (13) in the guide rail (12), that can each rotate independently.
The inner surfaces of the V-slot are polished and covered with Teflon strips
to reduce the friction between it and the pivots.
(C) Clamped supports, where the half-shaft is again locked, as in (B), and the
plate is bonded or clamped in the rectangular slot of the guide rail (14) as
in (A).
All three boundary conditions are well defined.
The loading system aims at applying a uniform line load on the short transverse
ends of the plate. Following the concept of [8.63], this is achieved by water filled
tubes (4), acting as continuous springs. The tubes are commercially available PVC
tubes, of outer diameter 16 mm, which have been shaped to their rectangular
form at 80° C. Instead of water, the pressurized fluid could have also been oil or
glycerin, if less or more stiffness were required. The upper and lower tubes are
not connected, each distributing the load uniformly by adapting its pressure to the
local conditions, which may differ slightly. The load is therefore transmitted from
the bearing blocks via the pressurized tubes to the guide rails, (12) or (14), which
ride between roller bearings to minimize friction, and then to the plate.
For simple supports, the longitudinal edges were originally two round Teflon
rods pressing on the plate (see Figure 8.112b). Metal plates could be supported

Figure 8.112 IfL, RWTH Aachen test rig for buckling of plates unloaded edge, simple
supports (from [8.163]): (a) new arrangement of round, slotted Teflon segments
riding in a vertical guide channel, (b) original edge with pair of Teflon rods
588 Plate Buckling

directly between these round rods, whereas for carbon fiber plates additional Teflon
strips were put between their rougher surfaces and the Teflon rods. However,
difficulties in controlling the gap between the rods lead to poor reproducibility of
the test results. Hence a new longitudinal support was developed, which consists
of eight round Teflon segments, with milled slots for the plate, riding in a vertical
guide channel (see Figure 8.112a). A Teflon foil is introduced between the Teflon
segments and the guide channel to further reduce friction. This modified support
was found to be more accurate and repeatable.
For clamped supports, the unloaded edges of the plate slide in a slotted guide
bar, the friction being reduced by a Teflon foil inserted between them.
To conclude, it should be noted that the two examples, discussed in detail,
are representatives of many others, including some of much larger scale (see for
example Figure 8.20 or [8.55]). They typify two groups in modern plate buckling
experiments: one with the emphasis on mechanical simplicity for cost effective-
ness, and one with the emphasis on optimization of the mechanical details at the
expense of some complexity for the sake of well-defined boundary conditions and
precision. The general future trend of on-line correction of data and loading, based
on feedback from real time measurements, will apply to both groups.

8.7.3 Imperial College London High Stiffness Test Machine

As has been pointed out in the discussion of many buckling test setups in this and
other chapters, high stiffness is an essential character of these test rigs. A good
example of a recent plate testing machine designed to have the high stiffness neces-
sary for postbuckling investigations, is the Imperial College London 250 ton panel
testing facility, shown in Figures 8.113 and 8.114. It has recently been employed
in an extensive study of the buckling and postbuckling behavior of flat stiffened
graphite-epoxy composite (CFC) panels (see [8.165] and [8.166]).
The machine was designed as a special purpose research facility to investigate
postbuckling of fairly large panels. It was to be manufactured at a small fraction
of the cost of commercially available machines, using low cost industrial parts.
Indeed the total cost was less than $50 000.
The machine is rather hefty and can be used for testing panels up to 1.5 m
long, 1 m wide and 0.5 m deep. The maximum travel of the lower, loading platen
is 150 mm. Panels of different length are accommodated by repositioning the
intermediate cross beam, (13) in Figure 8.114, as for example for the smaller
I-stiffened CFC panels (865 mm ð 610 mm) of the recent test program, shown in
Figure 8.113. For full length, 1.5 m panels, the cross beam is removed completely.
This method of reconfiguring the machine is inconvenient, but it is commensurate
with low cost and the high stiffness desired.
The stiffness of the machine is 170 tons/mm [8.167], which is close to that
of other large rigid test machines. Since the columns of the machine are very
massive, this stiffness is virtually independent of the panel length. The stiffness of
the Imperial College 250 ton testing machine is actually similar to that of specially
rigid machines, commercially available, of the same capacity. For example, a recent
Guidelines to Modern Plate Buckling Experiments 589

Figure 8.113 Imperial College London, Department of Aeronautics, 250 ton panel testing
machine (courtesy of K.A. Stevens)

heavy four-column 250 ton MTS Model 311.41 Load Unit has a stiffness of K D
411 ton/mm (see [8.168]). This is about twice the stiffness of the IC machine, but
the spring rate of the MTS unit is for a 25 percent narrower and 15 percent shorter
test opening, which makes it inherently stiffer. Smaller capacity commercially
available universal testing machines have usually lower spring rates, whereas very
large capacity ones are stiffer. For example, a recent two-column 50 ton MTS
Model 318.50 Load Unit has a stiffness of K D 43.8 ton/mm (see [8.168]); or an
older two-column 27 ton Balwin Hamilton Lima testing machine (of the sixties)
had a stiffness of K D 13.2 ton/mm (see [9.81, Volume 2]). On the other hand, a
recent very large capacity four-column 1000 ton MTS Model 311.71 Load Unit
has a stiffness of K D 1, 305 ton/mm.
Figure 8.114 shows the construction and operation of the testing machine in more
detail and conveys an impression of its rigidity. The platens (12) are approximately
1350 ð 625 mm and the load is measured by four clevis type load cells (10). The
590
Plate Buckling

Figure 8.114 Imperial College London, Department of Aeronautics, 250 ton panel testing machine operating diagram (courtesy of K.A. Stevens)
References 591

setup also has rate of feed and platen displacement gages for precise indication of
these parameters, that determine the actual loading.
As described by the designer of the test machine, Peter Gasson, the displacement
drive operates as follows:
“1. A four h.p. squirrel cage motor (1) drives a variable displacement hydraulic
pump. The rotational speed of the variable displacement pump (2) may be
controlled remotely between 1500 and 35 r.p.m. at full torque but may be
driven at lower speeds subject to a reduced torque and efficiency.
2. The output shaft of the variable displacement pump drives through a 7.5:1
single worm reduction gear box with twin output shafts. In the latest research,
requiring a very low loading rate, the drive speeds were reduced further by
changing the gear sets to give a 20:1 reduction. This reduced the loading rate
to < 0.1 mm/min at full load, a very low one indeed.
3. The output from the single worm reduction gear (4) is fed into two separate
single worm reduction gears of ratio 25:1.
4. The two separate output shafts of the 25:1 single worm gear boxes (6) drive
two separate single worm screw jacks, (8) of ratio 36:1, with 20 mm pitch
screws. These screws drive the loading platen directly.”
The observation platform (above the motors and pumps) facilitates the instal-
lation and servicing of specimens. The instrumentation shown in Figure 8.113
includes an in-situ ultrasonic scanning facility (its telescopic arm can be seen to
the left of the test panel), an acoustic emission sensing apparatus as well as a data
logger and P.C.s. Moiré interferometry is also employed.
The design focused, however, on high stiffness as the main characteristic of
this setup, for providing a practically rigid frame for loading and displacements
reference. The good performance of this testing machine indicates the feasibility
of relatively low cost “build it yourself” special purpose test facilities, provided
the structures laboratory has the support of well equipped workshops.

References
8.1 Timoshenko, S., History of Strength of Materials, McGraw-Hill Book Company,
New York/Toronto/London, 1953, 156 162.
8.2 Clark, E., The Britannia and Conway Tubular Bridges, With General Inquiries on
Beams and on the Properties of Materials Used in Construction, 1, Day and Son,
Lincoln’s Inn Fields, and John Weale, High Holborn, London, 1850.
8.3 Kollbrunner, C.F., and Meister, M., Ausbeulen, Theorie und Berechnung von
Blechen, Springer-Verlag, Berlin/Göttingen/Heidelberg, 1958, 285 287.
8.4 Walker, A.C., A Brief Review of Plate Buckling Research, Behavior of Thin-Walled
Structures, J. Rhodes, and J. Spence, eds., Elsevier Applied Science Publishers,
London, 1984, 375 398.
8.5 Rhodes, J., and Harvey, J.M., Examination of Plate Post-Buckling Behavior, Journal
of the Engineering Mechanics Division, ASCE, 103, (EM3), 1977, 461 478.
8.6 Schuman, L., and Back, G., Strength of Rectangular Plates Under Edge Compres-
sion, NACA Technical Report No. 356, 1930.
592 Plate Buckling

8.7 von Kármán, T., Sechler, E.E. and Donnell, L.H., The Strength of Thin Plates in
Compression, ASME Applied Mechanics Transactions, 54, 1932, 53 57.
8.8 Sechler, E.E., The Ultimate Strength of Thin Flat Sheets in Compression, GALCIT
Publication 27, Guggenheim Aeronautics Laboratory, California Institute of Tech-
nology, Pasadena, 1933.
8.9 Cox, H.L., Buckling of Thin Plates in Compression, Aeronautical Research
Committee, R. & M. No. 1554, 1933.
8.10 Lahde-Wagner, Versuche zur Ermittlung der mittragenden Breite von verbeulten
Blechen, Luftfahrt-Forschung, 13, 1936, 214 223.
8.11 Sechler, E.E., Stress Distribution in Stiffened Panels under Compression, Journal
of the Aeronautical Sciences, 4, (8), 1937, 320 323.
8.12 Maguerre, K., Die mittragende Breite der gedrückten Platte, Luftfahrt-Forschung,
14, (3), 1937, 121 128 (Translated as The Apparent Width of the Plate in Compres-
sion NACA TM 833, 1937).
8.13 Dickinson, H.B., and Fischel, J.R., Measurement of Stiffener Stresses and Effective
Widths in Stiffened Panels, Journal of the Aeronautical Sciences, 6, (6), 1939,
249 254.
8.14 Hoff, N.J., Instability of Monocoque Structures in Pure Bending, Journal of the
Royal Aeronautical Society, 42, 1938, 291 346.
8.15 Sechler, E.E. and Dunn, L.G., Airplane Structural Analysis and Design, John Wiley
& Sons, New York, 1942, 201 234.
8.16 Kollbrunner, C.F., Das Ausbeulen der auf einseitigen, gleichmässig verteilten
Druck beanspruchten Platten im elastischen und plastischen Bereich (Versuchs-
bericht) Mitt. Inst. Baustatik, E.T.H. Zürich, H. No. 17, 1946, Lehmann, Zürich.
8.17 Winter, G., Strength of Thin Steel Compression Flanges, Trans. ASCE, 112, 1947,
527 554.
8.18 Stussi, F., Kollbrunner, C.F., and Walt, M., Versuchsbericht über das Ausbeulen
der auf einseitigen, gleichmässig und undgleichmässig verteilten Druck
beanspruchten Platten aus Avional M, hart vergütet, Mitt. Inst. Baustatik, E.T.H.,
Zürich, H. No. 25, 1951, Lehmann, Zürich.
8.19 Faulkner, D., A Review of Effective Plating for Use in the Analysis of Stiffened
Plating in Bending and Compression, Journal of Ship Research, 19 (1), 1975, 1 17.
8.20 Conley, W.F., Becker, L.A., and Allnutt, R.B., Buckling and Ultimate Strength of
Plating Loaded in Edge Compression, Progress Report No. 2 Unstiffened Panels,
U.S. Navy David Taylor Model Basin, DTMB Report 1682, May 1963.
8.21 Becker, H., Goldman, R., and Pozerycki, J., Compressive Strength of Ship Hull
Girders, Part I, Unstiffened Plates, Ship Structures Committee (USA) Report SSC-
217, 1970.
8.22 Fukumoto, Y., and Itoh, Y., Basic Compressive Strength of Steel Plates from
Test Data, Proc. of Japan Society of Civil Engineers (JSCE) (344/I-1) (Structural
Eng./Earthquake Eng.) April 1984, 129 139.
8.23 Botman, M., and Besseling, J.F., The Effective Width in the Plastic Range of Flat
Plates under Compression, NLL Report S 445, Amsterdam, 1954.
8.24 Besseling, J.F., De Experimentele Bepaling van de Meedragende Breedte van
Vlakke Platen in het Elastische en het Plastische Gebied, NLL Report S 414,
Amsterdam, 1953.
8.25 Supple, W.J., and Chilver, A.H., Elastic Post-Buckling of Compressed Rectangular
Flat Plates, Thin-Walled Structures, A.H. Chilver, ed., John Wiley & Sons, New
York, 1967, 136 152.
References 593

