You are on page 1of 177

NEW ANALYSIS AND DESIGN PROCEDURES FOR

ENSURING GAS TURBINE BLADES AND ADHESIVE


BONDED JOINTS STRUCTURAL INTEGRITY AND
DURABILITY
By
Hsin-Yi Yen, Ph.D.
The Ohio State University, 2000
M. -H. Herman Shen, Adviser

Most load-carrying structural systems under severe operating conditions such


as gas turbine engines usually demand durability, high reliability, light weight, and
high performance. In turn, as it has been reported, a number of structural failures
have occurred in aircraft engines during development testing and operational service.
In order to prevent failures of turbine engines, the turbine blade vibration must be
attenuated to an acceptable level. To achieve this goal, the blade has to be provided
with higher damping, either externally or internally. The objective of this study
is to explore the feasibility of using a stress dependent magnetomechanical surface
coating material for enhancing high damping capacity on turbine blades. The results
show that a 2% or 4% of blade thickness free surface magnetomechanical coating
layer has a signi cant contribution to the damping enhancement and the reduction
1

of vibratory stresses at various low and high frequency vibration modes under either
non-rotating or rotating conditions.
Similar to the blade failure, the structural reliability and safety of the adhesive bonded joint, one of the most commonly used structural joint designs in the
aerospace industry, is also a serious concern of the aircraft design community. Adhesive joints easily become weaker due to environmental degradation and/or improper
manufacturing procedures. This often reduces structural durability and reliability
signi cantly. This motivates us to develop a new nite element tool/procedure
for assessing the interfacial disbonding mechanics of the single-lap joint with various imperfectly-bonded conditions in order to predict the adhesive bonded joint's
strength more precisely during its service period. According to these conclusions,
a new three-dimensional graphic mesh has been created to display the maximum
stress variations under di erent amounts and sizes of disbonded area. This new
procedure can be used as a basis for the development of a bonded joint reliability
prediction method and accept/reject inspection criteria.

NEW ANALYSIS AND DESIGN PROCEDURES FOR ENSURING


GAS TURBINE BLADES AND ADHESIVE BONDED JOINTS
STRUCTURAL INTEGRITY AND DURABILITY
DISSERTATION
Presented in Partial Ful llment of the Requirements for
the Degree Doctor of Philosophy in the
Graduate School of The Ohio State University
By
Hsin-Yi Yen, M.S.
*****
The Ohio State University
2000
Dissertation Committee:

M. -H. Herman Shen, Adviser


Gerald M. Gregorek
Rama K. Yedavalli
Chia-Hsiang Menq

Approved by

Adviser
Aeronautical and
Astronautical Engineering
Graduate Program

c Copyright by

Hsin-Yi Yen
2000

ABSTRACT

Most load-carrying structural systems such as gas turbine engines are usually
under severe operating conditions. This type of structure demands durability, high
reliability, light weight, and high performance. In turn, as it has been reported, a
number of structural failures have occurred in aircraft engines during development
testing and operational service. In order to prevent failures of turbine engines, it
becomes necessary to attenuate the turbine blade vibration to an acceptable level.
To achieve this goal, the blade has to be provided with higher damping, either
externally or internally. The objective of this study is to explore the feasibility of
using a stress dependent magnetomechanical surface coating material for enhancing
high damping capacity on turbine blades.
In the beginning, an analytical procedure is developed for modeling the dynamic behavior of beams and blades coated with a plasma sprayed iron-chromium
based coating. The stress or strain dependent damping capability of the coating
material is evaluated experimentally and further quanti ed with a probabilistic distribution function. The equations of motion and associated boundary conditions
are derived for uniform coated beams and blades. Later, the resulting equation of
motion is solved for beams and blades with a free surface coating layer by a closed
form procedure and nite element approximation. The results show that a 4% of
blade thickness free surface magnetomechanical coating layer can have a signi cant
ii

contribution to the damping enhancement and the reduction of vibratory stresses


at various low and high frequency vibration modes under either non-rotating or
rotating conditions.
Similar to the blade failure, the structural reliability and safety of the adhesive bonded joint, one of the most commonly used structural joint designs in the
aerospace industry, is also serious concern by its imperfectly-bonded interface. This
imperfectly-bonded interface, caused by the environmental degradation and/or improper manufacturing procedures, has weaker interface strength and often reduces
structural durability and reliability signi cantly. This motivates us to develop a
new nite element tool/procedure for assessing the interfacial disbonding mechanics of the single-lap joint with several imperfectly-bonded conditions under various
impulse loading conditions in order to predict the adhesive bonded joint's strength
more precisely during its service period.
In this study, a new disbonded interface model and a nite element procedure
have been developed to calculate the stress distributions in the adhesive joints with
perfectly-bonded or imperfectly-bonded interfaces under tensile loading. The nite element modeling of the weakened strength of the disbonded interface is accomplished through a spring element procedure or a newly developed line element
procedure. Those nite element procedures consist of a new modeling technique
for assessing the e ects of interfaces on the stress elds of adhesive joints during the degradation process. The results from these procedures determined the
strength reduction of the interface depends on the amount of disbonded area and the
size of each individual disbonded area. According to these conclusions, new threedimensional information has been created to display the maximum stress changes
iii

under di erent amounts and sizes of disbonded area. This new procedure can be
used as a basis for the development of a bonded joint reliability prediction method
and accept/reject inspection criteria.

iv

This is dedicated to my parents.

ACKNOWLEDGMENTS

I would like to acknowledge my adviser, Dr. M. -H Herman Shen for his guidance,
encouragement, support, and care during the course of this dissertation. As my
adviser, his suggestions and observations contributed immensely to the success of
this work.
I would like to acknowledge Dr. Gerald M. Gregorek, Dr. Rama K. Yedavalli,
and Dr. Chia-Hsiang Menq for serving on my dissertation committee and for various
suggestions for this dissertation.
Additionally, I also would like to thank Ms. Sandy Rhoads for helping my
paper works , and thank Mr. Tommy George for his valuable suggestions in this
dissertation.
Finally, I thank my parents for their support and encouragement and all my
friends at The Ohio State University for their assistance during the past four years.

vi

VITA

Feb. 12, 1970 ............................. Born - Taipei, Taiwan


1992 ......................................B.S. National Cheng Kung University,
Tainan, Taiwan
1992 ......................................Mechanic Trainee
Far East Air Transportation Co.
1992-1994 .................................Legislative Ocer
R. O. C. Army
1996 ......................................M. S. The Ohio State University
1996-present .............................. Graduate Research Associate
The Ohio State University
PUBLICATIONS
Research Publications

H. -Y. Yen and M. -H. Herman Shen \Passive Vibration Suppression of Turbine
Blades Using Magnetomechanical Coating". Journal of Sound and Vibration, under
revision process.
H. -Y. Yen and M. -H. Herman Shen \Development of a New Modeling Procedures
for the Imperfectly-Bonded Interface in Single-Lap Joints". International Journal
of Adhesion and Adhesives, under review.
H. -Y. Yen and M. -H. Herman Shen \A New Finite Element Modeling Technique
for Disbonded Adhesive Joints". to appear in 2000 ASME International Mechanical
Engineering Congress and Exposition, November 2000.
vii

H. -Y. Yen and M. -H. Herman Shen \Development of Passive Turbine Blade
Damper Using Magnetomechanical Coating". International Gas Turbine & Aeroengine Congress & Exhibition, 2000-GT-0366, May 2000.
H. -Y. Yen and M. -H. Herman Shen \Magnetomechanical Coating Applications in
Vibration Suppression of Turbine Blades". 5th National Turbine Engine High Cycle
Fatigue Conference, March 2000.
H. -Y. Yen and M. -H. Herman Shen \Passive Vibration Suppression of Turbine
Blades using Magnetomechanical Coating". 4th National Turbine Engine High
Cycle Fatigue Conference, February 1999.
H. -Y. Yen and M. -H. Herman Shen \Passive Vibration Suppression of Beams Using
Magnetomechanical Coating". Vibration and Noise Control, DE-Vol. 97/DSC-Vol.
65, ASME, 1998.
H. -Y. Yen and M. -H. Herman Shen \The E ects of Fatigue Cracks on Free
Torsional Vibration of Shafts". International Gas Turbine & Aeroengine Congress
& Exhibition, 97-GT-249, June, 1997.
FIELDS OF STUDY

Major Field: Aeronautical and Astronautical Engineering


Studies in Structures: Prof. Shen, M. -H.,Herman

viii

TABLE OF CONTENTS

Page

Abstract . . . . . .
Dedication . . . . .
Acknowledgments .
Vita . . . . . . . .
List of Tables . . .
List of Figures . .
Chapters:

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

1. INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2 Problem Review and Literature Search . . . . . . . . . . . . . .
1.2.1 Vibration Suppression of Turbine Blades . . . . . . . . .
1.2.2 The Modeling Technique of the Bonded Joint . . . . . .
1.3 Problem Approach . . . . . . . . . . . . . . . . . . . . . . . . .
1.3.1 Magnetomechanical Coating on Turbine Blade . . . . . .
1.3.2 The Modeling Technique for Imperfectly-Bonded Joints .

. ii
. v
. vi
. vii
. xii
. xiii
.
.
.
.
.
.
.
.

1
1
7
7
12
15
15
16

2. PASSIVE VIBRATION SUPPRESSION WITH MAGNETOMECHANICAL COATING MATERIAL . . . . . . . . . . . . . . . . . . . . . . .


2.1 A Brief Review of the Energy Dissipation Mechanism . . . . . . .
2.2 Damping Measurement of the Coating Material . . . . . . . . . .
2.3 Theoretical Development . . . . . . . . . . . . . . . . . . . . . . .

19
19
22
30

ix

2.3.1 Kinematic Assumptions . . . . . . . . . . . . . . . . . . . 30


2.3.2 Variational Theorem . . . . . . . . . . . . . . . . . . . . . 32
3. APPLICATION TO BEAMS AND BLADES . . . . . . . . . . . . . .
3.1 E ect of Coating on Forced Response of Simply Supported Beams
3.2 E ect of Coating on Forced Response of Cantilevered Beams . . .
3.3 E ect of Coating on Low Frequency Forced Response of Cantilevered Blades . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4 E ect of Coating on High Frequency Stripe Mode Forced Response
of Cantilevered Blades . . . . . . . . . . . . . . . . . . . . . . . .
3.5 The Coating Thickness E ect of Rotating Cantilevered Blades . .

37
38
42
49
61
73

4. DEVELOPMENT OF A FINITE ELEMENT PROCEDURE FOR ADHESIVE BONDED JOINTS . . . . . . . . . . . . . . . . . . . . . . . . 85


4.1 Finite Element Model for Single-Lap Joint . . . . . . . . . . . . . 86
4.2 Finite Element Model for Double-Strap Joint . . . . . . . . . . . 95
5. DEVELOPMENT OF IMPERFECTLY-BONDED INTERFACE MODELS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1 The Distributed Spring Element Model . . . . . . . . . . . . . . .
5.2 The One-Dimensional Line Element Model . . . . . . . . . . . . .
5.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . .
5.3.1 The Distributed Spring Element . . . . . . . . . . . . . .
5.3.2 The One-Dimensional Line Element . . . . . . . . . . . .
5.4 The E ect of Disbonded Island Number and Size for ImperfectlyBonded Joints . . . . . . . . . . . . . . . . . . . . . . . . . . . .

100
101
105
109
109
111
120

6. CONCLUSIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
Appendices:
A. Collocation method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
B. The Abbreviated MSC/NASTRAN Routine for the Cantilevered Blade
with the Magnetomechanical Coating . . . . . . . . . . . . . . . . . . . 140
x

C. Dynamics of the Rotating System . . . . . . . . . . . . . . . . . . . . . 144


D. DMAP Routine for the Centrifugal Force on the Rotating System . . . 147
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

xi

LIST OF TABLES
Table

Page

3.1 The Maximum De ection (in) of Simply Supported Beams. (FEM:


Finite Element Results ; CF: Closed-Form Results) . . . . . . . . . 40
3.2 The Maximum Stress (ksi) of Simply Supported Beams. (FEM: Finite Element Results) . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.3 The Maximum Deformation (in) of Cantilevered Beams. (FEM: Finite Element Results ; CF: Closed-Form Results) . . . . . . . . . . 43
3.4 The Maximum Stress (ksi) of Cantilevered Beams. (FEM: Finite
Element Results) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

xii

LIST OF FIGURES
Figure

Page

1.1 Air Force Engine Fatal Accident Rate Caused by Engine Failures. .

1.2 High Cycle Fatigue Failure in the Turbine Engine Blade [22]. . . . .

1.3 Boeing 737 Lap Joint Con guration. . . . . . . . . . . . . . . . . .

1.4 Boeing 737 Lap Joint Con guration (continued). . . . . . . . . . . .

1.5 Turbine Engine Life Cycle Diagram [1]. . . . . . . . . . . . . . . . .

1.6 The Friction Dampers in the Jet Engine Blades. . . . . . . . . . . .

2.1 The Domain Structure of the Coating Material (


: point into the
paper; : point out the paper). . . . . . . . . . . . . . . . . . . . . 20
2.2 The Steel Beam Used in the Damping Measurement . . . . . . . . . 24
2.3 The Coated Beam Vibration Experiment. . . . . . . . . . . . . . . . 25
2.4 The Frequency Response of the Steel Beam Vibration. . . . . . . . 26
2.5 The Coated Beam Vibration Experimental Setup. . . . . . . . . . . 27
2.6 The Two Steel Beams Coated with the Magnetomechanical Coating
and One Magnetomechanical Material Beam Before Machining. . . 28
2.7 The Steel Beam Coated with One Magnetomechanical Coating Layer 29
2.8 The Geometry of the Coated Beams. . . . . . . . . . . . . . . . . . 30
xiii

2.9 The Comparison of Curve Fitting and Test Results . . . . . . . . . 31


3.1 The Finite Element Model for the Blades Containing the Coating Layer 39
3.2 First and Third Mode Shapes of Uncoated and Coated Simply Supported Beams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.3 Stress Distributions in the First and Third Modes of Uncoated and
Coated Simply Supported Beams . . . . . . . . . . . . . . . . . . . 46
3.4 First Three Mode Shapes of Uncoated and Coated Cantilevered Beams 47
3.5 Stress Distributions in the First Three Modes of Uncoated and Coated
Cantilevered Beams . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.6 The Finite Element Mesh of Turbine Blade . . . . . . . . . . . . . . 53
3.7 Haigh (or Goodman) Diagram for Ti-6Al-4V . . . . . . . . . . . . . 54
3.8 The von Mises Stress in the First Mode Vibration of the Uncoated
Cantilevered Blade . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
3.9 The von Mises Stress in the First Mode Vibration of the Coated
Cantilevered Blade . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
3.10 The von Mises Stress in the Second Mode Vibration of the Uncoated
Cantilevered Blade . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
3.11 The von Mises Stress in the Second Mode Vibration of the Coated
Cantilevered Blade . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.12 The von Mises Stress in the Third Mode Vibration of the Uncoated
Cantilevered Blade . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
3.13 The von Mises Stress in the Third Mode Vibration of the Coated
Cantilevered Blade . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
3.14 The Third Stripe Mode Shape of the Cantilevered Blade . . . . . . 66
3.15 The von Mises Stress in the Second Stripe Mode Vibration of the
Cantilevered Blade . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
xiv

3.16 The von Mises Stress in the Second Stripe Mode Vibration of the
Coated Cantilevered Blade . . . . . . . . . . . . . . . . . . . . . . 68
3.17 The von Mises Stress in the Third Stripe Mode Vibration of the
Cantilevered Blade . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
3.18 The von Mises Stress in the Third Stripe Mode Vibration of the
Coated Cantilevered Blade . . . . . . . . . . . . . . . . . . . . . . 70
3.19 The von Mises Stress in the Fourth Stripe Mode Vibration of the
Cantilevered Blade . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.20 The von Mises Stress in the Fourth Stripe Mode Vibration of the
Coated Cantilevered Blade . . . . . . . . . . . . . . . . . . . . . . 72
3.21 The Rotating Diagram of the Coated Blades . . . . . . . . . . . . . 76
3.22 The Third Stripe Mode Shape of the Cantilevered Blade about Rx
under 10000 rpm Rotation . . . . . . . . . . . . . . . . . . . . . . . 77
3.23 The Fourth Stripe Mode Shape of the Cantilevered Blade under 10000
rpm Rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
3.24 The von Mises Stress of the Fourth Stripe Mode of the Cantilevered
Blade Rotating about Rx under 10000 rpm . . . . . . . . . . . . . . 79
3.25 The von Mises Stress of the Fourth Stripe Mode of the Coated Cantilevered Blade Rotating about Rx under 10000 rpm with 2% Thickness Coating Layer . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
3.26 The von Mises Stress of the Fourth Stripe Mode of the Coated Cantilevered Blade Rotating about Rx under 10000 rpm with 4% Thickness Coating Layer . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
3.27 The von Mises Stress of the Fourth Stripe Mode of the Cantilevered
Blade Rotating about Ry under 10000 rpm . . . . . . . . . . . . . . 82
3.28 The von Mises Stress of the Fourth Stripe Mode of the Coated Cantilevered Blade Rotating about Ry under 10000 rpm with 2% Thickness Coating Layer . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
xv

3.29 The von Mises Stress of the Fourth Stripe Mode of the Coated Cantilevered Blade Rotating about Ry under 10000 rpm with 4% Thickness Coating Layer . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
4.1 The Geometry and Finite Element Mesh of the Bonded Joints . . . 90
4.2 Adhesive Stress Distributions Obtained from the Present Finite Element Model with Linear Geometry (Uniform Loads)(GR: Goland
and Reissner; OP: Oplinger; HS: Hart-Smith; FEM: Finite Element
Method )(y=0) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
4.3 Adhesive Stress Distributions Obtained from the Present Finite Element Model with Nonlinear Geometry (Uniform Loads)(y=0)(GR:
Goland and Reissner; OP: Oplinger; HS: Hart-Smith; FEM: Finite
Element Method) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
4.4 Adhesive Stress Distributions Obtained from the Present Finite Element Model with Nonlinear Geometry (Single Loads)(y=0)(GR:
Goland and Reissner; OP: Oplinger; HS: Hart-Smith; FEM: Finite
Element Method ) . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
4.5 Adhesive Stress Distributions Obtained from the Present Finite Element Model with Nonlinear Geometry (Uniform Loads)(y=0) . . . 94
4.6 Adhesive Stress Distributions Obtained from the Present Finite Element Model for Double-Strap Joint (OP: Oplinger; HS: Hart-Smith;
FEM: Finite Element Method) . . . . . . . . . . . . . . . . . . . . . 98
4.7 Adhesive Stress Distributions Obtained from the Present Finite Element Model Compared with TJOINTNL Program (OP: Oplinger;
HS: Hart-Smith; FEM: Finite Element Method) . . . . . . . . . . . 99
5.1 The Geometry and Finite Element Model of the Imperfectly-Bonded
Joints . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
5.2 Adhesive Stress Distributions Obtained from the Present Finite Element Model with Perfectly-Bonded Interface (Uniform Loads) . . . 116

xvi

5.3 Adhesive Stress Distributions Obtained from the Present Finite Element Model with 20% (Ad=0.2) Imperfectly-Bonded Interface by
Using the Spring Element (Uniform Loads) . . . . . . . . . . . . . . 117
5.4 Adhesive Stress Distributions Obtained from the Present Finite Element Model with 20% (Ad=0.2) Imperfectly-Bonded Interface by
Using the Line Element (Uniform Loads) . . . . . . . . . . . . . . . 118
5.5 Adhesive Stress Distributions Obtained from the Present Finite Element Model with Various Disbonded Areas by Using the Line Element.(30%
(Ad=0.3) near the End of Overlap Decreases to 0% (Ad =0) in the
Center) (Uniform Loads) . . . . . . . . . . . . . . . . . . . . . . . . 119
5.6 The Strength of the Interface with Various Disbonded Island Sizes . 121
5.7 Maximum Peel Stress Distributions at the Upper Level (y = t ) of
the Bonded Interface along the Percentage of Disbonded Area and
Disbonded Island Size (inches) . . . . . . . . . . . . . . . . . . . . . 128
2

5.8 Maximum Peel Stress Distributions at the Middle Level (y = 0) of


the Bonded Interface along the Percentage of Disbonded Area and
Disbonded Island Size (inches) . . . . . . . . . . . . . . . . . . . . . 129
5.9 Maximum Peel Stress Distributions at the Lower Level (y = t ) of
the Bonded Interface along the Percentage of Disbonded Area and
Disbonded Island Size (inches) . . . . . . . . . . . . . . . . . . . . . 130
2

5.10 Maximum Shear Stress Distributions at the Upper Level (y = t ) of


the Bonded Interface along the Percentage of Disbonded Area and
Disbonded Island Size (inches) . . . . . . . . . . . . . . . . . . . . . 131
2

5.11 Maximum Shear Stress Distributions at the Middle Level (y = 0) of


the Bonded Interface along the Percentage of Disbonded Area and
Disbonded Island Size (inches) . . . . . . . . . . . . . . . . . . . . . 132
5.12 Maximum Shear Stress Distributions at the Lower Level (y = t ) of
the Bonded Interface along the Percentage of Disbonded Area and
Disbonded Island Size (inches) . . . . . . . . . . . . . . . . . . . . . 133
2

xvii

LIST OF SYMBOLS
Ad
E
Eab
Ec
Es
Ea
Eb
Fint
F (x)
FN
Ft
G
I
Is
Kj
K
K 0
Kt
Kn

Area Ratio between Disbond and Bonded Area


Young's Modulus of Elasticity (coated beam)
E ective Young's Modulus of Elasticity for Adhesive Joint
Young's Modulus of Elasticity (coating material)
Young's Modulus of Elasticity (steel)
Young's Modulus of Elasticity for Adherends
Young's Modulus of Elasticity for Adhesives
Inertial Force
Excitation Forcing Function
Normal Force per Unit Length
Shear Force per Unit Length
Shear Modulus of Elasticity (coated beam)
Moment of Inertia (coated beam)
Moment of Inertia (uncoated beam)
j th Sti ness Component of the Bonded Interface
Magnetic Parameters
K=2
Shear Sti ness of the Bonded Interface
Normal Sti ness of the Bonded Interface
xviii

Kr
Ks
M
N
Uc
Ud
Uj
a
bl
b

2Cb
2d
db
h
l
lb
m
s
t
ux
u
u; v; w
y1 ; y2

The Sti ness Matrix for the Rotating System


The Sti ness Matrix for the Stationary System
Mass of the Turbine Blade
Number of Distributed Spring Elements
Strain Energy Density (coating material)
Strain Energy (coating material)
Local Strain of the Bonded Interface
The Crack Length of the specimen
The Interface Length of the specimen
Width of the Beam
Length of the Overlap
Thickness of the Uncoated Beam
Length of the Line Element
Thickness of the Coating Layer
Length of the Beam
Length of the Outer Adherends
Mass of the Bonded Interface
i =  i =
Thickness of Adherends
Normal Displacement of the Bonded Interface
Nodal Displacement
Coordinates for the Three-Dimensional System
Coordinates of two ends of the line element
Displacement of the Bonded Interface
xix

K
Uc
b

c
s



i
j
loc
l
cr


i


a
b
x
!

The Sti ness E ect from the Centrifugal Force


Energy Dissipation Density (coating material)
Thickness of the Adhesive Bonded Layer
Loss Factor (coated beam)
Loss Factor (coating material)
Loss Factor (uncoated beam)
Rotating Speed
Density (coated beam)
Vibratory Stress
Average Internal Stress
j th Component of Force Per Unit Length in Adhesive Joint
Local Internal Stress Barrier
Local Internal Stress
Critical Stress of the Saturation Point

t
Vibratory Strain
Average Internal Strain
Shear Stress of the Adhesive Bonded Joint
E ective Poisson's Ratio
Poisson's Ratio for Adherends
Poisson's Ratio for Adhesives
Shear Deformation of the Bonded Interface
Free Vibration Natural Frequency
xx

CHAPTER 1
INTRODUCTION
1.1 Background

Many load-carrying structural systems such as aircraft gas turbine engines are
typically under severe operating conditions. This type of structure demands durability, high reliability, light weight, and high performance. Traditionally, lifetime
failure-free design criteria based on the Goodman Diagram and Miner's Rule have
been adopted by the aircraft engine design community for ensuring safety of critical
structural components.
These design criteria, design guides, or design codes are often established using
the results of a simple deterministic analysis procedure without taking into account
information such as degradation of material properties, scatter in test data, previous successful design experience, and uncertainties inherent in real world operating
conditions. In turn, as has been reported, a number of structural failures have
occurred in aircraft engines during development testing and operational service.
Among those failures, the Sioux City incident is one of the most famous examples.
In 1989, United Airlines ight 232 enrout from Denver to Chicago crash landed
in Sioux City, Iowa, due to the defected fan disk failure of its No. 2 engine. In
1

Figure 1.1: Air Force Engine Fatal Accident Rate Caused by Engine Failures.
this incident, the failed disk broke the turbine blades which rendered all three hydraulic systems on board inoperative. This made the DC-10 almost uncontrollable
and resulted in a crash landing killing 111 of 296 passengers and crew members.
These incidents triggered an awareness of the fact that although current aircraft
engine critical structural components satisfy the lifetime failure-free design criteria
they sometimes fail, blading systems in particular. Although the Sioux City incident was caused by a material defect during the manufacturing process, high cycle
fatigue of the blades in jet engines is often the major concern in aviation safety,
2

especially in high performance military jet, as shown in Fig. 1.1. High cycle fatigue directly causes blade cracking, which increases maintenance and inspection
costs, reduces operational readiness, and sometimes even results in the loss of the
aircraft and crew members on broad. In the past one year alone, the U. S. Air
Force spent roughly one third of its total maintenance expenses on the high cycle
fatigue incidents [25]. According to their records, there are at least two F-16 ghter
crashes related to high cycle fatigue incidents within the last twelve months. One
was caused by a catastrophic failure in high pressure turbine assembly when two
turbine blades separated due to the high cycle fatigue, as shown in Fig. 1.2. The
other incident was caused by rst-stage compressor failure when one compressor
blade broke away due to high cycle fatigue. In this incident, an ultrasound inspection test had been performed on that particular engine, but still failed to detect
the developing cracks. Until recently, a newly designed U.S. navy F/A-18E Super
Hornet ghter was grounded due to the compressor blade failure of its new 22; 000
lb thrust GE F414 engine. After a full examination, fatigue cracks were found near
the tip of two compressor blades. A root cause analysis conducted by the U.S.
Navy pinpointed the cause of this cracking problem which was primarily due to the
high cycle fatigue. This cracking problem is directly caused by the extreme maneuvering the aircraft involved during its regular testing procedure. Unfortunately,
extreme maneuvering is one of the activities that military ghters cannot avoid
during daily training missions. This makes avoiding high cycle fatigue problems the
priority mission in the new integrated high performance turbine engine technology
(IHPTET) for the next generation Joint Strike Fighter (JSF). This project has been
corresponded by all the current major engine research groups, including Air Force,
3

Figure 1.2: High Cycle Fatigue Failure in the Turbine Engine Blade [22].
Navy, NASA, GE Aircraft Engine, Allison Advanced Development, Pratt-Whitney,
Williams International, Teledyne Ryan Aeronautical, Allied Signal Aerospace, and
a few smaller companies as well as other government agencies. Therefore, preventing turbomachinery blade failures caused by high cycle fatigue is one of the major
objectives of current aircraft engine design and in-service maintenance.
Similar to the blade failure, the structural reliability and safety of the environmentally degraded adhesive bonded joint, one of the most commonly used structural
joint designs in the aerospace industry, is also a serious concern of the aircraft design
4

community. The bonded joints have been widely used in Boeing 737, 747, 757, and
767s as bonded tear strap designs, shown in Fig. 1.3 and Fig. 1.4, or in common
airplane fuselage windows and doors as bonded doubler designs.

Figure 1.3: Boeing 737 Lap Joint Con guration.


During their operational periods, the bonded joints need to resist thousands of
signi cant loading cycles caused by the commercial airliner pressurized cockpits. In
addition to these internal airplane structural designs, adhesive bonding of composite parts is now widely used in the vertical stabilizers of Boeing 777 airplanes, the
5

Figure 1.4: Boeing 737 Lap Joint Con guration (continued).


leading edge of the horizontal stabilizer of the Boeing 767, and even the newly designed Sikorsky S-92 helicopter. These new bonded joint designs are mechanically
equivalent to, or even stronger than conventional assemblies, but built at lower cost
and weight. Also, adhesive bonded joints can avoid the low cycle fatigue problem caused by traditional metal joints, because the adhesive bonding layer has the
ability to absorb energy. Successfully reducing the manufacturing cost and eliminating rivet-induced fatigue problems, the adhesive bonded joint is easily to be
selected by the engineers as the primary method of fastening structural members.
However, the interface degradation, caused by environmental factors or improper
manufacturing procedures, leads to weaker interface strength and often reduces the

adhesive joint's lifetime tremendously. Recently, the French Civil Aviation Directorate issued a warning about a potential problem related to the weaker strength
of adhesive bonds in A-320 series and A330-200 composite vertical stabilizers. This
alarmed the aerospace industry of the potential reliability and safety problem of
adhesive bonded joints. Unfortunately, the current non-destructive inspection technologies, such as ultrasonic wave and X-ray, have diculties to monitor the quality
of the bonded joints. Recently, the composite honeycomb liquid hydrogen tank of
the NASA X-33 delaminated during testing when the adhesive bonds appeared to
be in perfect condition under ultrasonic test. This leads the urgency to develop a
method or technology in order to monitor and detect the imperfectly-bonded (poor)
interface of the adhesive bonding joints.
1.2 Problem Review and Literature Search
1.2.1 Vibration Suppression of Turbine Blades

These failures of turbine blades generally result from high vibratory or alternating stresses in various stage turbomachinery blades during resonant response.
Resonant response occurs, in a non-uniform ow eld operating environment when
the excitation frequency from unsteady aerodynamic loading is coincident with any
blade's natural frequency. Hence, an adequate blade design would exist if all the resonant conditions could be avoided. Realistically, it is impossible due to the fact that
insucient real world loading information is available at the time of the analysis
and design.
To prevent blade failure, the maximum stress of the excited resonant response
needs to be attenuated to an acceptable level through the control of blade vibration.
7

As an example in Fig. 1.5 obtained from Pratt-Whitney, the engine life cycle drops
tremendously when the maximum vibratory stress slightly increases [1]. Among
the solutions to reduce blade vibratory stress, adding damping is the most attractive and achievable method. In the last thirty years, several investigators have
presented di erent approaches to suppress blade vibratory stress by providing additional damping through blade dampers. For example, dry friction dampers [15]
which include blade-to-ground, blade-to-blade, and shroud dampers are the most
common vibration suppression devices employed by aircraft engine designers, shown
in Fig. 1.6.

Figure 1.5: Turbine Engine Life Cycle Diagram [1].


8

F-100-PW-2200 Engine

Shrouded Friction Damper

Base Friction Damper

Stator Row
Friction Damper

Figure 1.6: The Friction Dampers in the Jet Engine Blades.