8.26 Stein, M., The Phenomenon of Change in Buckle Pattern in Elastic Structures,
NASA Technical Report R-39, 1959.
8.27 Stein, M., Loads and Deformations of Buckled Rectangular Plates, NASA Technical
Report R-40, 1959.
8.28 Uemura, M., and Byon, O-I., Secondary Buckling of a Flat Plate Under Uniaxial
Compression, Part 1: Theoretical Analysis of Simply Supported Flat Plate, Inter-
national Journal of Non-Linear Mechanics, 12, 1977, 355 370.
8.29 Nakamura, T., and Uetami, K., The Secondary Buckling and Post-Secondary-
Buckling Behaviours of Rectangular Plates, International Journal of Mechanical
Sciences, 21, 1979, 265 286.
8.30 Ojalvo, M., and Hull, F.H., Effective Width of Thin Rectangular Plates, Journal of
the Engineering Mechanics Div., Proc. ASCE, 84, (EM3), 1958, 1718 1 20.
8.31 Stroebel, G.J., and Warner, W.H., Stability and Secondary Bifurcation for von
Kármán Plates, Journal of Elasticity, 3, 1973, 185 202.
8.32 Supple, W.J., Changes of Wave Form of Plates in the Post-Buckling Range, Inter-
national Journal of Solids and Structures, 6, 1970, 1243 1258.
8.33 Domburian, E.M., Smith, C.V., and Carlson, R.L., A Perturbation Solution to a
Plate Postbuckling Problem, Journal of Non-Linear Mechanics, 11, 1976, 49 58.
8.34 Uemura, M., and Byon, O-I., Secondary Buckling of Flat Plate Under Uniaxial
Compression, Part 2: Analysis of Clamped Plate by F.E.M. and Comparison with
Experiments, International Journal of Non-Linear Mechanics, 13, 1979, 1 14.
8.35 Boucif, M., Wesfreid, J.E. and Guyon, E., Role of Boundary Conditions on Mode
Selection in a Buckling Instability, Journal de Physique, Lettres, 45, 1984,
L413 L418.
8.36 Clement, M., Guyon, E., and Wesfreid, J.E., Multiplicite de modes de deformation
d’une plaque sans compression. Experience, C.R. Acad. Sci., Paris, Series II, 293,
1981, 87 89.
8.37 Maaskant, R. and Roorda, J., Mode Jumping in Biaxially Compressed Plates, Inter-
national Journal of Solids and Structures, 29, (10), 1992, 1209 1219.
8.38 Yamaki, N., Postbuckling Behavior of Rectangular Plates with Small Initial Curva-
ture Loaded in Edge Compression, Journal of Applied Mechanics, 26, Trans. ASME
81, Series E, 1959, 407 414.
8.39 Yamaki, N., Experiments on the Postbuckling Behavior of Square Plates Loaded in
Edge Compression, Journal of Applied Mechanics, 28, Trans. ASME, 83, Series E,
1961, 238 244.
8.40 Rhodes, J., Harvey, J.M., and Fok, W.C., The Load-Carrying Capacity of Initially
Imperfect Eccentrically Loaded Plates, International Journal of Mechanical Scien-
ces, 17, 1975, 161 175.
8.41 Walker, A.C., Flat Rectangular Plates Subjected to a Linearly-Varying Edge
Compressive Loading, Thin-Walled Structures, A.H. Chilver, ed., Chatto and
Windus, London, 1967, 208 247.
8.42 Frieze, P.A., Dowling, P.J., and Hobbs, R.E., Ultimate Load Behavior of Plates in
Compression, Steel Plated Structures, P.J. Dowling, J.E. Harding, and P.A. Frieze,
eds., Crosby Lockwood Staples, London, 1977, 24 50.
8.43 Bradfield, C.D., Tests on Plates Loaded in In-Plane Compression, Journal of
Constructional Steel Research (UK), 1, (1), 1980, 27 37.
8.44 Bradfield, C.D., Tests on Single Plates Under In-Plane Compression with Controlled
Residual Stresses and Initial Out-of-Flatness, University of Cambridge, Dept. of
Engineering, Report CUED/D-Struct/TR. 78, 1979.
594 Plate Buckling

8.45 Dwight, J.B., and Ractliffe, A.T., The Strength of Thin Plates in Compression,
Thin-Walled Steel Structures, K.C. Rockey, and H.V. Hill, eds, Crosby Lockwood,
London, 1969, 3 34.
8.46 Dwight, J.B., and Moxham, K.E., Welded Steel Plates in Compression, The Struc-
tural Engineer, Journal of the ISE (UK), 47, (2), 1969, 49 66.
8.47 Crisfield, M.A., Full-Range Analysis of Steel Plates and Stiffened Panels Under
Uniaxial Compression, Proceedings of the Institution of Civil Engineers, 59, 1975,
595 624.
8.48 Harding, J.E., Hobbs, R.E., and Neal, B.G., Ultimate Load Behavior of Plates
Under Combined Direct and Shear In-Plane Loading, Steel Plated Structures,
P.J. Dowling, J.E. Harding, and P.A. Frieze, eds., Crosby Lockwood Staples,
London, 1977, 369 403.
8.49 Little, G.H., Rapid Analysis of Plate Collapse by Live Energy Minimization, Inter-
national Journal of Mechanical Sciences, 19, 1977, 725 744.
8.50 Moxham, K.E., and Bradfield, C.D., The Strength of Welded Steel Plates Under
In-Plane Compression, University of Cambridge, Dept. of Engineering, Report
CUED/C-Struct/TR. 65, 1977.
8.51 Korol, R.M., and Thimmhardy, E.G., Geometric Imperfections of Steel Box Girder
in Bridges in Canada, Stability of Metal Structures, Proceedings of 3rd Interna-
tional Colloquium, George Winter Memorial Session, SSRC, Toronto, Canada, 1983,
231 251.
8.52 Ueda, Y., Yao, T., and Ominami, R., Ultimate Strength of Square Plates Subjected
to Compression (1st Report) Effects of Initial Deflection and Welding Residual
Stresses, Journal Society of Naval Architects of Japan, 137, 1975, 210 221.
8.53 Hoff, N.J., Boley, B.A., and Coan, J.M., The Development of a Technique for
Testing Stiff Panels in Edgewise Compression, Proceedings of the Society for Exper-
imental Stress Analysis, 5, (2), 1948, 14 24.
8.54 Schlack, A.L., Experimental Critical Loads for Perforated Square Plates, Experi-
mental Mechanics, 8, (2), 1968, 68 74.
8.55 Fischer, M., and Harre, W., Ermittlung der Traglastkuren von einachsig gedrückten
Rechteckplatten aus Baustahl der Seitenverhältnisse ˛  1 mit Hilfe von Versuchen,
Der Stahlbau, 47, (7), 1978, 199 204, and (8), 239 247.
8.56 Duffy, D.J., and Allnutt, R.B., Buckling and Ultimate Strength of Plating Loaded
in Edge Compression, Progress Report 1 6061-T6 Aluminum Plates, U.S. Navy
David Taylor Model Basin, DTMB Report 1419, April 1960.
8.57 Moxham, K.E., Buckling Tests on Individual Welded Steel Plates in Compression,
University of Cambridge, Department of Engineering, Report CUED/C-Struct/TR.
3, 1971.
8.58 Mofflin, D.S., and Dwight, J.B., Buckling of Aluminum Plates in Compression,
Behavior of Thin-Walled Structures, J. Rhodes, and J. Spence, eds., Elsevier
Applied Science Publishers, London, 1984, 399 427.
8.59 Mofflin, D.S., and Dwight, J.B., Tests on Individual Aluminum Plates Under In-
Plane Compression, University of Cambridge, Department of Engineering, Report
CUED/D-Struct/TR. 100, 1983.
8.60 Bradfield, C.D., Analysis of Measured Distortions in Steel Box Girder Bridges,
University of Cambridge, Department of Engineering, Report CUED/C-Struct/TR.
42, 1974.
References 595

8.61 Minguez, J.M., An Experimental Investigation of How Accurate Simply Supported


Boundary Conditions can be Achieved in Compression Testing of Panels, Experi-
mental Mechanics, 26, 1986, 238 244.
8.62 Sharman, P.W., and Humpherson, J., An Experimental and Theoretical Investi-
gation of Simply-Supported Thin Plates Subjected to Lateral Load and Uniaxial
Compression, The Aeronautical Journal (of the Royal Aeronautical Society) 72,
(689), 1968, 431 436.
8.63 Rosen, A., and Libai, A., Stability and Behavior of an Annular Plate Under Uniform
Compression, Experimental Mechanics, 16, 1976, 461 467.
8.64 Zaal, K., Buckling and Postbuckling of a Square Plate Subjected to Uniform Edge
Loads An Experimental Investigation, ASL-138 Technion Israel Inst. of Tech-
nology, Dept. of Aeronautical Engineering, Feb. 1988.
8.65 Majumdar, S., Buckling of a Thin Annular Plate under Uniform Compression, AIAA
Journal, 9, (9), 1971, 1701 1707.
8.66 Rogers, N.A., Compression Tests on Plain Flat Outstands, University of Cambridge,
Dept. of Engineering, Report CUED/C-Struct/TR. 52, 1976.
8.67 Fok, C.D. and Murray, N.W., The Effect of Initial Imperfections on the Elastic
Behaviour of Isolated Thin Steel Plates with In-Plane Compression, Aspects of the
Analysis of Plate Structures, Volume in Honour of W.H. Wittrick, Oxford University
Press, Oxford 1985, 225 249.
8.68 Denston, R.J. and White J.D., An Electrical Demountable Extensometer, University
of Cambridge, Dept. of Engineering, Report CUED/C-Strut/TR. 61, 1977.
8.69 Horton, W.H., Cundari, F.L. and Johnson, R.W., A Review of the Applicability of
the Southwell Plot to the Interpretation of Test Data Obtained from Stability Studies
of Elastic Column and Plate Structures, SUDAAR 296, Department of Aeronautics
and Astronautics, Stanford University, Dec. 1966.
8.70 Walker, A.C., Local Instability in Plates and Channel Struts, Journal of the Struc-
tural Division, ASCE, 92, (ST 3), June 1966, 38 55.
8.71 Queinec, A., Thermal Buckling of Centrally Heated Circular Plates, SUDAAR 106,
Department of Aeronautics and Astronautics, Stanford University, June 1961.
8.72 Fernandes-Sintes, J., Horton, W.H. and Hoff, N.J., Thermal Buckling of Annular
Plates, SUDAAR 143, Department of Aeronautics and Astronautics, Stanford
University, Dec. 1962.
8.73 Majumdar, S., Buckling of Thin Annular Plates due to Radial Compressive Loads,
Aeronautical Engineer Thesis, California Institute of Technology, Pasadena, Cali-
fornia, 1968.
8.74 Rosen, A. and Libai, A., Stability, Behaviour and Vibrations of an Annular Plate
Under Uniform Compression, TAE Report 229, Dept. of Aeronautical Engineering,
Technion Israel Institute of Technology, Haifa, Israel, Oct. 1974.
8.75 Carlson, R.L. and Datta, P.K., On the Analysis of Plate Stability Experiments,
Proceedings of the 14th South American Conference of Structural Engineering and
4th Pan American Symposium of Structures, S.A.E.M., Buenos Aires, Argentina,
October 1970.
8.76 Carlson, R.L., Zielsdorff, G.F. and Harrison, J.C., Buckling in Thin Cracked Sheets,
in Proceedings of the Air Force Conference on Fatigue and Fracture of Aircraft
Structures and Materials, AFFDL TR 70 144, 1970, 193 205.
596 Plate Buckling

8.77 Zielsdorff, G.F. and Carlson, R.L., On the Buckling of Thin Tensioned Sheets with
Cracks and Slots, Engineering Fracture Mechanics, 4, 1972, 939 950.
8.78 Datta, P.K. and Carlson, R.L., Buckling and Vibration of a Thin Tensioned Sheet
with an Elliptical Hole, Experimental Mechanics, 13, July 1973, 280 286.
8.79 Datta, P.K., Static Stability Behaviour of Plate Elements with Non-Uniform In-
Plane Stress Distribution, Journal of Mechanical Engineering Science, IMechE, 21,
(5), 1979, 363 365.
8.80 Souza, M.A., Fok, W.C. and Walker, A.C., Review of Experimental Techniques for
Thin-Walled Structures Liable to Buckling Part II Stable Buckling, Experimental
Techniques, 7, October 1983, 36 39.
8.81 Fok, W.C. and Yuen, M.F., Modified Pivotal Plot for Critical Load Calculation of
a Rectangular Plate Under Edge Compression, Journal of Mechanical Engineering
Science, IMechE, 23, (4), 1981, 167 170.
8.82 Chailleux, A., Hans, Y. and Verchery, G., Experimental Study of the Buckling of
Laminated Composite Columns and Plates, International Journal of Mechanical
Sciences, 17, 1975, 489 498.
8.83 Segall, A. and Springer, G.S., A Dynamical Method for Measuring the Critical
Loads of Elastic Flat Plates, Experimental Mechanics, 26, Dec. 1986, 354 359.
8.84 Fok, W.C., Evaluation of Experimental Data of Plate Buckling, ASCE Journal of
Engineering Mechanics, 110, (4), April 1984, 577 588.
8.85 Weller, T., Abramovich, H. and Yaffe, R., Dynamic Buckling of Beams and Plates
Subjected to Axial Impact, Computers and Structures, 32, (3/4), 1989, 835 851.
8.86 Abramovich, H., Weller, T. and Yaffe, R., Application of a Modified Donnell Tech-
nique for the Determination of Critical Loads of Imperfect Plates, Computers and
Structures, 37, (4), 1990, 463 469.
8.87 Fok, W.C., Personal communication to J. Singer, 30 August 1989.
8.88 Vann, W.P. and Sehested, J., Experimental Techniques for Plate Buckling, Proc. of
the Second Speciality Conference on Cold Formed Steel Structures, St. Louis, MO,
1973, 83 105.
8.89 Venkataramaiah, K.R. and Roorda, J., Analysis of Local Plate Buckling Experi-
mental Data, in Proc. Sixth International Speciality Conference on Cold-Formed
Steel Structures, St. Louis, MO, 1982, 45 74.
8.90 Schuette, E.H., Charts for the Minimum-Weight Design of 24S-T Aluminum-Alloy
Flat Compression Panels with Longitudinal Z-Section Stiffeners, NACA ARR
L5F15, 1945 (Wartime Report).
8.91 Heimerl, G.J. and Roy, J.A., Column and Plate Compressive Strength of Aircraft
Structural Materials, 17S-T Aluminum-Alloy Sheet, NACA ARR L5F08, June 1945
(Wartime Report).
8.92 Gallaher, G.L., Plate Compressive Strength of FS-1h Magnesium-Alloy Sheet and
a Maximum-Strength Formula for Magnesium-Alloy and Aluminum-Alloy Formed
Sections, NACA TN 1714, October 1948.
8.93 Timoshenko, S., Über die Stabilität versteifter Platten, Der Eisenbau, 12, 1921,
147 163.
8.94 Stein, M. and Neff, J., Buckling Stress of Simply Supported Rectangular Flat Plates
in Shear, NACA TN 1222, 1947.
8.95 Budiansky, B. and Connor, R.W., Buckling Stresses of Clamped Rectangular Flat
Plates in Shear, NACA TN 1559, 1948.
8.96 Wilson, J.M., On Specifications for Strength of Iron Bridges, Transactions of the
American Society of Civil Engineers, 15, Part I, 1886, 401 403, 489 490.
References 597