However, friction dampers have their limitations. These dampers are based on
the relative motion of components, which causes energy dissipation via frictional effects. Therefore, friction dampers often work well for low frequency blade vibration
modes in which displacements are relatively high, but are much less e ective for the
small displacements at high frequency blade vibration modes, especially in the high
stripe modes, in which there are almost no or very small displacements near the base
of the blades. Another drawback for the friction damper is that new design turbine
9

blades no longer have the shrouds in order to improve their aerodynamic performance, which even further reduces the installation location of traditional friction
dampers. At the same time, the small displacements of the high frequency vibration modes make aerodamping also relatively small. Therefore, since the structural
damping from the dry friction dampers as well as the aerodamping are negligible
at high frequency vibration modes such as second stripe, third stripe, and fourth
stripe vibration modes, the remaining damping of the blades is solely from the energy dissipation from the material. Unfortunately, modern low aspect ratio blades
frequently vibrate like this plate-like stripe modes resulting in higher modal density
in the engine operating range. Consequently, the low material damping results in
high vibratory stress, increased failure risk, and signi cantly reduced reliability and
safety [29]. This has motivated the recent research activities for the development
of high frequency dampers.
Recently, numerous studies [14, 19] have been undertaken regarding the integration of viscoelastic damping materials into rotating blades for the purpose of
reducing vibratory stresses in high frequency modes. In this viscoelastic patch, the
additional vibratory energy dissipation is accomplished through high internal friction of the viscoelastic material patches inserted into milled cavities, which are then
sealed with a coversheet to maintain the structural integrity and the original airfoil
contour. However, this internal constrained layer damping treatments are not be
widely used because the viscoelastic material creep under the high loads and its
temperature limitation.
Recently, another interesting magnetic vibration suppression system has been
introduced by Ho man [18]. This magnetic damping system contains three groups
10

of electromagnets embedded near the blade tip. Each electromagnet is powered and
controlled by one central electronics module. The module can turn on the magnetic
elds when the rotor is at a known critical speed or passing through a vibration
mode. By powering the electromagnets, the energy from the blade vibration can
be dissipated through the eddy currents caused by the additional magnetic elds
from the electromagnets. In addition to the temperature limitation problem, adding
damping patches to blades with either internal cavities or internal electromagnets
also has manufacturing and durability considerations, especially with the cooling
lm design in the modern turbine blade. To avoid those diculties, a surface high
damping layer, such as a high damping coating layer, is still likely to be more
practical.
Cross, Lull, Newman, and Cavanagh [10] presented an alternative damping
method, using several candidate materials as graded plasma coating on aluminum
blades for accomplishing higher structural damping. Their test data demonstrated
the use of three to six layers of coating, such as magnesium aluminate, molybdenum, and Hastelloy-X, to enhance the structural damping of the blades. The
coating consists of an outermost layer portion formed of an oxdie ceramic or refractory carbide an intermediate portion formed of a mixture of one of the above
alloys, and the material forming the outermost layer. However, the graded plasma
coating materials introduced in this study are used only on the aluminum blades,
and it is still questionable about their high damping capacity when operating under
high temperatures and high rotating speed environments. It becomes necessary to
consider another alternative coating method to suppress the blade vibration.
11

1.2.2 The Modeling Technique of the Bonded Joint

The bonded adhesive joint is one of the most commonly used structural joint
con gurations. Therefore, the stress analysis of the adhesive joint, especially for
the single-lap joint, has been in development for more than a half century. Goland
and Reissner [13] developed a stress analysis which models the single-lap joints
as bending plates. In their analysis, the stress distribution of a single-lap joint
was formulated in accordance with the joint geometries and material properties
of the adherend and adhesive. Based on Goland and Reissner's assumption, the
eccentricity of the loading path creates a bending moment, the so-called joint edge
moment. This bending moment has a major e ect on the stress distributions of
the overlap of the adhesive joint. Furthermore, Goland and Reissner acknowledged
the critical role of the transverse normal (peel) stress near the ends of the adhesive
layer.
Later, Hart-Smith [16] improved the Goland and Reissner analysis by treating
the adherends as elastic beams, and calculated the stress distributions for linear elastic and elastic-plastic adhesives. Therefore, the Hart-Smith analysis is not restricted
by the lumped overlap which is assumed in the Goland and Ressiner analysis, and
provides more accurate linear elastic or elastic-plastic solutions. Recently, Oplinger
[26] proposed a more realistic layered beam model by considering the in uence of
the large de ection of the overlap, so called the geometrical nonlinear e ect.
In order to further understand the stress distribution along the thickness of
the adhesive and to ensure the boundary conditions are satis ed at the free ends,
several researchers have proposed two-dimensional analytical solutions. Chen and
12

Cheng [8] rst computed the stress distribution at the overlap of the single-lap
joint by using the variational principle of complementary energy. In their study, a
more realistic stress distribution in the adhesives was obtained. However, due to
the calculation diculties in Chen and Cheng's analysis, the longitudinal normal
stresses were assumed to be zero and the moment and shear force at the end of
the overlap are treated as external forces obtained by the Goland and Reissner
formulations.
In addition to Chen and Cheng's work, Barthelemy, Kamat, and Brinson [4]
develop a two dimensional nite element program called STAP to analyze a singlelap joint. After testing several di erent two-dimensional shell elements, an eightnoded isoparametric element was chosen to be used in their program, and proved to
be relatively accurate in comparison with their experimental data. Later, Lang and
Mallick [20] also created a two-dimensional nite element analysis to investigate
the recessed bonded joints. In their investigation, the stress distributions at the
adhesive mid-thickness and interface were determined for joints with various levels
of recessing, and compared to a continuous single-lap joint solution. More recently,
Tong [33] developed another two-dimensional nite element procedure to simulate
the relationship between surface displacement and adhesive peel stress in bonded
double-lap joints. However, in spite of tremendous numbers of papers related to the
perfectly-bonded structure analysis, the mechanics of imperfectly-bonded joints,
the common practical situation, are still not been studied. Therefore, the aerospace
and airline industry has begun to question the imperfectly-bonded joints' safety and
reliability.
13

First in 1984, Bail and Thompson [3] developed a quasi-static model for the
ultrasonic transmission and re ection at the imperfectly-bonded interface. In their
work, the imperfectly-bonded interface is assumed to be represented by a distributed
spring. This spring sti ness has been derived from the solutions of fracture mechanics [32] for the elastic displacement of materials containing cracks and inclusions under static loads. Later, Margetan, Thompson, and Gray [23] extended this
quasi-static spring model to represent the ultrasonic re ectivity of an imperfectlybonded interface as a function of frequency and angle of incidence. Those results
were then incorporated in a new model for the corner re ection from a di usionbonded joint between two plates. In their study, the distributed spring model used
for modeling the cracks can represent the imperfectly-bonded interface relatively
well. In addition, Lavrentyev and Rokhlin [21] further applied this distributed
spring model in several adhesive bonded joint non-destructive evaluations to simulate the environmental degradation. The distributed spring model in Lavrentyev
and Rokhlin's research has been used as the transverse spring boundary conditions
in order to simulate the ultrasonic wave re ection and the frequency shift caused by
the imperfectly-bonded interface lled by absorbed water. Unfortunately, the previous three studies are restricted to using the distributed spring model to simulate
the ultrasonic characterization and interactions, such as ultrasonic amplitude reduction and re ection frequency shift of the imperfectly-bonded interface. In turn,
the stress and strength changes in the disbonded joints caused by the environmental degradation still can not be predicted, which increase the safety concerns and
inspection diculties.
14

1.3 Problem Approach


1.3.1 Magnetomechanical Coating on Turbine Blade

In this study, the objective is to explore the feasibility of using a magnetomechanical surface coating material [7] for enhancing the high frequency damping of
turbine blades. The additional damping is achieved through internal friction via
the magneto-elastic e ects caused by stress or strain-induced irreversible movement
of magnetic domain walls. According to the domain theory [9], every ferromagnetic
material consists of so-called domain walls which are more or less randomly oriented
in an unmagnetized material. The domain walls of the coating material rotate and
generate a higher magnetic eld strength under external loading, because the domain walls tend to align themselves in the direction of the stress or stain. As the
loading is removed, the domain walls rotate to a di erent pattern which corresponds
to a lower magnetic eld strength. This movement of the domain walls results in an
irreversible magnetostrictive strain. During this process, it has been observed and
shown to produce a signi cant amount of magnetomechanical hysteresis energy loss
as a result of the magnetomechanical e ect.
The strain or stress level dependence of the magnetomechanical damping capacity or loss factor has been experimentally determined in Chapter 2, which in general
is to linearly reach a pronounced peak and then decay slowly to its initial value as
the stress or strain increases. A theoretical foundation is laid down to characterize the energy dissipation due to the irreversible movement of magnetic domain
boundaries of the coating material. Following the energy dissipation result, an analytical model for a uniform beam coated with a plasma sprayed magnetomechanical
15

layer is presented by using a generalized variational principle extended from the


Hu-Washizu principle [34] to derive the governing equations.
In Chapter 3, a closed-form procedure [11] from the previous Chapter 2 as well
as a nite element procedure are then applied to predict the forced responses of the
coated beams, for simply supported and cantilevered con gurations. The results
show that the nite element solutions are found to agree well with those closedform solutions.
Additionally, a nite element approach on resonant response of coated blades
is presented. Four-node isoparametric and eight-node solid elements are used to
model the coating layer across the blade's outer pressure side and the blade, respectively. Due to the stress or strain dependent damping, the present nite element
approach requires one to formulate the damping matrix in accordance with the
strain distribution.
1.3.2 The Modeling Technique for Imperfectly-Bonded Joints

The development of a basic understanding of the e ects of imperfectly-bonded


interfaces becomes urgent, when the structural reliability and safety of the adhesive joints is jeopardized. Therefore, in this study, a new nite element procedure
is developed for assessing the interfacial disbonding mechanics of single-lap and
double-strap adhesive joints under various impulse loading conditions.
In Chapter 4, the procedure involves the development of an optimal nite element mesh using Quad-8 isoparametric elements and taking into account geometrical and material nonlinearities. The accuracy of the nite element model has been
veri ed by comparison with the theoretical solution of Goland and Reissner [13],
16

Hart-Smith [16], and Oplinger [26]. The optimal nite element mesh was then used
as a basis in Chapter 5 for the development of a modeling technique for predicting
the e ects of disbonded interfaces in the adhesive joints.
In Chapter 5, the disbonded interface modeling technique is accomplished by
introducing a customized spring element as the connections between the adhesive
bonds and adherends to simulate the weaker imperfectly-bonded interface. The
sti ness of the spring elements represent the strength of the imperfectly-bonded
interface and depend on the percentage of disbonded area to the overall interface
area under a xed individual disbonded island size.
Additionally, in this study, a new one-dimensional six-noded line element has
also been developed and integrated into the procedure for assessing the damage
progress of the imperfectly-bonded interface. Similar to the spring element model,
the line element also simulates the strength changes of the disbonded interface by
modifying the sti ness matrices in the nite element procedure. The advantage of
this line element approach is that the imperfectly-bonded joint problems can be
solved directly with the modi ed sti ness matrices without adding more boundary
conditions like the distributed spring element procedure. This improvement with
the line element model provides a better ability to predict the disbonding mechanics
than the spring model.
By using the previous results, a three-dimensional graphic mesh has been created
to represent the maximum stresses of the imperfectly-bonded interface change under
di erent disbonded island size and disbonded area in the last part of Chapter 5.
The purpose of this mesh is to understand the e ects of the disbonded island size
and the disbonded area on the quality of imperfectly-bonded joints. Furthermore,
17

by using this information, the safety and reliability of imperfectly-bonded joints can
be determined during the environmental degradation. Therefore, these results can
be used as a design or inspection criteria to predict the bonded joint reliability and
create an accept/reject inspection procedure for the bonded joints.

18

CHAPTER 2
PASSIVE VIBRATION SUPPRESSION WITH
MAGNETOMECHANICAL COATING MATERIAL
2.1 A Brief Review of the Energy Dissipation Mechanism

Before investigating the magnetic energy dissipation mechanism, the physics of


magnetic energy has to be studied rst. Magnetic energy stored in a ferromagnet causes the magnetization of all magnetic atoms to be parallel. Magnetic poles
are induced on both ends, and increase the strength of the magnetizing eld. According to Domain Theory [7], all ferromagnetic materials are composed of many
small domains or magnets. Each domain consists of many atoms. The specialty
of this domain is that it can spin freely and usually is oriented with respect to
the crystal lattice of the material. The direct result of this is that the magnetic
energy can be increased by straining the material isotropically about the direction
of magnetization, shown in Fig. 2.1. By using this magnetic-elastic relation, the
external strain can orient and deorient the domain, and cause the domain walls to
rotate. However, this process requires higher strain energy to orient the domain
than energy released during its deorienting process. This movement creates an energy dissipation similar to the hysteresis loop caused by other damping materials
energy loss. This energy dissipation mechanism due to the irreversible movement
19

Demagnetized

Saturated (Domain Rotated in High Magnetic Field)

Crystal Axes

Magnetic Field

Figure 2.1: The Domain Structure of the Coating Material (


: point into the
paper; : point out the paper).
of magnetic domain boundaries, has been studied by Smith and Birchak [30, 31].
They divided the energy dissipation density per cycle, Uc, into two basic regions,
each associated with a stress range de ned as

Uc = K


if

 < cr

(2.1)

Uc = Kloc

if

 > cr

(2.2)

loc

20

where K is a constant depending on the shape of the hysteresis loop,  is the


saturation magnetostriction,  is the vibratory stress, loc is the local internal stress
barrier, and cr is the stress corresponding to the saturation point of the domain
walls irreversible movement.
According to Rayleigh's law [7], K , , loc, and cr are constants. Therefore,
from Eq. 2.1 and Eq. 2.2, the energy dissipation keeps increasing while the vibratory stresses become higher, as shown in Eq. 2.1. However, the energy dissipation
becomes a constant, as in Eq. 2.2, when the vibratory stress reaches the saturation
level, which means the magnetic domain has already reached its highest possible
magnetic eld strength.
In order to sum the contributions from these two di erent regions, a probabilistic
distribution function satisfying the previous two limiting forms of behavior de ned
in Eq. 2.1 and Eq. 2.2 has been introduced as follows:
N ( ) = (

4 )e

i

(2.3)

2
i

where i is the average internal stress.


Later, using this probabilistic distribution given by Eq. 2.3, Smith and Birchok
[30, 31] reconstructed Eq. 2.1 and Eq. 2.2 to a single energy dissipation formulation
as:
Ud =

ZVZ
0


0

Kl [ l e
i

where l is the local internal stress.


21

2l

i

]dl dV

(2.4)

After integrating Eq. 2.4, it becomes


Ud =

Z
0

Ki f1 e

2

i

(1 + 2 + 2  )gdV
2

(2.5)

So, the energy dissipation density function is as follows:


Uc = Kif1
where s is de ned by
s=

e

i

2s

(1 + 2s + 2s )g
2

= 

Therefore, the loss factor, c, of the coating material de ned by


Uc
c =
2Uc
Because the strain energy is de ned as
1
Uc = 
2E
the loss factor, c, can be represented as
2

c =

KE
[1 e
i

2s

(2.6)

(1 + 2s + 2s )]=s
2

(2.7)
(2.8)
(2.9)

where E is the Young's modulus and Uc is the vibratory energy density of the
coating material.
Eq. 2.9 shows the loss factor or the energy dissipation density per cycle of the
coating material is independent of the vibratory frequency.
2.2 Damping Measurement of the Coating Material

The above energy dissipation concept is applied to examine the e ects of the
coating damping on forced response of beams and blades. Initially, as shown in
22

Fig. 2.8, simply supported and cantilevered beams with a coating thickness of 10%
of the beam's thickness are considered, where the coating layer is located on the
top surface of the beams. The bending rigidity, EI (1 + i), of the coated beam is
calculated by solving the Ross-Kerwin-Unger equation [28]
EI (1 + i ) = Es Is f(1 + is ) + e2 h32 (1 + ic )

+3(1 + h ) [ 1 +e eh h(1(1+ +ici) ) ]g


2

where

e2 =

Ec
;
Es

h2 =

h
d

(2.10)
(2.11)

and h is the thickness of the coating layer, d is the thickness of the beam, and
Es Is (1 + is ) is the bending rigidity of the uncoated portion of the beam.
This RKU analysis was developed and solved using sinusoidal expansions for
the modes of vibration for simple beams and plates. The analysis also assumes
rigid connections between the coating layer and the beam itself. In this research,
the RKU equation has been successfully applied to predict the performance of the
unconstrained-layer coating damping treatment.
To further measure the damping properties from the magnetomechanical coating, a damping test has been accomplished. In accordance with the ASTM standard
E756-93 [12], a Ti-6Al-4V beam and a steel beam were machined having dimensions
7:087in  0:3937in  0:079in. A sinusoidal signal produced by a Siglab 20-22 digital
signal processor and ampli ed by a 75V power ampli er (Bruel and Kjaer Power
Ampli er 2706) was used to drive a dynamic shaker (Bruel and Kjaer Vibration Excitor 4809), exciting the clamped edge of the beams. The excitation frequency was
23

Figure 2.2: The Steel Beam Used in the Damping Measurement


swept from 50 Hz to 500Hz. The excitation frequency used in the later frequency
computation was measured from the input voltage of the shaker, and transferred to
the rst input channel of the digital signal processor. The output was determined
via a strain gage bonded 0:25in toward the clamped edge of the coated beam, shown
in Fig. 2.2. After using a pre-ampli er to boost the output signal, the signal from
the strain gage is sent to the second input channel of the digital signal processor.
The experimental set-up can be seen in Fig. 2.3. The input and output signals were
then employed to produce the frequency-dependent transfer function. This is so
called the swept-sine measurement technique. In turn, the damping loss factors of
24

the rst four vibration modes of the beams were calculated using the half power
point method. This method measures the bandwidth of resonant frequencies where
p
the amplitudes of response are 1= 2 times the maximum response located at the
resonant frequencies of the system. Then, the loss factor can be written as the
bandwidth of the resonant frequency divided by the resonant frequency. Following
this procedure, the loss factor of the steel beam, as shown in Fig. 2.2, and Ti-6Al-4V
beam were measured as (approximately) 0.003 and 0.002, respectively. The steel
beam result is presented in Fig. 2.4.

Figure 2.3: The Coated Beam Vibration Experiment.

25

Steel Beam (ASTM E75693)


0.035

0.03

0.025

Voltage

0.02

0.015

0.01

0.005

0
220

225

230

235

240
Frequency (Hz)

245

250

255

260

Figure 2.4: The Frequency Response of the Steel Beam Vibration.


Another steel plate of 5:118in  3:937in  0:79in coated with a 0:079in thickness
coating was cut into beams, similar to the beam shown in Fig. 2.6 and Fig. 2.7.
The excitation input and acceleration was measured by an accelerometer mounted
just above the xed end of the beam, and the output was determined via a laser
vibrometer at the free end of the beam, as shown in Fig. 2.5. The input and output
signals were employed to produce the frequency-dependent transfer function. By
using the swept-sine measurement technique as mentioned previously, the frequency
responses of the coated beam can be determined. In this test, the upper surface
strain at the xed end of the beam has been calculated from simple beam theory,
26

based on the de ection at the end of the beam as measured by the laser vibrometer.
By adjusting the input velocity of the dynamic shaker, the damping loss factor was
determined over a wide range of strain levels (0 4  10 ) from the frequency
response function of the coated beam using the half power point method at the
rst vibration mode. Then the damping loss factor c, as shown in Fig. 2.9, of the
coating layer was calculated using the Ross-Kerwin-Ungar equation Eq. 2.10.
4

Laser Vibrometer
Accelerometer
Beam
Shaker

Pre Amplifier

Computer

Power Amplifier

SigLab Digital
Signal Processor

Figure 2.5: The Coated Beam Vibration Experimental Setup.


Once the damping loss factor c of the coating layer is determined, damping
properties (K), and the internal stress (i) of the coating material can be obtained
by tting the loss factor function Eq. 2.9 in a least squares sense to the testing data
over the strain amplitude range that was investigated. Although any vibration mode

27

Figure 2.6: The Two Steel Beams Coated with the Magnetomechanical Coating and
One Magnetomechanical Material Beam Before Machining.
response can be used to evaluate the damping properties, only the fundamental (1st
mode) frequency was considered here for simplicity.
Due to the diculties of hysteresis loop measurement, it is almost impossible to
directly create the hysteresis loop for this magnetomechanical coating material in
order to obtain magnetic parameters K from the hysteresis loop's slope. So, an
alternative way has been used in this research. In the beginning, the dependence
of the damping loss factor on the strain amplitude is obtained from the previous
experiment, shown in Fig. 2.9. Then, the least-squares t between the testing
28

Figure 2.7: The Steel Beam Coated with One Magnetomechanical Coating Layer
data and the loss factor determined the damping properties (K) and the internal
stress to be 0:36, and 2:5  10 psi, respectively. Since the damping properties
are observed to be stationary at high strain range, only a strain amplitude up to
4  10 was considered in the least squares procedure. A least-squares procedure
that leads to better agreement for high strains could have been performed, but then
a discrepancy would occur for small strains. Thus, the least-squares t chosen here
is believed to be a good compromise because it leads to excellent results for strains
up to 4  10 , while large strains are of no practical importance since noticeable
magnetomechanical damping would only occur at lower strain amplitude.
3

29

h
b

2d
l

Cantilevered Beam
h
b

2d
l

Simply Supported Beam

Coating Layer

Beam

2d

Beam Cross Section

Figure 2.8: The Geometry of the Coated Beams.


2.3 Theoretical Development
2.3.1 Kinematic Assumptions

The energy dissipation density per cycle of the coating can be characterized by
a single function de ned in Eq. 2.6. Using this function, the energy dissipation
through the coating layer can be expressed as:
30

0.35
Test Results
Curve Fit
0.3

loss factor c (coating layer)

0.25

0.2

0.15

0.1

0.05
cr/Ec
0

0.5

1.5

2
strain amplitude

2.5

3.5

4
4

x 10

Figure 2.9: The Comparison of Curve Fitting and Test Results

Uc =

Z lZ
0

d+h

Ki [1 e

2s

(1 + 2s + 2s )]dzdx
2

(2.12)

By using the beam theory, the strain of the beam can be rewritten as:
 = zw

00

(2.13)

where z represents the distance from the neutral axis of the beam.
So, substituting Eq. 2.13 into Eq. 2.12, Eq. 2.12 can be rewritten as follows:
Uc =

Z lZ
0

d+h
d

Ki [1 e

00

2 zw
i

zw00
(1 + 2 
i

31

zw00 2
+ 2(  ) )]dzdx
i

(2.14)

After integrating in respected to z, the Eq. 2.14 becomes


Zl
Uc = fKEci [hw00 + (2(h + d)w00 + (h + d) w00 + 32 i)e
i
00
( 32 i + 2dw00 + d w00 )e ]=w00 gdx
2

(d+h)w

00

i

dw
2 
i

(2.15)

In accordance with the proposed free layer coating concept, the thickness of the
layer should be relatively small in comparison with the thickness of the beam or
blade. By using this thin coating layer assumption, the lower surface strain of the
coating layer is equal or similar to the upper beam surface, which can be written as
 = (d)w00 . Therefore, the stresses or strains are assumed to be the same through
the coating thickness. Also, this assumption leads that the overall thickness of the
coated blade 2d + h can be replaced by the thickness of the uncoated ones 2d during
the following computation. Hence,
e

(d+h)w

i

00

e

(d)w

00

(2.16)

i

Then, substituting (d + h)w00 by (d)w00 , Eq. 2.15 can be simpli ed as:


Uc =

Z
0

KEc i h[1 + (2 +

(h + 2d)w00 )e
i

2dw

i

00

]dx

(2.17)

2.3.2 Variational Theorem

Using classical variational principles such as the Hu-Washizu principle [34], the
entire energy expression for the coated beam problem has been formulated by taking
into account the energy dissipation in the coating layer. This yields the following
functional:
Zt Zl1
 = t f [ 2 EI (1 + is)w00 + K 0Eci h[1 + (2+
32
2

1 b(2d)w_ 1 bhw_ ]dxgdt


(2.18)
]
i
2
2
where  is the mass density, i is the internal strain, l is length of the beam, 2d is
the thickness of beam, h is the thickness of the coating layer, b is the width of the
beam, and K 0 is equal to K=2.
Consider a coated beam subjected to an arbitrary set of surface tractions and
body forces. Then, there is only negligible di erence between the deformed body
geometry and the undeformed body geometry for small de ections. Therefore, the
boundary surface of the arbitrary coated beam is considered to be essentially unchanged by the external loads. By extending this assumption, the principle of
virtual work contends that a solid is in equilibrium if the sum of the external and
internal virtual work is zero for kinematically admissible virtual displacements. It
follows that the functional  in Eq. 2.18 is stationary for the solution w. Therefore,
the rst variation of  must be equal to zero, yielding
(h + 2d)w00 )e

 =

l
0

2dw

00

i

fEI (1 + is)w00 w00 + K 0 Ecih[ 4 d

(
h 2d) 2(h + 2d)dw00 00
]w e
+
i

2dw

00

i

i 2

b(2d + h)w_ ] w_ gdx

(2.19)

After some simpli cation, Eq. 2.19 can be rewritten as:


 =

Z
0

fEI (1 + is)w00 w00 + K 0 Ech[(h 6d)

(2hd + 4d )w00 ]e
2

i

2dw

i

00

00

b(2d + h)w
_ w_ gdx = 0

After integration by parts, Eq. 2.20 becomes


00

 = EI (1 + is )w w j0 +K 0 Ec h[(h


l

33

(2
hd + 4d )w00
]
6d)

2

(2.20)

2dw

00

i

jl b(2d + h)ww
_ jl
0

l
0

f[EI (1 + is)w000

+K 0 Ech[ 2(h  6d)d w000 2(h + 2d)d w000


i
i
00
 gdx = 0
+ 4(h + 2d)d w00 w000 ]e ]w0 + b(2d + h)ww
2

(2.21)

2dw

i

i

After another integration by parts, Eq. 2.21 becomes


00

 = EI (1 + is )w w

jl EI (1 + is)w000 w jl
0

00
(2
hd + 4d )w00
0
6d)
]
e
w jl
i
000
000
4
hdw
8
dw
4(
h + 2d)d w00 w000
0
K Ec h[
+  +
]


2

+K 0Ech[(h

2dw

i

00
2dw

i

jl b(2d + h)ww
_ jl + fEI (1 + is)wiv
0

4(h + 2d)d (w00 wiv + w000 ) + 8(h + 2d)d w00 w000


2

K 0 Ec h[

i
3
+ 162d w000 + 4hd wiv
i
i

i

8d

i

wiv ]e

00
2dw
i

8hd

i

+ b(2d + h)wgwdx = 0

0002

(2.22)

Therefore, for arbitrary independent variations, the equation of motion is obtained for the coated beams:
EI (1 + is )wiv + K 0 Ec h[

8(h + 2d)d
3i

00 0002

ww

4(h + 2d)d w00 wiv


2

2i

+ 4d (3h 2d) w000 + 4d(2d


2

! 2 b(2d + h)w = 0

h)

wiv ]e

2dw

00

i

(2.23)

where ! is the natural frequency.


Associated boundary conditions are as follows:
00

EI (1 + is )w w

34

jl = 0
0

(2.24)

000

EI (1 + is )w w jl0 = 0

2(
h + 4d)dw00
6d)
]e
i
h) 000 4(h + 2d)d w00 w000
w +
]e

K 0 Ec h[(h

4d(2d
K 0 Ec h[

2dw

i

00

i

i

(2.25)
w

2dw

i

00

jl = 0
0

b(2d + h)ww
_ jl0= 0

jl = 0
0

(2.26)
(2.27)
(2.28)

The above equation of motion shows that damping e ect of the magnetomechanical coating layer is dependent on three issues: the magnetic strength of the
coating material (K 0  and i ), the thickness of the coating layer (h), and the thickness of the beam (2d). Therefore, the beam coated with the thicker coating layer
did have a larger e ect than the thinner coating layer on suppressing the vibration.
Additionally, K 0 and i are related to the properties of the ferromagnetic material. In this research, the magnetomechanical coating has a unique large number
of K 0, which made the hysteresis loop wider than the regular ferromagnetic material. This creates the unusual high damping properties for this magnetomechanical
coating material. Finally, the coating layer damping properties are dependent on
the vibratory strain or stress in the beam. This is related to the thickness of the
beam, 2d, contained in those energy dissipation terms. In spite of its complicated
formula, the equation of motion obviously represents the unique movement of the
magnetomechanical coating material.
Clearly, if there is no coating layer, the coating parameters K 0  are zero. This
leads to the second, third, fourth , and fth terms of Eq. 2.23 being zero. Then,
Eq. 2.23 can be rewritten as follows:
EI (1 + is )wiv

! 2 b(2d + h)w = 0

35

(2.29)

Therefore, the complex equation of motion in Eq. 2.23 reduces to the simple
uniform Bernoulli-Euler beam equation.

36

CHAPTER 3
APPLICATION TO BEAMS AND BLADES

The e ects of coating, 10% of the beam thickness, on the forced response of
simply supported and cantilevered beams have been studied. Results have been obtained for the lowest three vibration modes by a closed form procedure [11] using the
collocation method [2], described in Appendix A, and the nite element approach.
In the nite element approach, there are two alternative ways which have been
chosen to model the coated layer. At rst, four-node isoparametric elements have
been used to simulate the simply supported or cantilevered beam with the coating layer. By using Eq. 2.10, each individual element has its own loss factor and
material properties. During this nite element calculation, the simply supported
or cantilevered beam is treated as a composite beam containing twenty di erent
materials. Each material has its own individual loss factor. Another approach used
in this study is to create another customized four-node isoparametric element to
represent the coating layer, which is connected to the eight-node solid elements of
the regular steel beam, shown in Fig. 3.1. Unlike regular four-node shell elements,
four-node isoparametric elements in this research need to have a customized stress
or strain dependent loss factor and their neutral axis is located on the bottom of the
element, which is connected to the solid elements, in order to correct the moment of
37

inertia error during the numerical computation. Due to the similarity of these two
nite element results, only the results from the rst approach have been presented
in the beams' cases.
The nite element approach has also been applied to examine the forced response of coated blades as shown in Fig. 3.6. Approximately ten thousand four-node
isoparametric and eight-node solid elements have been used to model the coating
layer and the blade. The number of the elements is chosen to capture the e ect
of the coating on low frequency bending and torsion modes as well as on high frequency stripe modes in the frequency range of 3K-20K Hz. To cover the uctuating
vibratory stress pattern, the coating is applied to the entire pressure side of the
blade and modeled with two-dimensional four-node isoparametric elements, which
is similar to the coated beam case. It was demonstrated numerically that the nite
element mesh shown in Fig. 3.6 gives a nearly optimal result for the present problem. Results for both low frequency bending and torsion modes and high frequency
stripe modes have been obtained and evaluated. Additionally, later in this section,
the e ect of high rotating speed also will be considered in the coated cantilevered
blades for high stripe mode vibrations.
3.1 E ect of Coating on Forced Response of Simply Supported Beams

A beam of rectangular cross-section of dimensions 3:937in  0:3937in  0:039in


with a 0:008in thick plasma sprayed iron-chromium based coating layer has been
considered in this investigation. Each beam was excited by a 0:5lb harmonic load.
38

Magnetomechanical
Coating
4-Node
Isoparametric Elements

Steel Beam or
Ti-6Al-4V Blade
8-Node
Solid Elements

Figure 3.1: The Finite Element Model for the Blades Containing the Coating Layer
Since the harmonic loading is applied at the mid-span of the beam, no even numbered mode vibrations appeared in the results for simply supported coated beams.
As shown in Fig. 3.2, the rst and the third forced vibration mode responses
of a simply supported beam with a coating layer were computed and compared
to the responses from the uncoated beam. The results show that the mode shape
of the simply supported uncoated beam is a sine wave function, as expected, but,
surprisingly, the mode shape of the coated beam is also close to a sine wave function.
This can be explained by the eigenvalue from the equation of motion, Eq. 2.23,
which did not su er a tremendous change by adding four more energy dissipation
39

terms for the magnetomechanical coating , because K and i are still relatively
small compared to the Young's modulus, E, and the moment of inertia, I, from the
potential energy of the coated beams. Therefore, the only major di erence between
the mode shapes of the uncoated and coated beams are their amplitudes.
From these simply supported beam results, the maximum displacement reduction in the third mode is signi cant, as much as 62% from the nite element approximation and 55% from the closed-form solution. However, the rst mode shape
reduction is relatively low (29:8% in closed-form solution and 22:8% by nite element approximation), as shown in Table 3.1. This result can be explained by noting
that a high strain level (as high as 0:0033) is produced in a major portion of the
beam during the rst mode vibration. Consequently, lower damping (0:004 0:068
in loss factor) was produced from the coating layer in the rst mode.
Vibration Mode 1st Mode
Uncoated
0.272
Coated (FEM)
0.191
Reduction in % 29.8%
Coated (CF)
0.210
Reduction in % 22.8%