8.97 Wagner, H., Ebene Blechwandträger mit sehr dünnem Stegblech, Zeitschrift für
Flugtechnik und Motorluftschiffahrt, 20, (8,9,10,11,12), 1929, translated as Flat
Sheet Metal Girder with Very Thin Metal Web, NACA TM Nos. 604 606, 1931.
8.98 Kuhn, P., Peterson, J.P., and Levin, L.R., A Summary of Diagonal Tension,
Part I Methods of Analysis, NACA TN 2661, 1952.
8.99 Kuhn, P., Peterson, J.P., and Levin, L.R., A Summary of Diagonal Tension,
Part II Experimental Evidence, NACA TN 2662, 1952.
8.100 Kuhn, P., Stresses in Aircraft and Shell Structures, McGraw-Hill, New York, 1956.
8.101 Hertel, H., Leichtbau, Springer, Berlin/New York, 1960.
8.102 Rockey, K.C., The Design of Web Plates for Plate and Box Girders A State of the
Art Report, in Steel Plated Structures, P.J. Dowling, J.E. Harding and P.E. Frieze,
eds., Crosby Lockwood Staples, London, 1977, 459 485.
8.103 Evans, H.R., Longitudinally and Transversely Reinforced Plate Girders, in Plated
Structures, Stability and Strength, R. Narayanan, ed. Applied Science Publishers,
London and New York, 1983, 1 37.
8.104 Basler, K., Strength of Plate Girders in Shear, Journal of the Structural Division,
Proc. ASCE, 87, (ST7), 1961, 151 180.
8.105 Rockey, K.C., The Design of Intermediate Vertical Stiffeners on Web Plates Sub-
jected to Shear, The Aeronautical Quarterly, 7, Nov. 1956, 275 296.
8.106 Rockey, K.C. and Skaloud, M., The Ultimate Load Behaviour of Plate Girders
Loaded in Shear, Structural Engineer, 50, (1), 1972, 29 47.
8.107 Rockey, K.C., Evans, H.R. and Porter, D.M., A Design Method for Predicting the
Collapse Behaviour of Plate Girders, Proceedings of the Institution of Civil Engi-
neers, 65, March 1978, 85 112.
8.108 Vilnay, O., The Behaviour of a Web Plate Loaded in Shear, Thin-Walled Structures,
10, 1990, 161 174.
8.109 Brown, K.E.P. and Evans, H.R., Theoretical and Experimental Investigation of the
Collapse Behaviour of Transversely Stiffened Aluminium Alloy Plate Girders, Thin-
Walled Structures, 18, 1994, 225 246.
8.110 Farley, G.L. and Baker, D.J., In-Plane Shear Test of Thin Panels, Experimental
Mechanics, 23, March 1983, 81 88.
8.111 Roberts, T.M. and Ghomi, S., Hysteretic Characteristics of Unstiffened Plate Shear
Panels, Thin-Walled Structures, 12, 1991, 145 162.
8.112 Bush, H.G. and Weller, T., A Biaxial Method for In-Plane Shear Testing, NASA
TM 74070, April 1978.
8.113 Rouse, M., Postbuckling and Failure Characteristics of Stiffened Graphite-Epoxy
Shear Webs, AIAA Preprint 87 0733, Proc. 28th AIAA SDM Conference,
Monterey, CA., April 1987, 181 193.
8.114 Horst, P. and Kossira, H., Theoretical and Experimental Investigation of Thin-
Walled Aluminum Panels under Cyclic Shear Load, in ESA-SP 89, Proc. of
the International Conference on Spacecraft and Mechanical Testing, Noordwijk,
Holland, Oct. 1988, 79 84.
8.115 Evans, H.R. and Hamoodi, M.J., The Collapse of Welded Aluminum Plate
Girders Experimental Study, Thin-Walled Structures, 5, 1987, 247 275.
8.116 Burt, C.A., Evans, H.R. and Vilnay, O., Further Experimental Studies of the Colla-
pse of Welded Aluminum Plate Girders, Thin-Walled Structures, 8, 1989, 19 39.
8.117 Herzog, M.A.M., Ultimate Static Strength of Plate Girders from Tests, ASCE Jour-
nal of the Structural Division, 100, (ST5), May 1974, 849 864.
598 Plate Buckling

8.118 Evans, H.R., An Appraisal, by Full-Scale Testing of New Design Procedures for
Steel Girders Subjected to Shear and Bending, Proc. Institution of Civil Engineers,
Part 2, 81, June 1986, 175 189.
8.119 Sfintesco, D., Beedle, L.S., Schulz, G.W. and Zandonini, R. eds., Stability of Metal
Structures/A World Review, American Institute of Steel Construction, Chicago, ILL.,
1982.
8.120 Evans, H.R. and Tang, K.H., The Influence of Longitudinal Web Stiffeners Upon
the Collapse Behaviour of Plate Girders, Journal of Constructional Steel Research,
4, 1984, 201 234.
8.121 Ari-Gur, J., Singer, J. and Libai, A., Repeated Buckling Tests of Stiffened Thin
Shear Panels, Israel Journal of Technology, 20, 1982, 220 231.
8.122 Kollet, M., Weller, T., Libai, A. and Singer, J., Durability Under Repeated Buck-
ling of Stiffened Shear Panels, TAE Report 509, Dept. of Aeronautical Engineering,
Technion Israel Institute of Technology, Haifa, Israel, March 1983.
8.123 Libai, A., Weller, T., Kollet, M. and Singer, J., Stiffened Panels Subjected to
Repeated Buckling Durability Studies, TAE Report 545, Dept. of Aeronautical
Engineering, Technion Israel Institute of Technology, Haifa, Israel, July 1984.
8.124 Weller, T., Kollet, M., Libai, A. and Singer, J., Durability Under Repeated Buck-
ling of Stiffened Shear Panels, in ICAS 1984 Proceedings of the 14th Congress of the
International Council of the Aeronautical Sciences, B. Laschka and R. Staufenbiel,
eds., ICAS, Toulouse, France, 1984, 932 942. Also Synoptic in AIAA Journal of
Aircraft, 24, (1), 1987, 6 7.
8.125 Weller, T. and Singer, J., Durability of Stiffened Composite Panels Under Repeated
Buckling, International Journal of Solids and Structures, 26, (9/10), 1990,
1037 1069.
8.126 Johansson, B., SAAB-340 Fuselage Side Panel with Window Cut-Outs, Static Shear
Test, SAAB-SCANIA 340 Test Report 72GTS5313, 16 Sept. 1983.
8.127 Herzog, M.A.M., Web Crippling with Bending and Shear of Thin-Walled Plate
Girders, Journal of Constructional Steel Research, 22, 1992, 87 97.
8.128 Roberts, T.M., Patch Loading on Plate Girders, Chapter 3 in Plated Structures,
Stability and Strength, R. Narayanan, ed., Applied Science Publishers, London and
New York, 1983, 77 102.
8.129 Elgaaly, M., Web Design Under Compressive Edge Loads, AISC Engineering Jour-
nal, Fourth Quarter, 1983, 153 171.
8.130 Tschamper, H., Interaktion zwischen Biegung und konzentrierter rippenloser
Lasteinleitung an schlanken Trägern, Publication No. 89 1, Baustatik und Stahlbau,
ETHZ, Zürich, 1989.
8.131 Elgaaly, M., Salkar, R. and Eash, M., Unstiffened and Stiffened Webs Under
Compressive Edge Loads, in Proceedings Structural Stability Research Council,
1992 Annual Technical Session, SSRC Lehigh University, Bethlehem, PA, April
1992, 279 290.
8.132 Girkmann, K., Die Stabilität der Stegbleche vollwändiger Träger bei Berücksich-
tigung örtlicher Lastangriffe, International Association for Bridge and Structural
Engineering (IABSE), 3rd Congress, Final Report, Berlin, 1936, 610 614.
8.133 Bergfelt, A. and Hövik, J., Shear Failure and Local Web Crippling in Thin-Walled
Plate Girders, Chalmers University of Technology, Göteborg, Institutionen för
Konstruktionsteknik Stal-Och, Träbyggnad, Int. skr S70:11b, 1970.
References 599

8.134 Granholm, C.A., Light Girders, Girders with Slender Flanges and Web, (reports
in Swedish with English summaries by A. Bergfelt), Chalmers University of
Technology, Göteborg, Inst. For Konst. Stal-Och, Träbyggnad, Int. skr S76:14,
1976.
8.135 Roberts, T.M. and Rockey, K.C., A Mechanism Solution for Predicting the Collapse
Loads of Slender Plate Girders when Subjected to In-Plane Patch Loading, Proceed-
ings of the Institution of Civil Engineers, Part 2, 67, 1979, 155 175.
8.136 Roberts, T.M., Slender Plate Girders Subjected to Edge Loading, Proceedings of
the Institution of Civil Engineers, Part 2, 71, 1981, 805 819.
8.137 Czechowski, A. and Brodka, J., Étude de la Résistance Statique des Assemblages
Soudés en Croix de Profilés Creux Rectangulaires, Construction Metallique, 3,
Paris, France, 1977, 17 25.
8.138 Packer, J.A., Web Crippling of Rectangular Hollow Sections, Journal of Structural
Engineering, ASCE, 110, (11), Oct. 1984, 2357 2373.
8.139 Davies, G. and Packer, J.A., Analysis of Web Crippling in a Rectangular Hollow
Section, Proceedings of the Institution of Civil Engineers, Part 2, 83, Dec. 1987,
785 798.
8.140 Santaputra, C., Parks, M.B. and Yu, W.W., Web-Crippling Strength of Cold-
Formed Steel Beams, Journal of Structural Engineering, ASCE, 115, 10, Oct. 1989,
2511 2527.
8.141 Hetrakul, N. and Yu, W.W., Cold-Formed Steel I-Beams Subjected to Combined
Bending and Web Crippling, in Thin-Walled Structures, J. Rhodes, and A.C. Walker,
eds., Granada, London, 1979, 413 426.
8.142 Zhao, X.-L. and Hancock, G.J., T-Joints in Rectangular Hollow Sections Subject to
Combined Actions, Journal of Structural Engineering, ASCE, 117, (8), Aug. 1991,
2258 2277.
8.143 Kutmanova, I., Skaloud, M., Janus, K. and Löwitova, O., Ultimate Load Behaviour
of Longitudinally Stiffened Steel Webs Subject to Partial Edge Loading, in Contact
Loading and Local Effects in Thin-Walled Plated and Shell Structures, Proceedings
IUTAM 1990 Symposium Prague, Academia Prague 1992 (and Springer Verlag,
Berlin), 148 164.
8.144 Raoul, J., Schaller, I. and Theillout, J.N., Tests of Buckling of Panels Subjected to
In-Plane Patch Loading, in Contact Loading and Local Effects in Thin-Walled Plated
and Shell Structures, Proceedings IUTAM 1990 Symposium Prague, Academia
Prague 1992 (and Springer Verlag, Berlin), 173 183.
8.145 Drdacky, M., Non-Stiffened Steel Webs with Flanges Under Patch Loading, in
Contact Loading and Local Effects in Thin-Walled Plated and Shell Structures,
Proceedings IUTAM 1990 Symposium Prague, Academia Prague 1992 (and Springer
Verlag, Berlin), 111 118.
8.146 Shimizu, S., Yoshida, S. and Okuhara, H., An Experimental Study on Patch-Loaded
Web Plates, in Stability of Plate and Shell Structures, Proceedings ECCS Interna-
tional Colloquium, Ghent University, April 1987, P. Dubas and D. Vandepitte, eds.,
1987, 85 94.
8.147 Herzog, M., Die Krüppellast von Blechträger- und Walzprofilstegen, Stahlbau, 55,
(3), 1986, 87 88.
8.148 Smith, C.S., Compressive Strength of Welded Steel Ship Grillages, Trans. Royal
Institute of Naval Architects, 117, 1975, 325 359.
600 Plate Buckling