3rd Mode
0.0029
0.0013
55%
0.0011
62%

Table 3.1: The Maximum De ection (in) of Simply Supported Beams. (FEM: Finite
Element Results ; CF: Closed-Form Results)
The mode shapes obtained by the closed form approach and nite element approximation are compared in Fig. 3.2. One can observe that the agreement between
closed form solution and the nite element result is reasonable. Most importantly,
40

both results indicate that the 10% coating layer is capable of suppressing vibratory
energy and consequently reduces the stress. In addition, these results also demonstrate that the strain patterns of the rst and third modes trigger the irreversible
movement of magnetic domain boundaries of the coating material.
As expected, the stress distributions, as shown in Fig. 3.3, are also sine wave
functions, but with ninety degree phase angle di erences from the previous deformation functions. This is consistent with simple beam theory. In turn, the coating
layer is able to reduce the maximum stresses by 32% for the rst mode vibration.
It follows that the e ect of the coating layer is heavily dependent on the stress or
strain in the simply supported beam.
On the other hand, since the majority portion of the beam under the third mode
vibration is experiencing strain in the range of 0 1:3  10 , a high damping (as
high as 0:2872 in loss factor) is therefore achieved in the third mode. Similar to
the displacement, the coating layer has been able to reduce the maximum stress in
third mode vibration by 63% from the nite element approximation, as shown in
Table 3.2 and Fig. 3.3.
4

Vibration Mode 1st Mode


Uncoated
106
Coated (FEM)
72
Reduction in %
32%

3rd Mode
4.3
1.6
63%

Table 3.2: The Maximum Stress (ksi) of Simply Supported Beams. (FEM: Finite
Element Results)
41

3.2 E ect of Coating on Forced Response of Cantilevered


Beams

The e ects of the coating layer on the rst three modes of vibration of cantilevered beams were studied. Similar to the simply supported beam case, a relatively small e ect of the coating layer on the rst mode response is observed. Again,
this is caused by the large strain during the rst bending vibration mode. With
the exception of a small portion of the beam near the free end, the majority of the
cantilevered beam has a strain range larger than 4  10 . This high strain range
directly re ects the relatively low damping performance from the magnetomechanical coating layer when strains are above 4  10 . In the Table 3.3 and Fig. 3.4,
the maximum deformation from the rst mode vibration is only reduced by 22:8%
from closed-form solution and by 29:8% from nite element approximation.
Then, as expected, a much larger e ect is achieved in the second and third
vibration modes, with 90:3% and 84% reduction in the maximum displacement by
nite element approximation and 87% and 76% by closed-form solution, respectively,
as shown in the Table 3.3.
The maximum strain of the second and third modes are 3:50  10 and 1:67 
10 , respectively, which produce far greater loss factors of 0:3095 and 0:2501. This
indicates that the coating layer is able to suppress the vibratory energy in these
strain ranges for both simply supported and cantilevered beams. Also, the second mode results showed incredibly close agreement between analytical and nite
element results, with nite element results only 5% lower than analytical results,
as shown in Fig. 3.4. In addition, the displacement of the coated beam is highly
damped at the second mode vibration.
42
4

Vibration Mode 1st Mode


Uncoated
1.07
Coated (CF)
0.81
Reduction in % 24.3%
Coated (FEM)
0.69
Reduction in %
15%

2nd Mode
0.062
0.008
87%
0.006
91%

3rd Mode
0.0031
0.0007
76%
0.0005
84%

Table 3.3: The Maximum Deformation (in) of Cantilevered Beams. (FEM: Finite
Element Results ; CF: Closed-Form Results)
Vibration Mode 1st Mode
Uncoated
66
Coated (FEM)
54
Reduction in %
18%

2nd Mode
20.4
4.7
77%

3rd Mode
4.8
2.0
58%

Table 3.4: The Maximum Stress (ksi) of Cantilevered Beams. (FEM: Finite Element
Results)
Finally, as shown in Table 3.4 and Fig. 3.5, the vibratory stress of the cantilevered coated beams appear to have a similar response to their deformation results. The addition of the coating layer has only reduced by 18% the maximum
stress in the rst mode, because its strain is much higher than the best coating
layer damping performance range. However, it is a totally di erent story for the
second and third modes. As mentioned in the previous paragraph, the strain in
the second mode and third mode is located within the range 0 4  10 . So,
4

43

as expected, the magnetomechanical coating has a large e ect on the stress reductions, which are 77% and 58% in the second and third mode resonant vibration,
respectively. The results can be compared in Table 3.4 clearly.

44

0.35
uncoatedFEM
coatedFEM
coatedCF

Displacement (in)

0.3
0.25
0.2
0.15
0.1
0.05
0

0.5

1.5

2.5

3.5

x 10

Displacement (in)

2
1
0
uncoatedFEM
coatedFEM
coatedCF

1
2
3

0.5

1.5

2
Length (in)

2.5

3.5

Figure 3.2: First and Third Mode Shapes of Uncoated and Coated Simply Supported
Beams

45

1.5

x 10

uncoatedFEM
coatedFEM

Stress (psi)

1
0.5
0
0.5
1
1.5

0.5

1.5

Stress (psi)

5000

2.5

3.5

2.5

3.5

uncoatedFEM
coatedFEM

5000

0.5

1.5

2
Length (in)

Figure 3.3: Stress Distributions in the First and Third Modes of Uncoated and
Coated Simply Supported Beams

46

Displacement (in)

1.5
uncoatedFEM
coatedFEM
coatedCF

1
0.5
0

Displacement (in)

0.1

0.5

1.5

2.5

3.5

1.5

2.5

3.5

1.5

2
Length (in)

2.5

3.5

uncoatedFEM
coatedFEM
coatedCF

0.05
0
0.05
0.1

Displacement (in)

0
3
x 10

0.5
uncoatedFEM
coatedFEM
coatedCF

2
0
2
4

0.5

Figure 3.4: First Three Mode Shapes of Uncoated and Coated Cantilevered Beams

47

Stress (psi)

10

coatedFEM
uncoatedFEM

5
0
5
2

Stress (psi)

x 10

0
4
x 10

0.5

1.5

2.5

3.5

0
2
4

coatedFEM
uncoatedFEM
0

0.5

1.5

2.5

3.5

Stress (psi)

5000
coatedFEM
uncoatedFEM
0

5000

0.5

1.5

2
Length (in)

2.5

3.5

Figure 3.5: Stress Distributions in the First Three Modes of Uncoated and Coated
Cantilevered Beams

48

3.3 E ect of Coating on Low Frequency Forced Response


of Cantilevered Blades

A case study was also conducted to illustrate a general procedure for enhancing
damping capability of gas turbine blades using magnetomechanical coating. The
gas turbine blade, shown in Fig. 3.6, is made of Ti-6Al-4V, having dimensions: 3.0
inches wide  7.0 inches long and 0:4 inches average thickness. The aspect ratio
of the blade (1.75) was higher than some modern blade designs, but resulted in
frequencies and mode shapes, especially in high frequency stripe modes, similar to
a modern blade. Also, the thickness of the coating is 0:016 inches, which is about
4% of the blade's average thickness. The coating layer is applied on the pressure
side of the blade. A NASTRAN routine used in this simulation has been presented
in Appendix B.
Even though adding the coating layer on the pressure side of blade can reduce
displacement signi cantly, reducing the high stress for high cycle fatigue prevention
is a more important issue. Therefore, only stress distributions of the blade are
presented in this section. The Goodman Diagram of Ti-6Al-4V in Fig. 3.7 was
obtained from the study by Maxwell and Nicholas [24]. These data are used to
form the failure criteria or failure boundary in the traditional high cycle fatigue
turbine blading system design procedure. This design process usually consists of
a structural dynamics analysis to determine natural frequencies and mode shapes
at certain operating speeds and a stress analysis to calculate the dynamic stress
distribution for identifying the maximum vibratory stress location or region under
a series of given excitations.
49

Once the maximum stresses for each vibration mode have been determined, high
cycle fatigue assessment can be achieved by measuring the margin between the
maximum vibratory stress and the material fatigue capability which is a straight
line drawn between the mean ultimate strength at zero vibratory stress and mean
fatigue strength at 10 cycles (or in nite life), as shown in Fig. 3.7. Therefore,
the rst case study was conducted to illustrate how to attenuate the vibration
stresses in the rst three vibration modes of the cantilevered blade through the
magnetomechanical coating, which was presented in the previous section.
Consider the case of the blade under a concentrated external harmonic excitation
force applied at the tip portion of the blade. The magnitudes of the excitation
forces may be adjusted to achieve a stress ratio equal to 1 and a strain range of
0  4  10 . The rst three modes forced responses were computed for the coating
strain-dependent loss factors obtained from the Eq. 2.10. The von Mises stress
distributions of the rst three mode vibrations excited by 200lb force are shown in
Figs. 3.83.13.
One should note that the maximum vibratory stress for mode one, as expected
for this bending mode, occurs near the trailing and clamped edges of the blade and
shows signi cant reduction in comparison with uncoated blades. The maximum von
Mises stress dropped from originally 428ksi to 52:3ksi, which is more than 87%, at
the rst mode frequency, shown in Fig. 3.8. Actually, the von Mises stress of 428ksi
is already over the yielding point of the Ti-6Al-4V. Therefore, it appears that the
original turbine blade model without the coating layer has actually failed at the
rst resonant frequency 304.8Hz. According to this gure, the trailing edge of this
turbine blade will be the rst one to break apart, similar to Fig. 1.2. However, the
7

50

maximum von Mises stress of the coated blade is still in the safe range, and far
below the yielding point. This demonstrates that the magnetomechanical coating
layer has the ability to increase the design boundary of turbine blades, or it can
improve an original poor design such that it is acceptable. Fig. 3.8 and Fig. 3.9 also
display that the maximum von Mises stress actually shifts from the trailing edge
to the leading edge. This leads that the maximum stress also can be moved from
certain areas of the turbine blade by adding this coating layer. According to these
observations, the possibility of the magnetomechanical coating improving turbine
blade safety and reliability has been further enhanced.
The further investigation of the drop in von Mises stress at the second mode has
been shown in Fig. 3.10 3.11. By comparing these two gures, the maximum von
Mises stress drop a relatively small 36% from 16:9ksi to 10:8ksi. Once again, the
4% thickness coating layer appears to successfully suppress vibration. In the second
mode vibration, the stress distribution is also slightly changed by adding the coating layer. After adding the coating layer, the maximum von Mises stress is located
on not only the trailing edge, but also at the blade base. This development causes
the stresses to concentrate on two di erent regions of the blade, both with a relatively smaller amplitude, instead of one region with a much larger stress amplitude.
Finally, the third mode results are displayed in Figs. 3.123.13. The additional
coating layer makes the maximum von Mises stress decrease 28%, which drop from
70:5ksi to 50:6ksi. Similarly to the previous second mode, the reduction is much
smaller than the rst mode, because the strain in the rst mode is distributed most
in a particular strain range (1  4  10 ), leading to high damping contribution
4

51

from the coating layer. Also, observing these rst three mode vibrations, the resonant frequencies appear to increase after the magnetomechanical coating has been
applied to the blade surface. This can be explained as the magnetomechanical coating material has higher Young's modulus than Ti-6Al-4V. So, the coated blade has
a higher sti ness than the uncoated blade.
Several interesting observations can be made from the above discussions. First,
it is believed that no matter what the mode shape is, the coating layer can still
eciently reduce stresses of the blades when strains are near the range of 1  4 
10 . This is because the energy dissipation from the magnetomechanical coating is
dependent on strain, but independent of the mode shape at the resonant frequencies.
Second, the stress distribution can be changed by adding this coating layer. Even
though the change is small, it is sometimes enough to avoid stress concentrations
in critical areas, such as the trailing and the leading edge of turbine blades.
4

52

Figure 3.6: The Finite Element Mesh of Turbine Blade

53

Figure 3.7: Haigh (or Goodman) Diagram for Ti-6Al-4V

54

Figure 3.8: The von Mises Stress in the First Mode Vibration of the Uncoated
Cantilevered Blade

55

Figure 3.9: The von Mises Stress in the First Mode Vibration of the Coated Cantilevered Blade

56

Figure 3.10: The von Mises Stress in the Second Mode Vibration of the Uncoated
Cantilevered Blade

57

Figure 3.11: The von Mises Stress in the Second Mode Vibration of the Coated
Cantilevered Blade

58

Figure 3.12: The von Mises Stress in the Third Mode Vibration of the Uncoated
Cantilevered Blade

59

Figure 3.13: The von Mises Stress in the Third Mode Vibration of the Coated
Cantilevered Blade

60

3.4 E ect of Coating on High Frequency Stripe Mode Forced


Response of Cantilevered Blades

In addition to the low frequency rst three mode vibrations, another case study
was also conducted to illustrate a general procedure for enhancing damping capability of gas turbine blades using a magnetomechanical coating for high frequency
stripe mode vibration. The goal in this case study is to investigate the feasibility of
using the magnetomechanical coating for enhancing damping at high frequencies, as
shown in Figs. 3.143.20. As mentioned in the introduction, this plate-like stripe
mode has low or sometimes no movement near the base of blade. This leads to
traditional friction dampers having no ability to dissipate the energy. Also, aerodamping has very little e ect on this high frequency vibration, due to their low
deformations. These realities cause current technologies to have diculty achieving
the desired stress reduction. Unfortunately, the high frequency stripe modes are the
typical causes of the actual hardware high cycle fatigue. It is necessary to develop
an alternative method, such as magnetomechanical coating, to avoid the high cycle
fatigue problem, which has been addressed in Chapter 1.
The gas turbine blade is made of Ti-6Al-4V, having the same dimensions as the
previous low frequency simulation model of Fig. 3.6. The aspect ratio of the blade
is still 1:75. Also, as indicated in the previous section, this blade model is similar
to modern blade behavior in the high frequency stripe vibration modes. Again, the
pressure side of the blade has been coated with the magnetomechanical coating,
which is about 4% of the blade's average thickness.
As mentioned in the previous section, even though adding the coating layer on
the pressure side of blade can reduce displacement signi cantly, eliminating the
61

high stress for high cycle fatigue prevention is the major concern in high cycle
fatigue problems, which is the purpose of this research. Therefore, only the stress
distributions of the second, third, and fourth stripe modes of coated and uncoated
beams are presented and examined.
Therefore, a case study was conducted to illustrate how to attenuate the vibration stresses in the three di erent stripe high frequency vibration modes using the
magnetomechanical coating damping.
Consider the case of the blade under a similar concentrated external harmonic
excitation force applied at the tip portion of the blade, as used in the previous
section. Once again, the magnitudes of the excitation force may be adjusted to
achieve a stress ratio equal to 1 and a strain range of 0 4  10 . The von
Mises stress distributions of coated blades for the second, third, and fourth stripe
modes excited by a 50lb force on the central free end of the blades are computed
and compared to those of the uncoated blades in Figs. 3.153.20. Observe that
the nite element results con rm the previous experimental predictions that high
vibratory stresses will be produced at high stripe modes.
One should note that in Fig. 3.16 the maximum vibratory stress, as expected
in this second stripe mode, occurs near the clamped edges of the blade and shows
signi cant reduction in comparison with the uncoated blade in Fig. 3.15. The drop
in the maximum von Mises stress is from 94:4ksi to 77:8ksi, which is about 17:6%.
Comparing Fig. 3.15 to Fig. 3.16, the stress concentrations are on two regions near
the clamped edge, where the color is red and yellow. Among these regions, the
uncoated blade appears have a larger size than the coated blade. Actually, the
stress concentrated areas of the blade coated with the magnetomechanical coating
4

62

are almost only displayed as a smaller yellow area. This is because the magnetomechanical coating successfully reduced the von Mises stress of this second stripe mode,
and makes the stress concentrated regions become much smaller. However, unlike
the stress amplitude, the original stress distribution was not changed by adding this
magnetomechanical coating.
Similar to the second stripe mode, the magnetomechanical coating was able to
suppress the fourth stripe mode vibration by 9:4%. In this fourth stripe mode,
the maximum von Mises stress has been reduced from 22:4ksi to 20:3ksi, shown
in Figs. 3.193.20. These two gures show a typical example of the high stripe
mode problem. In these two gures, almost no von Mises stress appears near the
clamped edge of blade. However, the maximum von Mises stress can be found near
the tip of blade. This indicates the stress of this stripe mode at 18956Hz resonant
frequency cannot be reduced by the traditional friction dampers. These two gures
also clearly display the larger von Mises stresses only located on the six ellipses
shape regions. With the exception of these six areas, the von Mises stresses are
either small or zero. This results in the diculties of energy dissipation through
the traditional approach.
Compared to the 9:4%  17:6% reduction at the second and fourth stripe mode,
Figs. 3.173.18 present a far greater drop of 42:9% occurring in the third stripe
mode. The maximum von Mises stress of the coated blade was 12:4ksi, which
dropped from 21:7ksi for the uncoated blade. Similar to the beam case, the differences in the stress reduction can be explained by noting that an adequate (in
the sense of inducing coating damping) strain level or pattern is produced in the
majority of the blade during the third stripe mode vibration. Consequently, higher
63

damping is therefore produced from the coating layer. By observing the Figs. 3.17
3.18, the stress level in those stress concentrated regions near the tip of blade is obviously suppressed after the blade has been coated with the magnetomechanical
coating, where the color is changed from red-yellow (20ksi) to yellow-green (10ksi).
In accordance with the stress patterns presented in Figs. 3.153.20, the maximum stress of the second, third, and fourth stripe vibration modes have been
reduced between 9:4% and 42:9% by using only a 4% blade's thickness magnetomechanical coating. This leads to the fact that even though the maximum stress of
the blade can be as high as 94; 400 psi, the majority portion of the blade as shown
in Fig. 3.15 responds to a much lower vibratory stress level ranging from 400-6000
psi, which in fact is an ideal stress level for triggering the irreversible movement of
magnetic domain boundaries of the coating material. Therefore, even though the
highest strain at one particular resonant mode vibration sometimes is higher than
4  10 , which is over the primary target range for this magnetomechanical coating
material, the coating material is still able to perform well in some portions of the
blade, where the strain level still appears on the range of 1 4  10 . This means
the maximum strain or stress of the coated blade is still lower than the original
blade by adding this magnetomechanical coating.
Several interesting observations can be made from the above discussions. First,
the maximum vibratory stress of the coated blade can be signi cantly reduced at
a speci c mode and reasonably suppressed at the other modes through the magnetomechanical coating under a certain loading condition. Second, the strain range
or loading condition becomes a major factor on the magnetomechanical coating
damping capacity for a speci c vibration mode. Third, after simulating the low
4

64

frequency and high frequency vibrations, it is proven that the magnetomechanical


coating can still eciently reduce stresses of the blades, regardless of their resonant
frequencies.

65

Figure 3.14: The Third Stripe Mode Shape of the Cantilevered Blade

66

Figure 3.15: The von Mises Stress in the Second Stripe Mode Vibration of the
Cantilevered Blade

67

Figure 3.16: The von Mises Stress in the Second Stripe Mode Vibration of the
Coated Cantilevered Blade

68

Figure 3.17: The von Mises Stress in the Third Stripe Mode Vibration of the Cantilevered Blade

69

Figure 3.18: The von Mises Stress in the Third Stripe Mode Vibration of the Coated
Cantilevered Blade

70

Figure 3.19: The von Mises Stress in the Fourth Stripe Mode Vibration of the
Cantilevered Blade

71

Figure 3.20: The von Mises Stress in the Fourth Stripe Mode Vibration of the
Coated Cantilevered Blade

72

3.5 The Coating Thickness E ect of Rotating Cantilevered


Blades

A case study was performed to investigate the feasibility of using two di erent
thicknesses of magnetomechanical coating on the rotating cantilevered blade. Both
2% and 4% of the blade thickness were chosen as the magnetomechanical coating
thickness in this study. Since turbine blades typically rotate at high speed, the
e ects of the coating damping capability is also investigated in this study. It is
well known that rotating the blade creates another new source of force, the centrifugal force, as presented in Appendix C. Such force produces a non-negligible
amount of stress or strain along the length of the blade, and directly in uences the
damping capacity of the magnetomechanical coating. Figs. 3.223.29 are presented
the fourth stripe mode results of the simulated turbine blade for which traditional
dampers have some diculties to achieve the desired stress reduction.
In this section, the sti ening e ect of centrifugal loading has been considered
during the forced response simulation, as shown in Appendix D, in order to demonstrate the ability of magnetomechanical coating to enhance damping of the turbine
blade during its high rotating speed operations. Due to the complexity of the real
world twisted blade con guration, two di erent blade sitting angles (0o and 90o),
as shown in Fig. 3.21 by rotation Rx and Ry , were investigated.
In Figs. 3.243.26, the turbine blade with 0o sitting angle, modeled by rotation
Rx in Fig. 3.21, under 10; 000 rpm has been investigated. Similar to the non-rotating
cases, a maximum 50lb concentrated external harmonic excitation force has been
applied at the tip portion of the blade in the rotating blade cases. However, unlike
the previous cases, the magnitude of the overall stresses or strains are expected to be
73

much larger due to the extra centrifugal force and Coriolis acceleration. Comparing
Fig. 3.14 to Fig. 3.22 clearly shows that the third stripe mode deformation and the
resonant frequency are increased by considering the rotating motion, even though
the mode shapes are similar.
One fourth stripe mode vibration is presented in this section, shown in Fig. 3.23.
The maximum von Mises stress at this fourth stripe mode, as expected, occurs near
the tip edges of the blade and shows signi cant reduction in comparison with uncoated blades, shown in Fig. 3.24, in either 2% thickness coating case, shown in
Fig. 3.25, or 4% thickness coating case, shown in Fig. 3.26. The drop of the maximum von Mises stress is 6:5% by adding 2% thickness of the coating layer, which
drop from 26:2ksi in the uncoated blade to 24:5ksi in the coated blade. When the
coating layer thickness becomes 4% of the overall blade thickness, the maximum von
Mises stress decreases from 26:2ksi to 22:5ksi. The reduction of the maximum von
Mises stress increases to 14:1%. However, the drop is still slightly higher than the
9:4% in the non-rotating case. Like the previous non-rotating case, the di erences
in stress reduction can be explained by noting that an adequate (in the sense of
inducing coating damping) strain level or pattern is produced in the majority of
the blade during the fourth stripe mode vibration. Another important observation
needs to be noted from these three gures. The thickness of the magnetomechanical
coatings did a ect their performance. By cutting the coating thickness to half, the
coating layer only makes the maximum von Mises stress drop 6:5% from 14:1%.
On the other hand, reducing by 6:5% the overall maximum stress is still considered
acceptable. The purpose of simulating the blades with these two di erent coating
thicknesses is to demonstrate that the design engineers would be able to optimize
74

their design by adding the proper thickness of coating to meet their reliability and
safety requirements. More precisely, by knowing the limitation of the maximum
stress during the blade design process, the engineers can design the coating layer
as thin as possible, such that it is enough to suppress the stress to an acceptable
level, without adversely increasing the overall blade weight and reducing their performance and fuel eciency.
Consequently, there are a couple of observations that can be made of these results. First, the thicker coating layer produced the higher damping, as expected.
Second, the addition of centrifugal force in uenced the performance of the coating layer. However, the coating layer is still able to suppress the vibration in a
reasonable manner.
As mentioned in the previous section, the turbine blade in the real world usually
has a twisted shape and is rotated with a particular angle to the blade surface in
order to have the best aerodynamic performance. However, as previously noted the
damping performance of the magnetomechanical coating is sensitive to the stress
in the blade. Therefore, an additional study about the magnetomechanical coating damping capacity with 90o blade sitting angle was investigated. As shown in
Figs.3.273.29, for the fourth stripe mode vibration the von Mises stress has been
reduced 5:1% by the 2% thickness of the coating layer or 13:5% by the 4% thickness of the coating layer, because the average strain is not located in the highest
damping range of the magnetomechanical coating. These results show the magnetomechanical coating has the ability to enhance the turbine blade damping under
di erent sitting angles and amplitudes of the centrifugal force, even though the
stress distributions and the resonant frequencies are changed.
75

Blade

Rotor

Shaft
Z

Ry

Rx
X

Figure 3.21: The Rotating Diagram of the Coated Blades


Simplifying the previous observations. It has been shown that the magnetomechanical coating is still able to successfully reduce the maximum von Mises stresses
under the high rotating speed and di erent coating thickness, regardless of its sitting
angle.

76

Figure 3.22: The Third Stripe Mode Shape of the Cantilevered Blade about Rx
under 10000 rpm Rotation

77

Figure 3.23: The Fourth Stripe Mode Shape of the Cantilevered Blade under 10000
rpm Rotation

78

Figure 3.24: The von Mises Stress of the Fourth Stripe Mode of the Cantilevered
Blade Rotating about Rx under 10000 rpm

79

Figure 3.25: The von Mises Stress of the Fourth Stripe Mode of the Coated Cantilevered Blade Rotating about Rx under 10000 rpm with 2% Thickness Coating
Layer

80

Figure 3.26: The von Mises Stress of the Fourth Stripe Mode of the Coated Cantilevered Blade Rotating about Rx under 10000 rpm with 4% Thickness Coating
Layer

81

Figure 3.27: The von Mises Stress of the Fourth Stripe Mode of the Cantilevered
Blade Rotating about Ry under 10000 rpm

82

Figure 3.28: The von Mises Stress of the Fourth Stripe Mode of the Coated Cantilevered Blade Rotating about Ry under 10000 rpm with 2% Thickness Coating
Layer

83

Figure 3.29: The von Mises Stress of the Fourth Stripe Mode of the Coated Cantilevered Blade Rotating about Ry under 10000 rpm with 4% Thickness Coating
Layer

84

CHAPTER 4
DEVELOPMENT OF A FINITE ELEMENT
PROCEDURE FOR ADHESIVE BONDED JOINTS

The main objective of this chapter is to develop a numerical tool capable of predicting stress and strain distribution in an adhesive bonded joint. Traditionally, the
closed form analytical methods, as mentioned in the introduction section, successfully present the stress or strain distribution for a small number of simple bonded
structures. However, the classical methods have diculties to further explore the
more complicated adhesive bonded joints. In turn, numerical methods, especially
the nite element procedure, have been developed to simulate a wide range of di erent adhesive bonded joints by using a set of given parameters, such as the adherend
thickness, adhesive thickness, joint width, and overlap length. In this chapter, a
two-dimensional nite element procedure has been developed for the adherends and
adhesive layer of the single-lap and double-strap joints. The results from this newly
developed nite element procedure are then compared to the earlier, classical, stress
analyses of adhesive joints, such as Goland and Reissner, Hart-Smith, and Oplinger.
Based on this nite element procedure, a modi ed procedure has been developed to
predict the stress distributions of the more complicated imperfectly-bonded joint in
the next chapter.
85

4.1 Finite Element Model for Single-Lap Joint

The stress distribution along the adhesive bonded layer of a single-lap joint,
one of the most commonly used bonding con gurations, has been computed by
using a two-dimensional nite element procedure. The geometrical and material
parameters of a single-lap joint have been chosen for this model development, as
shown in Fig. 4.1, where lb is the length of the outer adherend, 2Cb is the length
of the overlap, t is the thickness of the adhesive layer, and b is the thickness of
the adherend. The elastic modulus and Possion's ratio of adherend are Ea and a .
Respectively, Eb and b represent the elastic modulus and Poisson's ratio for the
adhesive layer. In this study, the widely used FM73 adhesive has been chosen for
the bonding layer. In the nite element analysis, ten elements were used across
the thickness of each adherend and four across the thickness of the adhesive layer.
Three hundred elements were used along the length of the bonded overlap and forty
elements along the length of the unbonded overlap. Due to the stress singularities
near the ends of the overlap, the number of elements was tripled in these areas in
order to have more accurate solutions, as shown in Fig. 4.1. The aspect ratio of
elements in the adhesive layer was 3. Since this con guration has four elements
across the thickness, the stress variations across the thickness of the adhesive layer
can be demonstrated. In addition, Quad-8 elements have been successfully used to
simulate the high stress and strain gradients near the end of the overlaps.
Due to the stress symmetry along the mid-span of the adhesive layer in the singlelap joint, only one half of the mid-span peel and shear stress distributions from the
two-dimensional nite element procedure are presented in this section. The results
86

show the peel and shear stress distributions under 1000lbs of uniform tensile loading
or a single concentrated tensile load, compared to three analytical results, Goland
and Reissner (dash line), Hart-Smith (dash-dotted line), and Oplinger (dotted line),
shown in Figs. 4.24.4.
In Fig. 4.2, a linear geometrical nite element procedure is compared to the
three theoretical solutions of the single-lap joint. This linear geometrical procedure
only considers the small de ection of bonded joints. It is assumed that the bonded
structure still has the same shape during the loading process. In this rst comparison, all three theoretical solutions, Goland and Reissner, Hart-Smith, and the
Oplinger approach, show the stresses slowly decrease from the center of the joint
overlap to near the end, but sharply increase near the free edges of the overlap. The
comparison between the three theoretical solutions and the nite element analysis
show the Goland and Reissner result is the closest to the nite element result, because Goland and Reissner consider the geometrical linearity like this nite element
procedure. The other two results appear to have both lower peel and shear stresses
than the nite element results. However, the nite element procedure also show
both peel and shear stresses suddenly decreasing to near zero while approaching
the free edge instead, which satis es the boundary conditions on the free edges. It
is believed that the classical theoretical results are not able to represent this stress
singularity because of their one-dimensional approach. Excluding the results near
these edges, generally, the four analyses show the stress distributions are symmetric,
and have lower stresses near the middle of the adhesive layer.
Considering nonlinear material and geometrical properties, shown in Fig. 4.3,
the stress distributions from the two-dimensional nite element procedure is slightly
87

lower than the previous linear solution. With the exception of the stress singularities
near the free edge of the overlap, this improvement made the peel stress results from
the nite element procedure agree well with the Hart-Smith and Oplinger results,
because of their geometrical nonlinear assumptions. However, the shear stress from
the nite element results shown in Fig. 4.3 is slightly higher than desired. This
requires further investigation into the possible cause of this shear stress di erence
between the theoretical solutions and the nite element results.
In order to have more accurate comparisons between the results, a single concentrated load has been applied to the single-lap joint to replace the original uniform
load con guration used in the nite element procedure, because of the single load
con guration used in the one-dimensional theoretical solutions. The nite element
result is expected to have much closer shear stress results in comparison with the
three theoretical results after a 1000lbs single load is applied to the middle level of
one end of the adherend. The results, shown in Fig. 4.4, show that the di erent
loading con guration did cause the maximum peel and shear stress distributions to
further drop about 5%. Both peel and shear stresses in this single load case agree
well with the theoretical results and nite element analysis, with the exception of
the stress singularities near the free edge of the overlap. This proves the accuracy
and eciency of this newly created nite element procedure.
Following the above successful comparison, the same nite element procedure
is used as a base tool for the imperfectly-bonded adhesive bonding research in the
next chapter. However, there is one more issue which needs to be addressed. Even
though these results better match the theoretical solutions, the two-dimensional
88

nite element procedure for the rest of this research will still use the uniform loading
con guration to avoid the stress concentration near the end of the joint lap.
Now, Fig. 4.5 demonstrates how powerful this two-dimensional nite element
procedure is. Because it contains four two-dimensional Quad-8 elements through
the adhesive bonding thickness, the peel and shear stress distributions can not
only show the results along the length of the overlap, but also show the peel and
shear stress variations along the adhesive bond thickness. This new information lets
engineers understand the peel and shear stress change inside the adhesive bonded
layer. In Fig. 4.5, the top (y = t ) surface of the adhesive bonded layer, presented as
a solid line, appears to have its maximum peel and shear stresses located near the
left end of the overlap, and another concentration of peel and shear stresses near
the right end. Obviously, those stresses near the left end are larger than those near
the right end. It is believed to be caused by the external uniform loads from the
upper adherend pulling the adhesive bonds from the left side. At the same time,
the lower adherend has been constrained by the right edge of its lap. This actually
leads to the maximum shear and peel stresses being located on either the upper left
or lower right corner of the adhesive bonded layer. According to this observation,
the bottom surface (y = t ) of the adhesive bonded layer has its maximum peel
and shear stress, shown as a dot line, near the right end of the overlap, and there is
another high stress region near the left end, but smaller than the one at the right
end. Comparing the peel stress to the shear stress, the shear stress distribution
reaches its local maximum value just slightly inside the left edge, and drops nearly
to zero at the left free edge of top surface, caused by the stress singularity, but the
peel stress has its local maximum value at the left edge. Therefore, the peel stress
2

89

y
lb

Cb

Cb

lb
x

Pb

b
b/2

The Double-Strap Joint


lb
Pb

2Cb

lb

x
t

The Single-Lap Joint

The Finite Element Mesh

Figure 4.1: The Geometry and Finite Element Mesh of the Bonded Joints
distributions at the top surface are no longer symmetric, where the free edge side has
nearly no stress and the side connected to the joint has the maximum peel stress.
The bottom surface stress distributions are asymmetric to the top surface stress
distributions. Instead of the asymmetry between these top and bottom surface
results, the middle level of the shear and peel stress distributions appear to be a
symmetrical curve between the right and left end of the overlap.