8.149 Williams, D.G. and Walker, A.C., Explicit Solutions for Plate Buckling Analysis,
Journal of the Engineering Mechanics Div., ASCE, 103 (EM4), Aug. 1977,
549 568.
8.150 Dowling, P.J., Plate Buckling Considerations in the Design of Steel Structures, in
Collapse: The Buckling of Structures in Theory and Practice, J.M.T. Thomson and
G.W. Hunt, eds., Cambridge University Press, 1983, 235 257.
8.151 Narayanan, R. and Shanmugam, N.E., Compressive Strength of Biaxially Loaded
Plates, in Plated Structures, Stability and Strength, R. Narayanan, ed., Applied
Science Publishers, London 1983, 195 219.
8.152 Harding, J.E., The Interaction of Direct and Shear Stresses on Plate Panels,
in Plated Structures, Stability and Strength, R. Narayanan, ed., Applied Science
Publishers, London, 1983, 221 225.
8.153 Valsgård, S., Ultimate Capacity of Plates in Biaxial In-Plane Compression, Det
norske Veritas Report 78 678, 1978, revised 1979.
8.154 Becker, H., Goldman, R. and Pazerycki, J., Compressive Strength of Ship Hull
Girders, Part I, Unstiffened Plates, Ship Structure Committee Report SSC-217 on
Small Hull Girder Model, US Coastguard Headquarters, Washington DC, 1970.
8.155 Becker, H. and Colao, A., Compressive Strength of Ship Hull Girders, Part III,
Theory and Additional Experiments, Ship Structure Committee Report SSC-267,
Washington DC, 1977.
8.156 Becker, H., Instability, Strength of Polyaxially Loaded Plates and Relation to
Design, in Steel Plated Structures, P.J. Dowling, J.E. Harding and P.E. Frieze, eds.,
Crosby Lockwood Staples, London, 1977, 559 580.
8.157 Stonor, R.W.P., Bradfield, C.D., Moxham, K.E. and Dwight, J.B., Tests on Plates
Under Biaxial Compression, University of Cambridge, Department of Engineering,
Report CUED/D-Struct/TR.98, 1983.
8.158 Kim, Y.S. and Hoa, S.V., Effects of Load Combination on Biaxial Buckling of
Laminated Composite Rectangular Plates, in Composites Design, Proceedings of the
9th International Conference on Composite Materials (ICCM/9), Madrid, July 1993,
A. Miravete, ed., University of Zaragoza, Woodhead Publishing, 1993, 495 502.
8.159 Rockey, K.C. et al. (members of the Panel for Standard Practice for Testing),
Recommended Standard Practices for Structural Testing of Steel Models,
Supplementary Report 254, Transport and Road Research Laboratory, Department
of Environment, Department of Transport, Bridge Design Division Structures
Dept., Crowthorne, Berkshire, U.K., 1977.
8.160 Davidson, H.L., Postbuckling Behavior of Long Rectangular Plates, Fritz Engi-
neering Laboratory Report No. 248.15, Department of Civil Engineering, Lehigh
University, Bethlehem, PA, June 1965.
8.161 Katzer, W. and Murray, N.W., Elastic Buckling of Stiffened Steel Plates of High
Aspect Ratio Under Uniaxial Compression, Behavior of Thin-Walled Structures,
J. Rhodes and J. Spence, eds., Elsevier Applied Science Publishers, London and
New York, 1984, 355 373.
8.162 Birkemoe, P.C., Stability: Directions in Experimental Research, Proceedings of
SSRC 50th Anniversary Conference, Lehigh University, Bethlehem, Pennsylvania,
June 1994, SSRC Fritz Engineering Laboratory, 1994, 349 357.
8.163 Dahlen, C.U., Analytische und versuchstechnische Untersuchungen des Stabilitäts-
verhaltens von anisotropischen Faserverbundwerkstoffplatten, Diplomarbeit,
supervised by R. Surjana, Institut für Leichtbau, RWTH Aachen, August 1989.
References 601

8.164 Fughe, C., Experimentelle Ermittlung der Beullasten von CFK-Platten und
Überarbeitung des Prüfstandes, Studienarbeit, supervised by R. Surjana, Institut
für Leichtbau, RWTH Aachen, May 1991.
8.165 Stevens, K.A., Ricci, R. and Davies, G.A.O., Postbuckling Failure of Composite
Compression Panels, ICAS Proceedings 1994, 19th Congress of the International
Council of the Aeronautical Sciences, Anaheim, CA, September 1994, AIAA, Wash-
ington, D.C., 1994, 3, 2975 2981.
8.166 Stevens, K.A., Ricci, R. and Davies, G.A.O., Buckling and Postbuckling of
Composite Structures, Composites, 26, (3), March 1995, 189 200.
8.167 Stevens, K.A., Department of Aeronautics, Imperial College London, personal
correspondence to J. Singer, 23rd October 1996.
8.168 MTS System Corporation, Eden Prairie, MN 55344 USA, Specifications of Series
311 and Series 318 Load Units, 1988 and 1995.
Author Index

Abramovich, H., 536, 596 Belytschko, T., 129


Adams, J.E., 285 Bergfelt, A., 561, 598 599
Allen, H.G., 124 Bernard, A., 398
Allnutt, R.B., 483, 592, 594 Bernard, E.S., 407
Almroth, B.O., 41, 84, 169, 124, 127, Bertero, V.V., 405 406
178 179 Besseling, J.F., 12, 580 581, 592
Amazigo, J.C., 145 146, 178 Biggers, S.B., 206, 215
Arbocz, J., 12, 127, 173, 178 179 Bijlaard, P.P., 344 345, 403,
Argyris, J.H., 406 Biot, M.A., 304, 398
Ariaratnam, S.T., 203, 215, Birkemoe, P.C., 357, 404, 600
Ari-Gur, J., 598 Blachut, J., 127
Aschendorff, K.K., 398 Bleich, F., 400
Atsuto, T., 125 Block, D.L., 145, 178
Avent, R.R., 13 Bodner, S.R., 41, 125, 252, 284 285,
Ayrton, W.E., 194, 214 416 420, 423 424, 426, 450
Azizinamini, A., 405 Bogdanoff, J.L., 451
Boley, B.A., 404, 473, 476, 499, 502, 517,
Babcock, C.D., 44, 12, 126, 406, 419, 582, 594
423 426, 451 Bolotin, V.V., 128
Back, G., 455 456, 470, 472 473, 499, 502, Booth, E., 241 242, 244 247, 250, 258, 284
508, 591 Boresi, A.P., 126
Bagchi, D.K., 37, 125, Botman, M., 580 581, 592
Baker, D.J., 597 Boucif, M., 593
Baker, W.E., 254 258, 285 Bradburn, J.H., 405
Balas, G.J., 406 Bradfield, C.D., 485, 488, 505, 593, 594, 600
Ballio, G., 399 Brazier, L.G., 85, 87, 128
Bannister, R.C., 273, 275, 286 Bresse, M., 44, 125, 409
Bansal, J.P., 402 Bridget, F.J., 202, 215
Baruch, M., 126 Bridgman, P.W., 219, 284
Basler, K., 540, 542, 597 Brivtec, S.J., 187, 214, 369, 370, 404
Batdorf, S.B., 125, 147, 516 Brodka, J., 599
Batista, E., 399 Brogan, F.A., 127, 178
Beck, M., 2, 11, 100, 128 Bromley, S., 401
Becker, H., 125 126, 571 572, 577, 592, Broughton, P., 434, 440, 451
600 Brown, H.N., 255, 285
Becker, L.A., 483, 592 Brown, K.E.P., 597
Beedle, L.S., 213, 598 Bruhn, E.F., 128, 130

Buckling Experiments: Experimental Methods in Buckling of Thin-Walled Structures: Basic Concepts, Columns, Beams and
Plates – Volume 1. J. Singer, J. Arbocz and T. Weller Copyright © 1998 John Wiley & Sons, Inc.
604 Author Index

Brunner, O., 406 Cox, H.L., 179, 214, 456, 485, 592
Brush, Don O., 41, 124 Crawford, R.F., 451
Bryan, C.W., 397 Crawley, E.F., 286, 407
Bryan, G.H., 18, 125 Crisfield, M.A., 594
Buchert, K.P., 12 Crockett, H.B., 309, 398
Buck, O., 398 Croll, J.G.A., 179, 189, 214
Buckingham, E., 219 221, 224, 225, 232, Cundari, F.L., 595
274, 284 Czechowski, A., 599
Budiansky, B., 4, 12, 64, 127, 139, 142, 147,
177, 539, 596 DaDeppo, D.A., 450
Bulson, P.S., 124 Dahlen, C.U., 600
Burr, W.H., 401 Datta, P.K., 595, 596
Burt, C.A., 597 Davids, A.J., 346, 399, 403
Bush, H.G., 543, 597 Davidson, H.L., 580, 600
Bushnell, D., 7, 13, 127, 129, 212, 215 Davies, G., 599
Butterworth, J.W., 404 Davies, G.A.O., 284, 405, 601
Byon, O-I., 460, 495, 498 499, 509, 593 Davison, J.B., 371, 405
Demuts, E., 452
Calladine, C.R., 241, 245 247, 248, 250, Denston, R.J., 595
258, 271, 285 Denton, D.R., 256, 285
Campbell, J.D., 284 Di Tommaso, A., 450, 452
Card, M.F., 178 Dickie, J.F., 434, 440, 451
Carlson, R.L., 522, 526 527, 593, 595 596 Dickinson, H.B., 592
Chailleux, A., 531, 596 Dodge, F.T., 254, 285
Chajes, A., 400, 401 Domburian, E.M., 593
Chan, H.C., 434, 451 Donelan, P.J., 259, 286
Charlton, T.M., 284 Donnell, L.H., 50 52, 55 58, 61 63, 67,
Chen, W.F., 125, 371 372, 402, 404 405 75, 126, 132 133, 139, 143, 161, 166,
Chen, Y., 178 170, 177, 198 203, 206, 214, 520 521,
Cheng, J.J., 343, 402 523, 533 534, 592
Cheng, S., 126 Dow, N.F., 12
Cheresh, M.C., 284 Dowling, A.R., 286
Cherry, S., 390, 406 Dowling, P.J., 10, 12 13, 279, 286, 406,
Cheung, M.C., 419, 423 426, 451 593 594, 597, 600
Chilver, A.H., 3, 12, 187, 214, 230, 233 234, Drdacky, M., 403, 599
304, 306, 369, 398, 400, 592 593 Driscoll, G.C., 404
Chini, S.A., 431 433, 451 Drucker, D.C., 1 2, 6, 11
Christoforou, A.P., 271, 286 Duffy, D.J., 483, 594
Clark, E., 591 Duffy, T.A., 257, 284
Clark, P.J., 286 Dumont, C., 332, 401
Clement, M., 593 Dunn, L.G., 400, 450, 456, 592
Coan, J.M., 21, 125, 167, 473, 476, 499, 502, Durelli, A.J., 284
517, 519 520, 536 537, 582, 594 du Plessis, D.P., 402
Cohen, G.A., 177 178 Dwight, J.B., 470, 485, 503, 505, 572, 594,
Colao, A., 600 600
Collier, D., 241, 284 Dym, C.L., 124, 136, 283
Conley, W.F., 483, 592
Connor, R.W., 539, 596 Eash, M., 598
Considère, A., 91, 183, 128 Eiden, M., 406
Conway, H.D., 409, 418, 450 Elgaaly, M., 125, 598
Costa Ferreira, C., 399 Ellinas, C.P., 398
Costello, M.G., 212, 215 Elmore, G.H., 401
Author Index 605

Elnashai, A.S., 406 Galletly, G.D., 78, 80, 127


El-Ghazouli, A.Y., 406 Gedies, R.W., 402
Engesser, F., 196, 198, 214 Geier, B., 145, 170, 178 179
English, R.W., 247, 285 Gent, A.R., 398
Epstein, H.I., 401 Gerard, G., 124 126, 310 311, 313, 399
Esslinger, M., 170, 179 Gere, J.M., 72, 124, 158, 412, 440, 521
Estuar, F.R., 398 Gerstle, K.H., 371, 376, 404
Euler, L., 16, 18, 22, 25 26, 28, 29, 42, 46, Ghomi, S., 597
55, 87, 91, 96 98, 101, 125, 137, 181, Girkmann, K., 598
183 184, 195 196, 198, 206, 213, 265, Gjelsvik, A., 416 420, 423 424, 426, 450
304, 322, 331, 344 345, 422, 442 Godden, W.G., 452
Evans, H.R., 546, 549, 597 598 Goel, S.C., 405
Evan-Iwanowski, R.M., 434, 451 Goldberg, J.E., 451
Ewing, W.O., 256, 285 Goldman, R., 592, 600
Ezra, A.A., 257, 285 Goodier, J.N., 228 230, 234 237, 284, 400
Gough, H.J., 214
Fahlbusch, G., 399 Goverdhan, A.V., 371, 405
Fang, P.J., 401 Granholm, C.A., 561, 599
Farley, G.L., 597 Graves-Smith, T.R., 399
Fasanella, E.L., 263, 286 Green, B.E., 215
Faulkner, D., 456, 459, 592 Gregory, M., 204, 215, 404
Fernandes-Sintes, J., 595 Griffith, A.A., 283, 287
Finzi, L., 399 Groot, W.J., 403
Fischel, J.R., 592 Gukhman, A.A., 284
Fischer, G., 84, 126, 127, Guyon, E., 593
Fischer, M., 476, 478, 502, 594
Fisher, G.P., 344 345, 403 Hall, D.B., 126
Fisher, H.R., 197, 214 Hamoodi, M.J., 597
Fitch, J.R., 177 Hancock, G.J., 346, 351, 399, 401, 403, 407,
Flathau, W.J., 256, 285 599
Fligg, C.M., 320, 322, 400 Hanks, B.R., 406
Flint, A.R., 401 Hanna, J.W., 256, 285
Florence, A.L., 252, 285 Hans, Y., 596
Flügge, W., 65, 72, 78, 126 127, 132 Hanson, R.D., 406
Fok, C.D., 512, 514, 595 Harari, O., 126
Fok, W.C., 476, 499, 502, 529, 533, 536, Harding, J.E., 12, 593 594, 597, 600
538, 593, 596 Hariri, R., 9, 13
Foss, G., 12 Harre, W., 476, 478, 502, 594
Foutch, D.A., 405 Harrison, J.C., 595
Frieze, P.A., 593 594, 597, 600 Harvey, J.M., 476, 499, 502, 591, 593
Fughe, C., 601 Hattori, R., 402
Fukumoto, Y., 33, 334 336, 398, 402, 403, Hausman, R.J., 286
592 Hayashi, 398
Fukuta, T., 406 Hayashi, T., 198, 214
Fung, Y.C., 178, 412, 413 415, 450 Hedgepeth, J.M., 126, 406
Föppl, L., 82, 127 Heimerl, G.J., 398, 596
Herr, R.W., 406
Gaber, E., 409 411, 439, 451 Herrman, J., 439, 452
Galambos, T.V., 333, 398, 402 403 Herrmann, G., 2, 11
Galerkin, B.G., 60, 64, 106, 108 110, 128, Hertel, H., 540, 597
145, 166, 171 Herzog, M.A.M., 561, 564, 597, 598, 599
Gallaher, G.L., 398, 596 Hetenyi, M., 1, 11
606 Author Index

Hetrakul, N., 599 Jones, R.P.N., 179


Hickman, W.A., 12 Jouri, W.S., 286
Hill, H.N., 332, 401 Ju, G.T., 128
Hill, R., 94, 128
Hirschfeld, T., 12 Kalyanaraman, V., 403
Hjelmstad, K.D., 406 Kampf, K.-P., 286
Ho, B.P.C., 126 Kaplan, A., 412 415, 450
Hoa, S.V., 500 Kappus, R., 320, 400
Hobbs, R.E., 593 594 Kato, B., 398, 404, 406
Hoff, N.J., 84, 128, 169, 179, 214, 368, 400, Katzer, W., 600
404, 456, 473, 476, 502, 517, 536 537, Kaufman, S., 404
582, 592, 594 595 Kawashima, K., 425, 427 429, 451
Hol, J.M.A.M., 127, 178 Kellas, S., 268, 286
Holloway, D.M., 206, 215 Kempner, J., 169, 178 179
Holt, M., 13 Kennedy, J.B., 326, 401
Hone, C.P., 400 Kerr, A.D., 45, 126
Hopkinson, B., 254 256, 285 Key, P.W., 399
Horne, M.R., 404 Khot, N.S., 145 146, 171, 178
Horner, G.C., 406 Kihira, M., 198, 214
Horst, P., 597 Kim, Y.S., 600
Horton, W.H., 204, 215, 521, 523, 528, 595 Kirby, P.A., 214, 405
Howe, D., 399 Kirsch, B., 182 183, 186, 214
Hsu, M.Y.H., 428, 430, 451 Kishi, N., 317, 372, 404, 405
Hu, P.C., 125, 167, 516 Klouman, F.L., 399
Huang, N.C., 11 Koiter, W.T., 2 3, 11 12, 46, 54, 81, 110,
Huber, A.W., 398 125 128, 133 135, 139, 143, 147, 154,
Huddleston, J.V., 409, 439 440, 450 160, 175, 178
Hull, F.H., 580, 593 Kolkka, R.W., 11
Humpherson, J., 495, 595 Kollbrunner, C.F., 295, 320, 398, 400, 479,
Hunt, G.W., 124, 423, 451, 600 480, 509, 591 592
Hurlbrink, E., 45, 126, 409 Kollet, M., 598
Hutchinson, J.W., 4, 12, 71, 74, 127, 139, Korol, R.M., 594
142, 145 147, 173, 177 179 Kossira, H., 597
Hövik, J., 598 Kouhia, R., 403
Krawinkler, H., 388, 406
Ipsen, D.C., 284 Kroll, W.D., 400
Ito, T., 425, 427 429, 451 Krätzig, W.B., 179
Itoh, Y., 402, 592 Kubo, M., 334, 402
Ivey, E.S., 283 Kuhn, P., 540, 597
Kuiken, G.D.C., 125
Jackson, K.E., 263, 267, 269, 286 Kuroki, F., 434, 440, 451
Jerome, C.C., 202, 215 Kutmanova, I., 599
Johansson, B., 598 Kwon, Y.B., 407
Johnson, C.P., 402 Kyriakides, S., 88, 128
Johnson, J.B., 397
Johnson, R., 204, 215 Lange, J., 452
Johnson, R.W., 595 Langhaar, H.L., 125, 219, 225, 284
Johnson, W., 251 252, 285 Lau, S.C.W., 399, 401
Johnston, B.G., 5, 12 Lee, G.C., 329, 333, 401
Jones, N., 241, 259, 284, 286 Lee, G.H., 287
Jones, R.E., 215 Lee, S.-J., 405
Jones, R.F., 212, 215, Legget, D.M.A., 179
Author Index 607

Leicester, R.H., 205, 215 Michell, A.G.M., 329 331, 401


Levin, L.R., 597 Michielsen, H.F., 179
Levy, S., 179, 197, 201, 214, 322, 400, 509 Midorikawa, M., 405
Libai, A., 505, 507, 595, 598 Mikulas, M.M. Jr., 178
Lin, E.M., 402 Miles, J., 241, 284
Lindström, G., 403 Miller, E., 127
Little, G.H., 594 Mindlin, R.D., 287
Llopiz, C.R., 405 Minguez, J.M., 278, 491 494, 502, 509, 595
Lo, C.F., 409, 418, 450 Mitchell, D.H., 434, 451
Lo, H., 434, 451 Mofflin, D.S., 505, 594
Loomis Richard, W., 401 Moore, H.F., 331, 401
Loomis Robert, W., 401 Morton, J., 259, 260 261, 267 269, 286
Loomis, R.H., 401 Moxham, K.E., 485, 594, 600
Loomis, R.S., 401 Mulligan, G.P., 403
Lorenz, R., 55, 126 Murphy, G., 284
Loughlan, J., 399 Murrey, N.W., 2, 11, 215
Lu, L.W., 404 Murty, K.S.M., 401
Lu, L.-W., 405 Müller, R.K., 283
Lui, E.M., 371, 404
Lundquist, E.E., 125, 132, 197 198, 200, Nachbar, W., 11
201 203, 214, 320, 322, 343, 345, 516, Nagwaney, A., 399
524, 526 528, 533, 400, 404 Nakamura, T., 403, 593
Luongo, A., 403 Nakashima, M., 390, 403, 406
Löwitova, O., 599 Narayanan, R., 125, 213, 398, 402 404,
Lévy, M., 44, 125, 409 597 598, 600
Nardo, S.V., 404
Maaskant, R., 593 Nash, W.A., 428, 430, 451
MacNeal, R.H., 283, 287 Neal, B.G., 594
Madsen, W.A., 179 Needham, A, 310, 398
Madugula, K.S.M., 326, 401 Neff, J., 538, 596
Maguerre, K., 456, 592 Nemat-Nasser, S., 11, 12
Mahin, S.A., 405, 406 Nethercot, D.A., 125, 214, 333, 334, 371,
Majumdar, S., 508, 522, 595 402, 405
Mang, F., 398 Neut, A. van der, 25 26, 28, 125, 128, 178,
Manjoine, M.J., 284 344
Marburg, E., 331, 401 Newell, J.S., 400
Marley, M.J., 404 Niles, A.S., 321, 400
Marlowe, M.B., 178 Nishiyama, I., 405 406
Marsh, K.J., 284 Nontanakorn, D., 449 450, 452
Martin, J.B., 251 253, 285 Noor, A.K., 129
Massey, C., 205, 215 Notenboom, R.P., 125
Massey, P.C., 344, 402 Nuismer, R.J., 286
Masur, E.F., 45, 50, 126, 419, 424, 426 Nurick, G.N., 251 253, 285
Mayers, J., 179
McElman, J.A., 74, 127 Ojalvo, M., 449, 450, 452, 580, 593
McMinn, S.J., 434, 451 Okuhara, H., 599
McPherson, A.E., 197, 201, 214, 509 Ominami, R., 594
Meck, H.R., 126, 343, 402 Osgood, W.R., 13, 92, 128, 228
Meirovitch, L., 128 Owens, G.W., 279, 286
Meister, M., 591
Menken, C.M., 352, 403 Packer, J.A., 599
Merchant, W., 203, 215, 404 Palmer, A.C., 214
608 Author Index

Pankhust, R.C., 284 Roberts, T.M., 562, 597 599


Papangelis, J.P., 443 449, 452 Robertson, A., 196, 214
Parks, M.B., 599 Robinson, W.H., 401
Pazerycki, J., 600 Rodney, K.C., 179
Pearce, H.T., 285 Roeder, C.W., 380, 405
Pekoz, T., 400, 403 Rogers, N.A., 595
Penning, F.A., 285 Rondal, J., 399
Perry, J., 194, 214 Roorda, J., 148, 178, 187 191, 207, 209,
Peterson, H.T., 127 211 215, 369, 410, 419 421, 423 424,
Peterson, J.P., 597 451, 520, 533, 536 537, 593, 596
Pfeiffer, P.A., 452 Rosen, A., 505, 595
Pflüger, A., 125 Rothwell, A., 403
Phillips, E.A., 284 Rouse, M., 597
Pignataro, M., 403 Roy, J.A., 596
Pilkey, W., 128
Pindera, J.-T., 12 Sachs, R.G., 255, 285
Pinson, L.D., 406 Saczalski, K., 128
Plumier, A., 398 Saitoh, M., 434, 440, 451
Poowannachaikul, T., 402 Salkar, R., 598
Popov, E.P., 406 Salmon, E.H., 181 182, 213
Porter, D.M., 597 Salvadori, M.G., 287
Potier-Ferry, M., 179 Sandhu, J.S., 405
Potters, M.L., 127 Sandhu, R.S., 449, 452
Pozerycki, J., 592 Santaputra, C., 599
Prandtl, L., 2, 181, 183, 213, 282, 287, 329, Schaeffer, H., 128
330 Schaller, I., 599
Prescott, J., 401 Schlack, A.L., 476 477, 499, 502, 529, 530,
Pretschner, W., 320, 400 594
Prion, H.G.L., 357, 404 Schmidt, R., 450
Procter, A.N., 401 Schnell, W., 126
Pugsley, A.G., 400 Schreyer, H.L., 45, 50, 126, 419, 424, 426
Pyle, J.S., 407 Schuette, E.H., 12, 398, 596
Schulz, G.W., 598
Qian, Y., 260, 268, 270, 286 Schuman, L., 455 456, 470, 472 473, 499,
Queinec, A., 595 502, 508, 591
Schwerin, E., 63, 126
Ractliffe, A.T., 470, 485, 503, 594 Sechler, E.E., 5 6, 12, 173, 178 179, 206,
Radhamohan, S.K., 127 215, 456, 460, 472, 592
Radkowski, P.P., 127 Segall, A., 596
Radziminski, J.B., 405 Seggelke, P., 145, 178
Ramberg, W., 92, 128, 197, 207, 214, 228, Sehested, J., 536, 537, 596
322, 400, 509 Seide, P., 127
Ramirez, D.R., 402 Sewell, M.J., 94, 128
Raoul, J., 599 Sfintesco, D., 598
Rasmussen, K.J.R., 399, 403 Shanley, F.R., 18, 91, 125
Rentschler, G.P., 404 Shanmugam, N.E., 600
Reynolds, T.E., 212, 215 Sharman, P.W., 495, 595
Rhodes, J., 399, 476, 499, 502, 591, Shaw, W.J.D., 451
593 594, 599, 600 Shimizu, S., 599
Ricci, R., 601 Shing, P.-S.B., 405 406
Riks, E., 172, 179 Simitses, G.J., 124
Rivello, R.M., 17, 125, 310 Simo, J.C., 129
Author Index 609