90

0.8
GR
OP
HS
FEM

0.4

xy

/P

0.6

0.2

0.05

0.1

0.15

0.2

0.25

0.15

0.2

0.25

Cb
0.8
GR
OP
HS
FEM

0.6

y/P

0.4
0.2
0
0.2

0.05

0.1
Cb

Figure 4.2: Adhesive Stress Distributions Obtained from the Present Finite Element Model with Linear Geometry (Uniform Loads)(GR: Goland and Reissner;
OP: Oplinger; HS: Hart-Smith; FEM: Finite Element Method )(y=0)

91

0.8
GR
OP
HS
FEM

0.4

xy

/P

0.6

0.2

0.05

0.1

0.15

0.2

0.25

0.15

0.2

0.25

Cb
0.8
GR
OP
HS
FEM

0.6

y/P

0.4
0.2
0
0.2

0.05

0.1
Cb

Figure 4.3: Adhesive Stress Distributions Obtained from the Present Finite Element
Model with Nonlinear Geometry (Uniform Loads)(y=0)(GR: Goland and Reissner;
OP: Oplinger; HS: Hart-Smith; FEM: Finite Element Method)

92

0.8
GR
OP
HS
FEM

0.4

xy

/P

0.6

0.2

0.05

0.1

0.15

0.2

0.25

0.15

0.2

0.25

Cb
0.8
GR
OP
HS
FEM

0.6

y/P

0.4
0.2
0
0.2

0.05

0.1
Cb

Figure 4.4: Adhesive Stress Distributions Obtained from the Present Finite Element
Model with Nonlinear Geometry (Single Loads)(y=0)(GR: Goland and Reissner;
OP: Oplinger; HS: Hart-Smith; FEM: Finite Element Method )

93

0.7
y=t/2
y=0
y=t/2

0.6

0.4

xy

/P

0.5

0.3
0.2
0.1
0
0.25

0.2

0.15

0.1

0.05

0
Cb

0.05

0.1

0.15

0.2

0.25

0.05

0.1

0.15

0.2

0.25

y=t/2
y=0
y=t/2

0.8

y/P

0.6
0.4
0.2
0
0.2
0.25

0.2

0.15

0.1

0.05

0
Cb

Figure 4.5: Adhesive Stress Distributions Obtained from the Present Finite Element
Model with Nonlinear Geometry (Uniform Loads)(y=0)

94

4.2 Finite Element Model for Double-Strap Joint

In addition to the single-lap con guration, another common bonded structure,


the double-strap joint, has been studied in this section in order to evaluate the
accuracy of this two-dimensional nite element procedure, shown in Fig. 4.1. With
a similar nite element procedure as in the previous single-lap joint case, the doublestrap joint has been xed on the left end of the adherend, and 2000lbs uniform load
is applied on the right end of the adherend. Due to the geometrical symmetry,
only one of the four adhesive layers has been presented in this study, and compared
to the theoretical solution from Oplinger [27] to verify this two-dimensional nite
element procedure.
Fig. 4.6 shows the shear and peel stress distributions obtained along the bottom
(dotted line), mid (dash line), and top (solid line) interface of the adhesive bonding
layer at the upper right corner of the joint. Because of its di erent con guration, the
peel and shear stress distributions at the top and bottom surfaces of the adhesive
bonding layer are no longer heavily concentrated on the left or right side of the
overlap. At the left free edge of the adhesive bonded layer, the shear and peel
stresses at the top interface and the shear stress at the bottom are nearly equal to
zero, but the peel stress reaches its maximum value at the bottom interface. On
the right edges, the peel stress reach its maximum value at the top interface, but
approaches zero at the bottom. The shear stresses at all three levels (top, middle,
bottom) inside the adhesive bonded layer are basically symmetric between the right
edge and left edge. Similar to the shear stress distribution of the single-lap joint,
the shear stress is higher near both ends of the overlap, but much lower at the center
95

of the overlap. However, the shear stress near both ends of the double-strap joint
only slightly increase initially, then drop to zero at the free edge of the adhesive
bonding layer, instead of the stress singularities in the single-lap joint case. This
is because the loads are distributed through two overlaps in the double-strap joint.
Unlike the shear stress, the maximum peel stresses are located near the left edge
of this overlap, regardless of their interface locations. Then, the peel stress almost
constantly decreases to zero near the right end of the overlap. Additionally, at
the right end of the overlap, the peel stress becomes compressive at the bottom
interface, but it is still tensional stress at the top interface. This is believed to be
caused by the large de ection occurring in the double-strap patch that connects the
two adherends.
In order to further evaluate the accuracy of this double-strap joint result, the
stress distributions at the middle level of the adhesive layer from the nite element results have been compared to those from the theoretical results developed by
Oplinger [27], shown in Fig. 4.7. Oplinger's results have been computed by using
the Visual Basic based TJOINTNL program. Fig. 4.7 shows that the nite element
results agree well with Oplinger's theoretical solutions in both peel and shear stress
distribution, with the exception of the peel and shear stresses near both ends of the
overlap. Because of Oplinger's one-dimensional computation, this theoretical result
is not decreasing to zero near the free edge. However, the nite element result does
appear to be zero, which satis es the boundary conditions. Excluding the results
near the edges, both methods yield consistent shear and peel stress distributions.
After testing two di erent con gurations, a good comparison with theoretical
results has increased con dence in the shear and peel stress distributions predicted
96

by this two-dimensional nite element analysis. Additionally, it also proves the


accuracy of the two-dimensional model in computing the peel and shear stresses
near the joint boundaries, and the ability to provide detailed stress distributions
inside the adhesive bonded layer. Therefore, this nite element procedure has been
used as a base tool to further predict the peel and shear stress of the imperfectlybonded joint in the next chapter.

97

1.5
y=t/2
y=0
y=t/2

xy

/P

0.5

0.05

0.1

0.15

0.2

0.25

0.15

0.2

0.25

Cb
1.5
1

y/P

0.5
0

0.5
y=t/2
y=0
y=t/2

1
1.5

0.05

0.1
Cb

Figure 4.6: Adhesive Stress Distributions Obtained from the Present Finite Element
Model for Double-Strap Joint (OP: Oplinger; HS: Hart-Smith; FEM: Finite Element
Method)

98

1.5
OP
FEM

xy

/P

0.5

0.05

0.1

0.15

0.2

0.25

Cb
0.6
OP
FEM

0.4

y/P

0.2
0
0.2
0.4

0.05

0.1

0.15

0.2

0.25

Cb

Figure 4.7: Adhesive Stress Distributions Obtained from the Present Finite Element
Model Compared with TJOINTNL Program (OP: Oplinger; HS: Hart-Smith; FEM:
Finite Element Method)

99

CHAPTER 5
DEVELOPMENT OF IMPERFECTLY-BONDED
INTERFACE MODELS

The computational model developed during the last chapter has proved to be
accurate and reliable. Based on these two-dimensional nite element models, the
newly developed nite element procedure is now extended to consider the environmental degradation of adhesive joints. Among those adhesive bonded joints
damaged by environmental degradation, the imperfectly-bonded joint is one of the
most common and serious defects. Furthermore, it is also the most dicult structural problem to be detected by current non-destructive inspection techniques. The
main objective of this chapter is to predict the peel and shear stress distributions
of the imperfectly-bonded joint. To represent the change in the mechanics of the
imperfectly-bonded interface, a distributed spring element model has been created
to simulate this imperfectly-bonded interface. Later, improving this distributed
spring model, a newly developed one-dimensional line element also will be introduced in this chapter. By using these simulations as a base, a more detailed inspection or design criteria can be created in order to detect the failure of the adhesive
bonded interface due to environmental degradation.
100

5.1 The Distributed Spring Element Model

To simulate the imperfectly-bonded interface, the nature of the mechanics for


the imperfectly-bonded interface need to be rst explored. According to Tada,
Paris, and Irwin [32], the far eld displacement will be increased by the localized
deformations due to the absence of the interface, when a tensile load is applied to
an elastic solid containing a planar collection of cracks. In the imperfectly-bonded
interface, it does not contain any cracks, but it does have some "gaps" between
adherend and adhesive. The mechanics of the imperfectly-bonded interface are
therefore similar to a series of cracks. By adopting this assumption, the additional
deformation caused by these voids can be related to the stresses on the interface
under periodic oscillations with a time harmonic variation by applying Newton's
second law. This can be written as follows:
j (0+ ) j (0

)=

m! 2

(5.1)
2 [Uj (0 ) + Uj (0 )]
where j = x; y; z directions, Uj represents the j th component of a particle's displacement from its equilibrium position, j is the j th component of the force per
unit area exerted on the bonded area, m is the mass of the adhesive bonded layer,
! the natural frequency, and the interface is located at y = 0.
By using this relationship from Eq. 5.1, Baik and Thompson [3] developed a
distributed spring model to simulate this extra deformation due to the imperfectlybonded interface, where the spring sti ness per unit area, K , is a function of the
topography of the partially contacting surface. The idea is that the spring sti ness is
101

in nity when two objects are connected by a rigid connection, such as the perfectlybonded interface, but it becomes zero when there is no connection, like the totally
disbonded interface. So, Eq. 5.1 can be rewritten as:
1 [ (0 )  (0 )] = K [U (0 ) + U (0 )]
(5.2)
j
j
j
j
2 j
where Kj is the sti ness of the interface.
With the exception of when the adherend and adhesive are directly connected,
there is a virtual spring between the adherend and adhesive to simulate the imperfectlybonded interface. This spring model was proved successful in predicting the frequency dependence of re ection and transmission in ultrasonic measurements of
partially contacting interfaces in adhesive bonded joints [3]. Furthermore, Margetan, Thompson, and Gray [23] extended this spring model to di usion bonds,
which can predict the dependence of bond re ectivity upon both angle of incidence
and ultrasonic polarization. Recently, Lavrentyev and Rokhlin [21] described the
degradation as an interfacial disbond lled by an absorbed uid, which is also modeled by the transverse spring boundary conditions, with the complex spring sti ness
representing the quality of the bond. In their research, the disbond due to the degradation can be characterized by slip boundary conditions, where the normal spring
sti ness, de ned as Kn0 , is in nity and only the shear spring sti ness, de ned as
Kt0 , need be determined in the spring model [23].
Therefore, the next step is to determine the spring sti ness Kt . At rst, the
imperfectly-bonded interface between adherend and adhesive might be assumed
to be a complex pattern of cracks and contacts. It is dicult to nd the exact
stress distribution in the imperfectly-bonded region. An alternative approach is to
102
+

approximate the interface by a periodic array of contacts or cracks. Discretizing


this approach furthermore, the interface can be separated into smaller pieces, where
each piece has only one individual crack. Using this assumption, the displacement
of the imperfectly-bonded interface can be determined by Tada, Paris, and Irwin's
formula [32] for the specimen with a strip crack under tensile loading written as
follows:
1 ln(sec a )
 = 2 (1E  )a 2 a=b
2b

(5.3)

2EaEb
Eb (1 b ) + Ea (1

(5.4)

ab

where  is the shear stress,  is the Poisson's ratio,  is the deformation, a is the
crack length, and bl is the interface length.
The e ective Young's modulus Eab in Eq. 5.3 can be written as:
Eab =

a2 )

where Ea and Eb represented the Young's modulus from two side of the crack.
The equilibrium sti ness of an imperfectly-bonded interface can be computed
by the potential energy of the imperfectly-bonded layer. According to Tada, Paris,
and Irwin's formula [32], the sti ness Kt0 can be written as:
Kt0 =

a
 Eab 1
ln
(
sec
4 (1  2) b
2b )

(5.5)

According to Lavrentyev and Rokhlin [21], the imperfectly-bonded interface can


be represented as an array of circular water lled disbonds with the disbonding
thickness approaching zero. Unfortunately, the exact solution of the periodic distribution of circular cracks is not available. However, when the fractional area of
103

disbonding is low, the cracks are nearly noninteracting. Therefore, Kt can be calculated by assuming the strength per unit length of a cylindrical unit cell containing
a single centered circular crack multiplied by the overlap length and divided by
number of the spring elements as follows:
 2   C E 1
b
ab
1=2
Kt =
2 4N 1  2 a (1:2997Ad

0:9952Ad + 0:6672Ad= 0:43231Ad


3 2

0
:1491Ad=
1:8687ln(1 + Ad= ) 0:4199ln(1
0:0295Ad +
=
1 + Ad
1 2

+0:1407Ad

5=2

1 2

1 2

A1d=2 ))

(5.6)
where Ad =  a ab is the disbonded area and N is the number of the spring elements.
The model of the imperfectly-bonded interface has been developed as an array
of very thin interfacial disbonds lled by water, where the shear direction sti ness
of the spring has been determined by the area ratio between disbonded and bonded
area as Ad, so-called the disbonded area, in Eq. 5.6. By using this relationship,
a new two-dimensional nite element procedure has been developed to simulate
the imperfectly-bonded interface by adding the spring elements as the boundary
conditions between adherend and adhesive bonded layers. The shear sti ness of this
customized spring element is Kt , and the sti ness of the spring element for the other
ve degrees of freedom are assumed to be in nity, and therefore are assumed to be
constrained perfectly in this nite element procedure. In this procedure, one eightnoded isoparametric shell element from the adherend layer has been connected with
three spring elements to another eight-noded isoparametric shell element from the
adhesive bonded layer, as shown in Fig. 5.1. There are, in total, 601 spring elements
connecting these two interfaces to simulate the imperfectly-bonded interface. The
equilibrium sti ness of each spring elements is equal to the Kt in Eq. 5.6.
104
(

)2

4 2

5.2 The One-Dimensional Line Element Model

After the spring model simulated the imperfectly-bonded interface in the previous section, the spring model appears to have some disadvantages. First, the spring
element has a particular e ect which stores more potential energy in the heavy
loading regions than the relatively light loading regions. This causes the displacements to be similar along the adhesive bonding side of the spring elements, just
like a car's suspension system. Also, each spring element acts individually according to the external shear or peel force between the two nodes it is connected to,
which doesn't happen in the real imperfectly-bonded interface. In reality, the occurrence of one small disbonded area will lead to the stress changes in the surrounding
bonded region. Second, the spring elements only connects the nodes between the
adherends and adhesive layer, which causes discontinuities when the environmental
degradation only appears in some portions of the interface. For example, consider if
the imperfectly-bonded interface only happened on one end of the spring element,
while the interface was still perfectly-bonded on the other end. This will cause
the distributed spring element model to show a discontinuous stress distribution,
which does not represent the reality of the imperfectly-bonded interface. In order to
modify the limitations of the spring element and evaluate more complex degradation processes between the adherend and adhesive in an accurate manner, a newly
developed one-dimensional line element model has been proposed in this section to
improve the simulation of the imperfectly-bonded interface.