Singer, J., 12, 13, 44, 126, 178 179, 598, Thompson, J.M.T., 124, 403, 420, 423, 451,
601 600
Skaloud, M., 542, 597, 599 Thomson, W.T., 234 237, 284
Skan, S.W., 61, 126, 538 Timoshenko, S.P., 45, 55, 62, 72, 124, 126,
Smith, C.S., 270, 276 277, 286, 599 158, 214, 401, 409, 412, 440, 443,
Smith, C.V., 593 Tokarz, F.J., 449 450, 452
Smith, E.A., 401 Trahair, N.S., 333 334, 336, 402, 443 449,
Smith, N.L., 286 452
Soiter, M.T., 126 Trefftz, E., 97, 102, 105, 119, 128, 137
Soong, T.C., 84, 128 Tsai, W.T., 206, 215
Southwell, R.V., 55, 59, 61, 126, 186, Tsao, C.H., 284
194 209, 212 215, 520 521, 524, Tschamper, H., 598
526 528, 538 Tsien, H.S., 132, 168 169, 177, 412,
Souza, M.A., 536, 596 450 451
Spencer, H.H., 214, 528 529 Tsuyoshi, 398
Spier, E.E., 399 Tuckerman, L.B., 201, 203, 215, 323, 345
Springer, G.S., 596 Tulk, J.D., 279, 287
Sridharan, S., 399 Turneaure, F.E., 397
Srinivasan, G.V., 287
Uang, C.M., 405 406
Stallenberg, G.A.J., 403
Ueda, Y., 594
Starnes, J.H., 127
Uemura, M., 460, 495, 498 499, 509, 593
Stavrinidis, C., 406
Uetami, K., 593
Stein, M., 74, 84, 127, 460, 495 496, 502,
Urbano, C., 399
538, 582, 593, 596
Usami, T., 298, 398, 403
Stephens, W.B., 127, 453
Stevens, K.A., 405, 589, 590, 601 Vacharajittiphan, P., 452
Stonor, R.W.P., 600 Valsgard, S., 12
Stroebel, G.J., 593 Valsgård, S., 600
Stussi, F., 592 Van Driest, E.R., 284
Sun, C., 434 439, 451 van Erp, G.M., 403
Supple, W.J., 398, 404, 592 593 Van Kuren, R.C., 403
Sutherland, R.L., 287 van Musschenbroek, P., 181, 213
Sutherland, S.H., 284 Vann, W.P., 536, 537, 596
Swanson, S.R., 260, 270, 286 Vaswani, H.P., 370, 404
Symonds, P.S., 252, 284 285 Venkataramaiah, K.R., 536 537, 596
Venkayya, V.B., 145, 178
Takanashi, K., 390, 406 Verchery, G., 596
Tall, L., 213, 290 291, 397 398 Vilnay, O., 179, 597
Tam, L.L., 246, 248, 250, 285 Vinogradov, A.M., 451
Tang, K.H., 598 Viola, E., 450, 452
Tappin, R.G.R., 286 Virkar, A.V., 287
Tasi, J., 126 Vlasov, V.Z., 67, 170, 443, 452
Taylor, E.S., 284 von Kármán, T., 91, 132, 156, 158, 159, 161,
Taylor, G.I., 283, 287 166, 168, 181 183, 185 186, 195 196,
Tebedge, N., 398 203, 213, 289, 304, 307, 329, 398, 456,
Tennyson, R.C., 12 458, 592 593,
Theillout, J.N., 599 von Tetmajer, L., 214
Thewalt, C.R., 406 Vosseler, A.B., 215
Thielemann, W., 126
Thimmhardy, E.G., 594 Wagner, H., 320, 400, 508 509, 540, 542,
Thomas, E.W., 400 552 554, 592, 597
Thomassen, P.O., 403 Wakabayashi, M., 403
610 Author Index

Walker, A.C., 189, 214, 279, 287, 398 399, Wittek, U., 179
455, 473 475, 499, 528, 582, 591, 593, Wolde-Tinsae, M.W., 451
595, 596, 599, 600 Wright, D.V., 273, 275, 286
Walkner, C., 399
Wallace, B.J., 388, 406 Yaffe, R., 596
Walt, M., 592 Yamaki, N., 63, 126, 476, 499, 500 502,
Wan, C.C., 133, 177 509, 582, 593
Warner, W.H., 593 Yamamouchi, H., 405, 406
Washizu, K., 125 Yao, T., 594
Watabe, M., 405 Yarmici, E., 404
Way, E.R., 78, 205, 215, 310 Yoo, C.H., 449, 452
Weingarten, V.I., 127 Yoshida, S., 599
Weller, T., 536, 543, 596 598, Yu, T.X., 246, 285
Wells, S., 13 Yu, W.W., 400, 599
Wen, R.K., 452 Yuen, M.F., 529, 596
Wesfreid, J.E., 593 Yura, J.A., 337, 343, 402, 404
Westergaard, H.M., 203, 215
Westine, P.S., 254, 257 258, 285 Zaal, K., 507, 509, 536, 595
White J.D., 595 Zagustin, E.A., 439, 452
Whittaker, A.S., 406 Zandonini, R., 3, 12, 371, 598
Wierzbicki, T., 252, 284 285 Zaras, J., 399
Willems, N., 2, 11 Zele, F., 127
Williams, D.G., 600 Zhang, T.G., 246, 285
Wilson, E., 189, 214 Zhao, X.-L., 599
Wilson, J.F., 206, 215 Ziegler, H., 128
Wilson, J.M., 540, 596 Zielsdorff, G.F., 595 596
Winter, G., 400 401, 403, 592 Zienkiewicz, O.C., 121, 124, 129
Subject Index

Adjacent-equilibrium criterion, 52, 68 Axisymmetric imperfections, cylinder with,


Admissible functions, 102 128, 147, 175
Airy stress function, 51, 71, 74, 139, 143
Analogies, 282 283, 287 Batdorf parameter, 147
electrical circuit analogies, 283 Beam columns, 125, 197, 263, 402 403
membrane analogy for torsional stress, 282 computerized test control, 365, 367
Annular plates, 505 506, 508, 522, lateral-torsional instability, 40
524 525, 595 maximum moment, 422
heating of loading ring, 508 maximum stress, 165, 232, 239 240, 251
Technion loading system with radially nonlinear bending, 40
moving segments, 505 507 profiling rig, 358, 360, 365
Arch rise, 46 special spherical bearings, 364
Arch rise parameter, 409 University of Toronto large scale tubular
Arches, 41, 409 410, 412, 414, 416, 418, member tests, 357 367
420, 422, 424, 426 428, 430, 432, 434, weld induced residual strains, 357
436, 438, 440, 442, 444, 446, 448, Beam-column connections, 371
450 452 connection data base, data banks, 10, 154,
circular, 215, 425, 450 371
clamped, 451 semi-rigid connections, flexible joints, 371
deep cross section, 410, 440 Beams, lateral instability, 328, 401, 553
high rise, 409, 434, 439 Bending, cylinder subjected to, 84, 179
ideal, 422 Bending energy, 42, 95, 154
prestressed, 422, 431, 439, 450 451 Bifurcation behavior, 186, 191
shallow, 45, 69, 72 75, 114, 209, 431, Bifurcation buckling, 52, 83, 114 115, 131,
451, 481, 504 134 135, 151, 186, 409
simply supported, 44, 411 412, 419, 441, effect of material nonlinearity, 81, 88 94
444 with linear prebuckling theory, 53 57
sinusoidal, 419, 422 423 with nonlinear prebuckling theory, 80 86
steep, 409, 439, 450 Bifurcation point, 26, 45, 48 49, 52, 68, 81,
Aspect ratio, plate, 525 526 95, 131 134, 137 138, 154, 168, 173
Asymmetric bifurcation, 50 Bridging gaps between disciplines, 9
Asymmetric imperfections, cylinder with, 175 Boundary conditions, 3, 6 7, 16, 20, 29, 35,
Asymptotic theory, 150 151, 153 154, 160 40, 43 45, 47, 49, 54, 58, 60, 62 63,
Axially compressed cylindrical shells, 81, 72, 80, 82, 84 85, 96, 98 104, 106,
127, 132 133, 170, 179 108 109, 131, 133, 137, 141, 145, 147,
Axisymmetric collapse, 82 83, 114 151, 155, 161, 173, 176, 178, 185 187,

Buckling Experiments: Experimental Methods in Buckling of Thin-Walled Structures: Basic Concepts, Columns, Beams and
Plates – Volume 1. J. Singer, J. Arbocz and T. Weller Copyright © 1998 John Wiley & Sons, Inc.
612 Subject Index

194, 203, 212, 227, 272, 277, 278, 282, of beam columns, 356
318, 364, 409, 459, 462, 464, 470, of columns, 18, 342
472 473, 475, 476 478, 485, 491 492, of cylindrical shells, 63
494, 495, 497 499, 501, 505, 508, 517, of plates, 533 534
533, 534, 538, 542, 546, 571, 576 581, of rings, 41, 437
584 585, 587 588, 593, 595 of shallow arches, 45
Boundary conditions for plates of shallow spherical caps, 69 72
double knife edges for clamped edges, 476 of shells of revolution, 66 69
Kollbrunner’s circular cylindrical riders, of toroidal shell segments, 127
479 481 Buckling mode, 16, 24, 27, 30, 36, 44, 48,
knife edges, 184, 209, 295, 297, 307, 321, 60 61, 63, 78, 80, 98, 109, 120, 124,
330, 345, 369, 412, 414, 419 421, 134 137, 140, 142, 144 147, 176, 195,
464, 473, 476, 480, 495 497, 206, 314, 334, 347, 351, 412, 414, 418,
499 501, 578, 580 583 430, 460, 492, 494, 508 509, 514, 520,
needle bearing blocks, 476 477 538
roller bearings, 295, 298, 308, 331, 473, Buckling and postbuckling of columns, 289
476, 524 525, 584, 587 column curves, empirical design formulae,
rollers in circular slots, 492 9, 185, 187, 289 290, 292
rolling plates (which house ball bearings),
column testing, 290, 294, 299 300
481
columns in offshore structures, 303
rounded knife edges, 476, 582
end fitting effects in column tests, 304
semi-circular needle bearings, 477
secondary effects, 11, 184
tensioned steel wires support, 491 493
Buckling under internal pressure, 78 80
V-grooves, 310, 313, 471 472, 481
Built-up structure, 344
Boundary conditions, influence of:
modal interaction, 28
on columns, 295 297, 304 309
on cylinders, 62, 176 Buoyancy force, 449 450
on plates, 577 580
Cambridge University “finger” supports,
Boundary value problem, 29, 155
484 491
Brazier effect, 85, 87, 128
Bradfield’s improved “fingers”, 488
Buckingham Pi theorem, 219
Dwight and Ractliffe “fingers”, 485
Buckling coefficient, rectangular plate, 20,
horizontal 100-ton plate test rig, 486 489
106
Centrifuge model testing, 431
Buckling experiment, 151, 175, 184 186,
197, 229, 237, 275 277, 303, 318, Certification tests, 7
321 322, 327 330, 332 333, 337, 345, aircraft certification test, 8
352, 368, 370, 377, 382, 460, 509, Circular cylindrical shells (see Cylindrical
514 515, 545, 553 554, 556 557, 577, shells), 50 51, 66, 84, 126 128, 167,
579, 582, 584, 588 179
aims, 1, 186, 324, 507, 578, 587 Circular rings, 41, 409, 451
basic elements, 185 kinematic relations, 50 51, 67 68, 71, 161
boundary conditions, 579, 580, 584 588 nonlinear equilibrium equations, 81, 95
design of proper experiment by its purpose, potential energy expression, 112
577 580 stability equations, 43, 50, 52 54, 57, 60,
effects of scaling, 186, 285, 578 62 63, 68 70, 73 74, 80 81, 84,
ideal perfect structures, 186 95, 97 98, 102, 104, 109, 140, 156
real imperfect structures, 186 Circumferential waves, 60, 84
Buckling of frameworks, 367 Classical buckling, 137, 186
no-sway frames, 368 Classical problem of fluid dynamics, 218
sway frames, 368 COLA pseudodynamic test system, 379, 389
Buckling load, 28, 179, 214, 369, 400, 402, loading history created in parallel to
404, 530 loading, 390
Subject Index 613

on-line earthquake response test technique, Crippling tests, 311 314, 317, 564, 569
389 local postbuckling behavior of corners, 314
pseudodynamic method, 379, 405 crinkly collapse, 314 317
Collapse, 94, 127, 178, 281, 285, 315, 399, Crown of the arch, 422 423
401, 540, 594, 597 600 Curvature, column, 315
near bifurcation point, 133, 188 Cusp type behavior, 424
Collapse stress, 233 234 Cylindrical shell buckling
Column buckling, 98, 317, 344, 346 axial compression, 53 57
bifurcation buckling, 52, 83, 114 115, discrepancy between test and theory, 132
131, 134 135, 151, 186, 409 effect of length, 54, 55
initial post-buckling behavior, 127, imperfection sensitivity
136 139, 207, 503 post-buckling behavior, 143 154
mode interaction, 6, 28, 344 short cylinders, 55
Column failure, 125, 321 322, 400 stiffened, 143 148, 173 175
Columns, 9, 13, 125, 177, 181, 197, 213, combined loads, 57 66
215, 290, 292, 294, 296, 298 300, 302, external pressure, 57 59
304, 306, 308, 310, 312, 314, 316, 318, torsion, 59 63, 153, 171
320, 322, 324, 326, 328, 330, 332, 334,
Cylindrical shells, 13, 50, 63, 69, 81,
336, 338, 340, 342, 344, 346, 348, 350,
126 128, 133, 143, 151, 167, 169,
352, 354, 356, 358, 360, 362, 364, 366,
177 179, 231, 256, 258, 270 271, 377
368, 370, 372, 374, 376, 378, 380, 382,
384, 386, 388, 390, 392, 394, 396, Data bank, 10, 154, 372
398 404, 406, 596 Dead loading, 42, 353
Column Research Committee of Japan, 290, Dead weight, 339, 369, 416, 422, 443, 505
398 Deflection measurements, 197, 335, 499, 555,
Combined loading, 172, 539, 564 576
cylinders Cambridge University potentiometer-type
Compatibility equations: transducer, 510, 511
cylindrical shell, 52, 60, 63, 71, 171
dial gages, 508
plate, 166
Fok’s plate scanner, 512
Complementary nature of experiment and
LVDT’s, 300, 321, 351, 365, 376,
theory, 2
380 381, 383, 429, 509, 570
Composite ship hull structure, 276 277
LTD’s, 492, 509
Computer simulated experiments, 5
Demonstration experiment, 187, 189,
Computer programs, 6 7, 178, 212, 283
BOSOR 4, 7, 112 191 193, 206, 370
BOSOR 5, 13, 127 Design codes, 7, 10, 377
MARC, 124, 129 Determination of critical load, 199, 319, 516,
STAGS, 66, 127, 555 556, 571 519, 534, 596
Contact buckling, 434 applications of Southwell plot to composite
Controlled initial deflections, “bumping”, 516 plates, 531, 532
Conservative system, 41, 95, 121, 123 Caltech applications to annular plates, 508,
Coped steel beams, 337, 402 522
application of the Southwell method, 197 extrapolation of strain difference curves
University of Texas lateral buckling tests method, 517, 518, 536
on coped beams, 339 343 Georgia Tech applications of Southwell
Correlation factor, 6 7 plot, 522 528
Critical load, 11, 178, 596 inflection point method, 518 519, 536
Crippling failure, 309, 311 315 least square fit technique for buckling load,
Crippling strength, 309, 311, 399 534 536
Crippling stress, crushing stress, 309 pivotal plots for plates, 528 531
Gerard’s method, 310 311 Southwell’s method for plates, 520
614 Subject Index

Determination of critical load (continued): knife edge end fittings, 304 309
Stanford University applications of position fixed, 295, 297
Southwell plot, 521 523 roller bearings, 295 298
sharp break in mean compressive strain End fixtures, 183 184, 187, 294 298
method, 536 length correction, 184, 304 307
sharp break in transverse strain method, Equal energy load, 416, 418
537 Equilibrium paths, 80, 86, 132 135,
strain reversal method, 516 518, 537 149 151, 167, 187, 189, 207, 208, 353
three-point technique for buckling load, complementary, 188, 422
533 experimental, 150, 188
top-of-the-knee method, 325, 462, for axially compressed cylinder, 132
516 520, 537 for initially straight column, 137
Diagonal tension, 237, 540 541, 549, 597 for rectangular flat plate
in plates without holes, 237 Equilibrium point, 121
in plates with holes, 237 Equilibrium state, 53, 94, 115, 123, 171
Dimensional analysis, 218, 220, 230, 236, Euler column formula, 16, 181
284 Euler equations for the calculus of variations,
Dimensionless products, 218 221, 224 226, 42
232, 236, 238 239, 251 Euler load (also see Column and Wide
Dimensionally homogeneous and column), 29, 183 184, 195 196, 206,
nonhomogeneous, 219 322, 422
Direct methods for determination of buckling European Convention for Constructional
loads in plates - summary, 533, 536, 537 Steelwork (ECCS), 398
Distortional buckling, 320, 326, 401, 407 Experiment as essential link, 1
distortional mode, local-torsional mode, Experimental mechanics, 11 12, 285, 399,
326 404, 452, 594 595, 597
University of Sydney experiments, Experimental optimization, 4
326 328 Experimental stress analysis, 1, 11, 284,
Donnell’s applications of Southwell plot, 198 286 287, 594
Donnell equations, 50 Experimental verification, 11, 148, 284
Drag coefficient, 219 Experiments on axially compressed plates,
Dynamic loading, 7, 238, 251, 254, 259, 267, 470 516
284 DTMB (David Taylor Model Basin) plate
buckling tests, 479, 483 484
Eccentricity of loading, load eccentricity, 182 ETH Zürich plate buckling tests, 479 481
Edge buckling, 120 U.S. Bureau of Standards test setup, 470
Effective length of columns, 300 Experiments on plates subjected to shear, 542
Effective slenderness ratio, 294, 388 aerospace industrial test setups, 558 560
Effective width, 125, 178 179, 214, Boeing 757 shear panel test, 558, 559
455 459, 492 593 SAAB 340 shear panel test, 558, 560
plate slenderness parameter, 459 picture frames, 542 543, 545
reduced effective width, 459 bending and extension of frame
Eigenvalues, 35, 54, 58, 62, 72, 101 102, members, 543
111 112, 119, 121, 123, 141 corner-pin location, 543 545
Elastic rotational restraint, 30 31 NASA Langley modified picture frame,
Elasto-plastic buckling, 232 543 546
Elliptic integral, 155 rotational clearance, 543
End conditions, 295 plate girders under shear, 546
practical pinned-ends, 295, 296 influence of relative rigidity of stiffeners,
ball bearing end fitting, 309 551
direction fixed, 295 sway collapse mechanisms, 551
hemispherical pin-end supports, 295, 297 Technion repeated buckling tests, 552 558
Subject Index 615

shadow-moiré technique, 499, 555, 556 Hydrostatic pressure loading, 58


strain measurement, 346, 349 Hyperbolic relationship, 194
Wagner beams with three point loading,
552 Imperfect column, 154 160
Wagner beams, 542, 552 553 plate, 166, 167
cylindrical shell, 167 177
Failure, 33, 125, 215, 286, 458, 597 598, Imperfections, 154, 177 178, 181, 464, 594
601 Imperfection-sensitivity, 133 134, 147, 178
Finite difference method, 119 asymptotic theory, 148, 153
Finite displacements, 169, 412 Koiter’s general theory, 133 135
Finite element (also see Discretization), imperfection sensitivity parameter b,
121 124 145 148
First variation, 116, 162 limitations of asymptotic theory, 153,
Flat panel hinged on three sides and free on 154
fourth, 199, 201 maximum load as function of
Flat plates, 125, 592, 596 imperfection sensitivity parameter
Flaws and local defects, 182 and imperfection, 151
Flexural-torsional stability, 440, 452 cylindrical shells
Fluid-pressure loading, 42, 44 axial compression, 177, 179
circular rings monocoque, 207, 592
Fourier series, 118, 167, 194 195, 347 oval cylinders, 154
Frame analysis, 367 397 stiffened, 143
Framework joint flexibility, connection imperfection shape, 6
flexibility, 371, 376 measured, 154, 173
Frames, 28, 197, 203, 214, 367 369 random, 151
Free edges, 20, 521 initial post-buckling behavior (see
Functional, 96 97, 176, 233 asymptotic theory), 127, 207
Fundamental path (also see load-deflection load deflection curves, 247
curve), 45, 52 53 load-imperfection-amplitude curves,
Fluid structures interaction, 7 147 149
Full scale (test), 7, 8, 485, 578 modal interaction, 28
nonlinear prebuckling effects, 114
Galerkin method, 106 110 spherical shells, 127, 231
Gaussian curvature, plates, 162 uniform external pressure, 41, 44, 58
General instability stiffened shells (also see Imperfection
Generation gap, 11 sensitivity-cylindrical), 10, 144
Geometric nonlinearity, 66 stable post-buckling behavior, 142, 143
Guidelines for future plate buckling tests: toroidal shell segments, 72 78
factors affection choice of model scale, 578 Imperial College London high stiffness 250
importance of boundary conditions, 580 ton test machine, 588 591
long or short plates, 578 in-situ ultrasonic scanning facility, 591
multi-bay panels, 580 rate of feed and platen displacement gages,
possibility of partial load transfer, 580 591
recommendations for future tests, 580 581 stiffness similar to specially rigid
recommendations for good practice, 579 machines, 588, 589
Impulsive normal loading on plates, 251
Hinge, 45, 92 93, 393, 549 Indirect methods for determination of
Homogeneous, 17, 43, 60, 64, 71, 98, 101, buckling loads in plates, 516 533, 538
103, 105, 109, 112, 218 220, 230, 279 Inelastic buckling, 89 92, 184, 281, 323,
Homologous, 221 223, 226, 254, 256 333, 402, 452, 466
Hooke’s law, 89, 224, 226, 229, 274 Inelastic effects, 131, 185, 232, 468
Hoop stress, 80, 434 Influence of imperfections in columns, 182
616 Subject Index

Initial curvature, 125, 182, 593 Prandtl’s lateral buckling experiments, 329,
Initial imperfections, 178, 451, 595 330
Initial out-of-flatness, geometric University of Illinois tests, 331
imperfections, 464, 509, 516, 581, 593 Lateral instability, 32, 36, 205, 215, 329, 401
deliberately introduced out-of-flatness, Lateral pressure, 57, 59, 75 76, 146,
490 491, 512 515 170 171, 429, 572
Initial-postbuckling analysis, 134 148 Lehigh gravity load simulator, 361 362
Initial postbuckling experiment, 187, Limit load, 8, 10, 147, 150, 160, 409, 416,
148 151 418 419
Initial postbuckling path, 422 Limit point, 48 50, 52, 81, 85, 88, 94 95,
In-plane buckling, 203, 440 121, 135, 147, 150, 173,
Intelligent instruments, 3 Load-deflection curve, 52, 155, 183, 208,
Interaction curve, 37, 233, 576 577 210, 246, 302, 374 375, 416, 432,
Interaction of local and general instability, 518 519, 536, 576
344 346 Load-deflection curves, 182 183, 185, 194,
Interactive buckling in columns and beams, 207, 210, 246 247, 303, 416 419, 433,
344, 345 457, 462
Cornell University early tests on mode Load-deflection data, 534
interaction, 344, 345 Load eccentricity, 149 150, 159 160,
Eindhoven University interactive buckling 182 183, 206 207, 210, 351, 421 425
tests, 352 356 Loading devices, 210, 294, 369, 420 421,
interaction between local buckling and 451
lateral torsional buckling in bending, deformation (screw-type) testing machines,
354 356 499
University of Sydney interaction tests on feedback computer control, 295, 302
columns, 346 351 predetermined displacement or loading
Johnson parabola, 289 rates, 295
Kinematic relations, 50 51, 67 68, 71, 161 pressure (hydraulic) testing machines, 295,
Kinematic stability criterion, 95 101 499
Knockdown factor (also see imperfection Loading systems, 230, 341, 383, 429 430,
sensitivity), 7 498 507, 546, 572, 574 576, 584, 587
Koiter circle, 54 “dead load” levered loading system, 504
Koiter postbuckling theory, 178 displacement-control loading, 499
Koiter theory of initial postbuckling analysis, flexible loading strips, 502
133 148 small test rigs with own loading system,
499
Laboratory scale (test), 7 Yamaki’s test setup and boundary
Large scale (test), 7, 377, 389, 510 conditions, 499 501
Large test-rigs, 503 504, 509 Local buckling, 24 26, 125, 185, 303 304,
Cambridge University plate test rigs, 309, 311 312, 314, 317 318, 327,
486 491 344 349, 351 354, 356, 365, 383, 403
ETH Zürich test-rigs, 504 Local elastic buckling, 234
relative stiffness of loading frames, 505 Lundquist plot, 197 198, 200, 202, 343, 524,
Lateral buckling of arches, 440, 452 526 528, 533
Lateral buckling of beams, 32, 39, 125,
203 204, 328, 402 Marriage of theory and experiment, 5
ALCOA experiments, 332, 333 Material nonlinearity (see nonlinear material),
application of the Southwell method, 81, 160
197 198, 214 Materials for buckling experiments, 230
Michell experiments, 330 Mathematical models, 3, 6, 94, 390
Nagoya University tests, 334 337 Mechanical model, 187, 189, 192 193
Subject Index 617

Modal interaction, mode interactions, 6, 28, Nonlinear stress-strain relation, 229


344 Numerical analysis 3, 127, 439
Model analysis, 217, 272 273, 278 279, Numerical methods, 15, 80, 84, 272 273
281 282, 284, 404
Araldite models and larger welded steel One way buckling, 434, 451
model, 279 281 Out-of-plane buckling, 368, 440, 443, 452
as a design tool, 272, 279 Ovalization (see Pipe-Brazier flattening), 44,
direct method, measuring stresses and 65, 84 88
deformations, 272, 533, 538
in vibration studies, 273 275 Photoelastic models, 281
of glass reinforced plastic minesweeper, Pipe, bending of, 53, 60, 64, 84, 179, 183,
275 279 353, 562
Model frame tests, 369, 370 Brazier flattening, 84 88
Cambridge University tests, 504, 509, Planar frames, 28 32
574 575 Plastic buckling, 88 94, 127 128
Polytechnic Institute of Brooklyn tests, 368 Plasticity, 5, 25, 125, 315, 464
University College London tests, 369 Plate buckling, 19, 141, 202 203, 316, 325,
Model laws, 221, 254, 256, 260 453 456, 458 460, 462, 464 466, 468,
Modeling, 217 218, 220, 222, 224, 226, 228, 470, 472, 474, 476, 478, 480, 482, 484,
230, 232, 234, 236, 238, 240, 242, 244, 486, 488, 490, 492 494, 496, 498, 500,
246, 248, 250, 252, 254, 256, 258, 260, 502, 504, 506, 508, 510, 512, 514 516,
262, 264, 266, 268, 270, 272, 274, 276, 518 519, 520, 522, 524, 526, 528, 530,
278, 280, 282 286, 404 532, 534, 536, 538, 540, 542, 544, 546,
dissimilar material, 223, 257 548, 550, 552, 554, 556, 558, 560, 562,
mathematical, 217 564, 566, 568, 570, 572, 574, 576 580,
physical, 217 582, 584, 586, 588, 590 592, 594, 596,
theory and practice, 217 598, 600
Modern plate tests, 502 effective width, 455 459
computer controlled adjustment of loading, Fairbairn’s and Hodgkinson’s tests,
584 453 456
Monash University Melbourne simple test historical background, 453
rig, 582 postbuckling behavior, 160 167
RWTH Aachen test rig for buckling of Plates, 124 125, 128, 161, 166, 251, 260,
composite plates, 584 588 285 286, 403, 470, 516, 570, 592 597,
clamped supports, 587 599, 600
loaded edges with ball or roller bearings, rectangular, 20 21, 179, 253, 460 462,
586 471, 473, 500, 507, 539, 572, 577,
simple supports with constant or variable 591, 599, 600
rotation, 585, 587 Plate under multiple loading, 570 577
More complete data recording, 196 biaxial in-plane compression tests, 570
Motivation for experiments, 5 Cambridge University biaxial compression
tests, 572 577
NACA Langley multiple-bay fixture, constant transverse load, 576
495 497 interaction diagrams, 577
Newton-Raphson method, 112 proportional loading, 576
Non-conservative loads, Poisson’s ratio, 120, 223 224, 227,
follower forces, 2, 99 230 233, 238, 240, 274, 521
configuration-dependent external loads, 2 Post bifurcation (see postbuckling)
Nonhomogeneity of material, 182, 294 Postbuckling, 6 7, 13, 21, 26, 81, 131 139,
Nonlinear bifurcation, 115 141 142, 144 145, 147, 151, 156, 160,
Nonlinear equilibrium equations, 81 165, 168 170, 178 179, 181, 185 186,
Nonlinear material, 89, 229 217, 229, 231, 276, 278 279, 294, 314,
618 Subject Index

318, 324, 330, 370, 376 377, 381, 392, 387 388, 426, 465, 467, 513, 515 516,
394, 403, 405, 458 459, 462, 464, 578, 593 594
499 500, 508 509, 534, 537 540, 542, due to cold-bending, 291
553, 578, 580, 588, 593, 595, 597, 601 due to flame-cutting, 291
imperfection sensitivity parameter, due to hot-rolling, 291
145 148 due to welding, 293, 465
stable, 168, 324 Residual stresses in plates
unstable, 421 welding stresses, 465
Postbuckling behavior, 3, 6, 9, 81, 131 134, Response of structures to blast loading,
136, 138, 140, 142 144, 146, 148, 150, 254 259
152, 154, 156, 158, 160, 162, 164, dissimilar material blast response scaling
166 170, 172, 174, 176 179, 181, 184, law, 254, 258
187, 209, 231, 237, 258, 272, 289, 369, elastic response, 256
431, 455, 459, 462, 484, 499, 509, 520, elastic-plastic response, 256
536, 540, 542, 593, 600 Hopkinson’s blast scaling law, 254
of columns, 136 139, 154, 156 160 large deflection response of cantilever
of cylinders, 143 154 beams, 256
of plates, 139 143, 160 167 replica response modeling for explosive
Postbuckling strength of plates, 164 forming, 256
Potential energy, 41 42, 46, 95 96, replica structural response, 256
103 104, 109, 112 113, 115, 119, Sachs’ scaling law, 255
121 123, 136, 154, 161 162, 169 small scale model tests, 256
minimum, 95 Reynold’s number, 219, 221, 226, 229
stationary, 41, 45, 51, 67, 137, 154, Rigid cavity, 439, 452
168 169 Rings, 215, 409 410, 412, 414, 416, 418,
Prebuckling, 43, 53 54, 57, 59, 60, 63, 66, 420, 422, 424, 426, 428, 430, 432, 434,
68, 70, 73 77, 80 85, 98, 104, 436, 438, 440, 442, 444, 446, 448,
108 109, 111 112, 115 116, 120, 450 452
126 127, 132, 134 135, 140, 142, 144, Ring buckling, 442
163, 168, 172, 330, 537 Ring-stiffened, 114, 126, 129, 178, 213, 231
deformations, 85 Role of experiments, 3
nonlinear prebuckling analysis, 147, 177
Prototype, 220 223, 226 231, 234, 236, Safety margin
238 240, 249 250, 254, 257 261, 263, Scale factors, 222, 238, 260, 263, 274, 284,
265, 268, 272, 274 275, 282, 383 386, 578
388, 396 397, 431, 578 Scale model testing for impact loading,
241 242, 245
Quadratic form, 97, 112, 128, 123 dynamic behavior of type I and type II
Quasi-shallow shells, 67 structures, 245 251
Scaling, 226, 230, 237, 240, 243, 247 248,
Radius of curvature, 157 250 251, 253 261, 263 264, 259,
Radius of gyration, 18, 195, 294, 306, 416 265 272, 274, 284 286, 386 387
Random imperfections, 151 for blast loading, 254
Rayleigh-Ritz method, 20, 102 106, 109, 204 for impact loading, 241
Rectangular flat plates, 470, 592 593, 596 of composite beams, 260 268
Redistribution of stress (see Stress of composite cylindrical shells, 270 272
redistribution), 164 of composite plates, 268 270
Replica modeling, 223, 254 156 of dynamically loaded structures, 258
Research type experimental programs, 6 of free vibrations, 238
Residual deformation, 329 Scaling rules for composites
Residual stresses, 3, 5, 9 10, 182, 185, 187, buckling can be scaled reliably:
279, 289 294, 299, 303, 316 317, 336, composite cylindrical shells, 270 272
Subject Index 619

laminated beams, 260 268 geometric, 220, 222 223, 226, 232,
laminated plates, 268 270 240 241, 254, 258
plates impacted by cylindrical projectiles, kinematic, 222 223
268 270 for buckling, 229
scaling of lamina, 259, 260 for shear panels, 234 237
strength and large deflections, 260 268 of mass distribution, 221
Secondary buckling, 459 460, 462 464, Simple buckling test, 181 82, 184 186, 188,
495, 498, 593 190, 192, 194, 196, 198, 200, 202, 204,
major changes in buckled form, 460 206, 208, 210, 212, 214
theoretical and numerical studies, 459 465 Slenderness ratio, column, 18, 24, 159 160,
Seismic loads on multi-storey frames, 377 195, 289, 306 307, 324, 351, 357, 386,
0.305 scale models, 383 385 547
Berkeley earthquake simulator, 383, 385 Smeared stiffeners (see Stiffened)
brace buckling, 381 Snap buckling, 416
computer on-line actuator (COLA) Snap-through (also see collapse), 114
pseudodynamic system, 379, 380 Southwell plot, Southwell’s method:
concentric braces (CBF), 384, 386 as nondestructive test method, 206
eccentrically braced dual system (EBF or expressed in strains, 200
EBDS), 382, 384 limitations of applicability, 207 209
eccentrically braced frames, 382, 384 “smoothing” of data, 204, 533
general conclusions on model testing, 388, for angle columns, 202, 203
389 for beam columns, 215
for columns, 206
half-scale model, 383
for frames, 203, 204
Stanford small-scale models, 388
for lateral buckling of beams, 205, 206
Tsukuba, Japan full-scale six storey frame,
for nonexperimental task, 212, 213
377, 378
for plates, 520 528
U.S./Japan cooperative research in
Space structures, 392
earthquake engineering, 377 392
buckling process in folding, 392
Semi-rigid, 210, 369, 371, 405, 421
combination of experiment and analysis to
Separation of variables, 74, 99 “remove” gravity effects, 396, 397
Shallow arch, 45, 49 50, 52, 409 410, 427 Olympus Astromast, 392, 394, 406
Shallow, 45, 69, 72 75, 114, 290, 431, 451, simulation of zero-g environment, 394
504 technique of deployment, 396
shell, 67 unfolding space frames, 392
Shear loading, flat plate subjected to, 21, Span, 46, 209, 279, 331 332, 335, 338 339,
542 543, 545, 549, 558, 560, 572 342, 453, 410 412, 416 417, 420, 422,
Shear panels, 234 236, 538, 540 542, 545, 424 425, 431
552 558, 597 598 Spherical body immersed in stream of
buckling and postbuckling behavior incompressible fluid, 218
civil engineering type plate girders Spherical caps, 69 72, 114 115, 177
diagonal tension, 546 552 Spherical shell buckling
Shells of revolution, 13, 66, 127, 129, 178 external pressure
Shortening of column, 300, 302 shallow spherical cap, 69 72, 114, 115
Sideways buckling, 409, 450 Spring balance, 420 422
Similarity, 2, 220 222, 226 227, 229, 231, Stable symmetric bifurcation, 191 193
234, 236 237, 239, 241, 251, 254, 256, Stability determinant, 105
258 259, 261, 274, 282, 284 285, 315, Stability equations, 43, 50, 52 54, 57, 60,
383, 385 62 63, 68 70, 73 74, 80 81, 84, 95,
complete, 221, 223, 226 97 98, 102, 104, 109, 140, 156
conditions, 222 223, 227 Stiffened plates, 540, 582, 583
dynamic, 223 Stiffened shells, 10, 143 147
620 Subject Index

Stiffener eccentricity, 143, 144 Windsor University tests on angle struts,


Stiffness matrix K, 123 124 324 326
Strain, 278, 284, 286 287, 365, 431, 537, Torsional instability of columns, 22 24, 322,
580 323
Strain energy, 85, 87, 112 113, 115, 122, Torsional rigidity, 34, 320, 441
162, 199, 400 Transverse normal stress, 428, 451
Strain energy, column, 85, 87, 112 113, 115, Transverse shear, 428, 451
122, 162, 199, 400 Trefftz criterion, 97, 103
Strain energy, plate, 85, 87, 112 113, 115, Tsien’s energy criterion, 412
122, 162, 199, 400 Tube (see Pipe)
Strain energy cylindrical shell, 87, 104, 112,
113 Uemura and Byon’s test setup for secondary
Strain number, 225 226, 229, 236 237 buckling, 460 463
Stress redistribution, 164 Understanding of buckling and postbuckling
Stress-strain curve, 89, 156 157, 228, 305 behavior, 6
Structural Stability Research Council (SSRC), Uniform pressure loading, 427, 431 432, 436
12, 290, 398, 405, 598 Upper buckling load, 416
Stub-column test procedure, 183
Swinging platform, 431, 432 Variation, 38, 65, 149, 159, 402, 517
Vibration correlation techniques, 6
Tangent modulus, 12, 18, 91 92, 305 306, Vibration, 127, 129, 286, 596
468 von Kármán columns, 182 183, 195
Technology transfer, 9 10 von Kármán column experiment, 182
Test procedures for columns, 290 303 von Kármán plate equations, 139
aligning specimen, 299 von Kármán-Tsien analysis of cylindrical
evaluation of test data, 303 shells, 167 169
instrumentation, 299, 300
preparing specimen, 299 Warping constant, 23
presentation of test data, 302 Warping functions
testing, 300 Warping of thin-walled open cross sections
Thames barrier gates, 279 381, 386 Warren truss, 209 213, 369
Thin-walled cold formed and welded Web crippling, 315, 558, 561 564, 567 568,
columns, 313, 315 598 599
University of Strathclyde column tests, cause of failure in tee and cross joints of
318 320 rectangular hollow sections (RHS),
Thin-walled open cross sections, 315 562
Thrust, 191, 409, 422, 425 426, 499 collapse mechanism of plastic hinges and
Tilt loading, 450 yield lines, 562
Torispherical shell buckling under: due to concentrated load, 561
external pressure, 72 76 due to patch load, 561
internal pressure, 78 80 local instability failure, web crippling, 561
Toroidal shells, 73, 78, 85 web buckling, 561
Torsion, cylindrical shell subjected to, 59, 85, web crippling under patch loading, a
126, 215, 287, 400 plastic buckling phenomenon, 561
Torsional constant, 23 web squashing, 561
Torsional-flexural buckling, 317, 320, Web crippling tests:
322 326, 400 401 bending interaction, 564
Cornell University experiments, 323, 324 concentrated load effects, 564
NACA tests, 323 four basic loading conditions, 568
Stanford University tests, 322, 524 EOF, end one-flange loading, 568 569
torsional buckling tests, 320, 337, 452 ETF, end two-flange loading, 568
Subject Index 621

IOF, interior one-flange loading, 568 Wrinkling (also see crippling), 309
ITF, interior two-flange loading, 568
patch loading tests, 564 Yield stress, 17, 21, 25, 88, 94, 159,
University of Missouri-Rolla experiments, 231 233, 241, 304, 390 310, 458, 516,
569, 570 549, 562, 566 567, 573
University of Sydney tests, 564 567 Yielding, 127, 416
University of Toronto tests, 567, 568 Young’s modulus, 16, 26, 89, 120, 227 228,
Wide column, 19 20, 55 230, 236, 238, 240, 273, 278, 289, 418
Wide column buckling, 55 57

You might also like