105

Similar to the circular cracks case [23], the imperfectly-bonded interface leads
to additional deformations between the adherend and adhesive for axial, rotation,
and shear deformation.
From Eq. 5.2, the sti ness Kj represents the j th component strength of the
imperfectly-bonded interface, which becomes weaker as the disbonded area of interface increases. In this study, K is associated with the normal (peel) stresses, similar
to Kt in the spring model, which tend to open the interface, and is in nity to ensure
the continuity of displacement in the normal direction for the water lled disbonded
interface, according to Lavrentyev and Rokhlin [21]. K is associated with the the
shear stresses , similar to the Kn in the spring model, which depend on the crack
area Ad . K is not considered in this research due to those two-dimensional element
models.
Unlike the spring element model using the boundary conditions to connect the
imperfectly-bonded interface, the line element model uses a constitutive relation
between the adherend and adhesive surfaces as follows:
3

"

FN
Ft

"

K3

#"

U3
U2

= 0 K
(5.7)
where FN is normal force per unit length, Ft is the shear force per unit length, U
is the normal strain, and U is the shear strain.
From Eq. 5.1 and Eq. 5.7, the virtual work of the line element can be constructed
as:
"
#
Zy
F
N
 = [U U ] F dy
(5.8)
y
t
2

where y and y are the coordinates of two ends of the line element.
1

106

By substituting Eq. 5.7 into the force FN and Ft , the Eq. 5.8 can be rewritten
as:
=

x1

[U

U2 ]

"

K3

#"

#
U3 dy
U2

(5.9)
0 K
Using a two-point Gaussian integration, the element sti ness matrices of Eq. 5.9
can be formulated.
In order to determine the sti ness of the imperfectly-bonded interface, the deformation of the interface needs to be calculated rst. This is done in a manner
similar to the previous spring element.
x2

 2   d E 1
b
ab
K2 =
(1
:2997A1d=2
2
2 8 1  a

+0:1407Ad= 0:0295Ad + 0:1491A=d


1 + Ad
5 2

1=2

1 2

0:9952Ad + 0:6672Ad= 0:43231Ad


3 2

1:8687ln(1 + Ad= ) 0:4199ln(1


1 2

A1d=2 ))

(5.10)

where db is the length of the line element.


K when no water is present in the disbond voids can be written in terms of K
as:
2  
K =
(5.11)
2 K
3

This newly developed line element which has six nodes, shown in Fig. 5.1, where
nodes 1, 2, and 3 are located on the adhesive layer, and nodes 4, 5, and 6 located
on the adherend has been used to simulate the sti ness of the interface. So, the
Quad-8 shell element joins the 6-node line element along their common intersection
of nodes 4, 5, and 6. Because the imperfectly-bonded interface is virtually in nitely
thin, the coordinates of nodes 4, 5, and 6 actually coincide with nodes 1, 2, and 3,
respectively. By using the appropriate conditions calculated by the above Eq. 5.10
107

and Eq. 5.11, the displacement between the interface connected nodes 1, 2, and 3
and the interface connected nodes 4, 5, and 6 can be represented as:
U3
U2

=
=

uy (y = 0+ ) uy (y = 0

)
x(y = 0 ) x (y = 0 )
+

(5.12)
(5.13)

where uy and x are the normal and shear displacements interpolated from the
nodes 1, 2, and 3 from the lower interface as well as nodes 4, 5, and 6 from the
upper interface.
After using the nite element procedure, the generalized strains and stresses
were recovered at the integration points.
Compared to the spring element, the one-dimensional line element actually
modi es the nite element sti ness matrices of the adhesive joint to represent the
imperfectly-bonded interface movement, which is di erent than the spring element
procedure which uses the boundary conditions to simulate the imperfectly-bonded
interface. This modi cation allows the line element to better simulate the interface under more complicated water di usion conditions. For example, when the
interface degradation occurs only in the area near the end of overlap of the adhesive joint, the stress distributions computed by the spring element procedure have
some discontinuties in both peel stress and shear stress. However, by solving the
same problem, the line element procedure can avoid those stress discontinuities by
allowing the sti ness of the interface to changes along the interface. Unlike the
spring elements only connecting the nodal points, this new line element procedure
connects the entire Quad-8 elements along the imperfectly-bonded interface from
108

the adherends and adhesive through the 6-noded line elements. This prevents the
unrealistic stress concentrations of the spring element analysis.
5.3 Results and Discussion

The spring element and the new one-dimensional line element developed in the
previous section were applied to examine the peel and shear stress distributions
of a single-lap joint with one imperfectly-bonded interface. According to these
results, the new procedure can be further compared to the nite element results
from the previous perfectly-bonded joint in order to study the change of the stress
distribution during the environmental degradation.
5.3.1 The Distributed Spring Element

Due to computing expense, only the single-lap joints with shorter 0:25in overlap
and 0:01in thick FM73 adhesive bonds, and 0:1in thick aluminum adherends were
investigated in this research. Two joints with 0% imperfectly-bonded, or perfectlybonded, and 20% uniform imperfectly-bonded area, Ad , occurring at the upper
interface of the adhesive bonded layer have been evaluated by using the distributed
spring elements, as shown in Fig. 5.2 and Fig. 5.3, respectively. The imperfectlybonded interface in this section was assumed to contain only one disbonded island
size, 0:01 inch. Each joint was loaded by a 1000lbs static uniform loading at the left
end of the joint, and constrained at the right end of the joint. Since this research
focuses on stress changes during the environmental degradation of the adhesive
bonded layer, the peel and shear stress distributions in the upper (y = t), middle
(y = 0), and lower surfaces (y = t) of the adhesive layer are presented.
1
2

1
2

109

The change in the stress distributions after the degradation can be found by
comparing Fig. 5.2 and Fig. 5.3. The peel and shear stresses increase about 23%
near the middle of the overlap area after the environmental degradation occurred as
shown in Fig. 5.3, but decrease 15% near both ends of the overlap area, where the
stress singularities are located. These changes in stress distributions are believed to
be caused by the weaker strength of the imperfectly-bonded layer no longer resisting
the high stresses near the end of the overlap, and shifting the loads to the middle of
the overlap. This activity is even more obvious in the top interface than the bottom
interface, where the maximum shear stress increases 12% more than the bottom
interface. The bottom interface between adhesive and adherend in this simulation
is assumed to be still perfectly-bonded. In addition to the stresses increasing at the
middle of the overlap, the imperfectly-bonded interface made the peel stress and
shear stress distributions become almost a straight line except near both ends of the
overlap after the environmental degradation. This situation is caused by the spring
element storing more potential energy in high stress regions near the end of the
overlap than the other regions which make the deformations in the adhesive bonded
layer equal along the interface, even though the shear or normal forces applied to
the spring elements in the adherend are di erent. However, although the spring
element is questionable to present accurate results for the stresses near the middle
of the interface, it still provides a relatively reasonable stress reduction near the
ends of the overlap, as expected. Since neither the peel nor the shear stresses in the
middle of overlap are higher than the stresses near both ends of overlap, only the
stress singularities (high stress regions) near the ends of the overlap take a major
role in the creep and the failure of the bonded joint. Actually, this is also the place
110

that kissing bond usually rst occurred. This makes the spring element model a
useful tool for the bonded joint reliability prediction.
5.3.2 The One-Dimensional Line Element

In order to improve the previous distributed spring element model, a 20% uniform disbonded area at the imperfectly-bonded upper interface of the adhesive layer
has been recalculated once again by using the newly developed one-dimensional line
element procedure. Similar to the previous spring element case, the individual disbonded island sizes are assumed to be 0:01 inch. The results, shown in Fig. 5.4,
appear to be pretty close to the previous spring element's results. However, the peel
stresses at the middle of the overlap are obviously lower than the spring element
case. This made the peel stress distribution nonlinear. Actually, it is more like the
curved line from the perfectly-bonded interface's results, as shown in Fig. 5.4, but in
a smaller scale. Similar results happened with the shear stress. The middle surface
shear stress distribution near the middle of overlap computed by the line element
is 6% lower than the spring element. Also, the upper and bottom interfaces have
higher shear stress than the spring element case. The explanation is that the line
element utilizes the whole element sti ness matrix directly, which is more accurate
than the spring element. Another issue is that the loads are not shifted to high
stress regions in the line element model, unlike what occurred in the spring element
model. This made the shear and peel stresses near the middle of the interface lower
and more accurate. Clearly, the newly developed line element procedure is believed
to have better ability to simulate the imperfectly-bonded interface not only at the
end of the overlap, but across the whole interface. Further evaluating the stress
111

through the adhesive thickness, Fig. 5.4 shows that the upper interface has larger
e ects from the imperfectly-bonded interface than the bottom interface. From this
interesting observation, the in uence from the upper imperfectly-bonded interface
rst changes the stress distribution of the upper interface, obviously, while the same
time, the maximum peel and shear stresses in the middle interface also begin to drop
but much slowly. However, the stress distribution at the bottom interface which
is away from the imperfectly-bonded interface almost have no e ect. In order to
further study this observation, it becomes necessary to investigate the imperfectlybonded joints under a more complicated situation, such as one containing a di erent
disbonded area percentage in the imperfectly-bonded interface.
So, another case of a single-lap joint with a non-uniform imperfectly-bonded
area at the upper interface has been computed by using the one-dimensional line
element procedure, shown in Fig. 5.5. In this more complex case, the upper interface degradation near the end of the overlap is assumed to reach as much as 30%
disbonded area. The disbonded area decreases while moving close to the middle of
the overlap. Finally, the environmental degradation near the middle of the overlap
is assumed not to have occurred, corresponding to 0% disbonded area. Or it is
still perfectly-bonded. The individual island size is assumed to have a 0:01 inch
diameter throughout the whole adhesive bonded layer. This leads to an interface
containing imperfectly-bonded areas as well as perfectly-bonded areas. This interface happens in the real world particularly when liquid lls the disbonded island [21]
during the environmental degradation. The one-dimensional line element simulates
this imperfectly-bonded interface with lower K and K values (corresponding to
30% disbonded area in this case) near the end, then slowly increases to a higher K
2

112

and K values (corresponding to a 0% disbonded area or perfectly-bonded in this


case) near the middle of the overlap. Fig. 5.5 shows the stress distributions for the
upper, middle, and lower surfaces of the adhesive layer. Interesting but reasonable
results appear for the upper surface, where the shear stresses concentrate at the
ends of the perfectly-bonded area, then slightly reduce to approximately zero near
the end of the overlap when the imperfectly-bonded interface has more disbonded
area Ad. This stress reduction contributes to the imperfectly-bonded area having
lower strength than the perfectly-bonded area. Now, more loads are supported
by the perfectly-bonded interface. Unlike the upper surface, the shear stresses in
the middle surface as well as the lower surface not only concentrate at the ends of
perfectly-bonded area, but also concentrate at the ends of the overlap. Unlike the
upper interface, the rest of the adhesive layer area still has the same strength as the
perfectly-bonded interface, which means the shear stress still concentrates at the
end of the overlap like the perfectly-bonded case. It causes totally four stress concentrations at the bottom interface located near both ends of the overlap, shown in
Fig. 5.5. Unlike the previous constant disbonded area case, these results prove that
the bottom interface also appears to have been e ected by the imperfectly-bonded
interface.
Several interesting observations can be made from the above discussion. First,
it is believed that the distributed spring element and the one-dimensional line element are able to simulate the reduction of the shear stress near both ends of the
overlap of the imperfectly-bonded single-lap joint. Second, the newly developed
one-dimensional line element can perform a better prediction for the shear stress
and peel stress distributions for the uniform imperfectly-bonded interface. Finally,
3

113

only the line element model has the ability to simulate the stress distributions of
the non-uniform imperfectly-bonded interface, unlike the stress discontinuties from
the spring element model.

114

lb

2Cb

lb

t
b

Pb

l
b
= 10 , b = 8
t
cb

Imperfectly-Bonded
Interface

a
Spring Model of Interface

Ad =

( a b) 2
4a 2

The Distributed Spring Element Model


lb

2Cb

lb

x
Pb

6
2

b
l
= 10 , b = 8
t
cb

Imperfectly-Bonded
Interface

3
Line Element of Interface

The One-Dimensional Line Element Model

Figure 5.1: The Geometry and Finite Element Model of the Imperfectly-Bonded
Joints

115

1.5
y=t/2
y=0
y=t/2
xy/P

0.5

0.05

0.1

0.15

0.2

0.25

0.15

0.2

0.25

Cb
4
y=t/2
y=0
y=t/2

/P

2
1
0
1

0.05

0.1
Cb

Figure 5.2: Adhesive Stress Distributions Obtained from the Present Finite Element
Model with Perfectly-Bonded Interface (Uniform Loads)

116

1.5
y=t/2
y=0
y=t/2
xy/P

0.5

0.05

0.1

0.15

0.2

0.25

0.15

0.2

0.25

Cb
3
y=t/2
y=0
y=t/2

/P

1
0
1
2

0.05

0.1
Cb

Figure 5.3: Adhesive Stress Distributions Obtained from the Present Finite Element Model with 20% (Ad =0.2) Imperfectly-Bonded Interface by Using the Spring
Element (Uniform Loads)

117

1.5
y=t/2
y=0
y=t/2
xy/P

0.5

0.05

0.1

0.15

0.2

0.25

0.15

0.2

0.25

Cb
4
y=t/2
y=0
y=t/2

/P

2
1
0
1

0.05

0.1
Cb

Figure 5.4: Adhesive Stress Distributions Obtained from the Present Finite Element
Model with 20% (Ad=0.2) Imperfectly-Bonded Interface by Using the Line Element
(Uniform Loads)

118

1.5
y=t/2
y=0
y=t/2
xy/P

0.5

0.05

0.1

0.15

0.2

0.25

0.15

0.2

0.25

Cb
4
y=t/2
y=0
y=t/2

/P

2
1
0
1

0.05

0.1
Cb

Figure 5.5: Adhesive Stress Distributions Obtained from the Present Finite Element
Model with Various Disbonded Areas by Using the Line Element.(30% (Ad =0.3)
near the End of Overlap Decreases to 0% (Ad=0) in the Center) (Uniform Loads)

119

5.4 The E ect of Disbonded Island Number and Size for


Imperfectly-Bonded Joints

According to the previous analysis, the strength of the imperfectly-bonded interface is highly dependent on the percentage of the disbonded area. However, the
strength of the imperfectly-bonded interface is also a ected by another issue, the
disbonded size, so-called the island size. In this section, the in uence from the disbonded island sizes is considered for the peel and shear stress distributions of the
imperfectly-bonded interface.
By using the relationship from Eq. 5.6, there are two major issues that need to
be considered in predicting the failure of the adhesive bonded joints. First, obviously, increasing the percentage of the disbonded area weakens the adhesive bonded
interface. Second, increasing the size of each individual disbonded island also weakens the strength of the interface as well, as shown in Fig. 5.6. By using these two
conclusions, a newly developed three-dimensional graphic mesh has been created to
represent the maximum peel and shear stresses of the imperfectly-bonded interface
under di erent disbonded island sizes and di erent percentages of disbonded areas.
This new three-dimensional information can be used to determine when the quality
of the bonded interface is no longer safe to resist the loads. Furthermore, it also
can be used as a base for the development of the bonded joint reliability prediction
method and accept/reject inspection criteria.
In Figs. 5.75.9, the maximum peel stresses of the imperfectly-bonded interface
with di erent percentages of the disbonded area and di erent individual disbonded
island sizes have been presented. In the rst gure, Fig. 5.7, it shows the maximum
peel stress at the upper level (y = t ) of the adhesive layer where the imperfect
120
2

High Strength
Small Disbonded Islands
(Ad=0.25)

Low Strength
Large Disbonded Islands
(Ad=0.25)

Figure 5.6: The Strength of the Interface with Various Disbonded Island Sizes
bond occurred. As presented in this gure, the maximum peel stress at this upper
interface of the imperfectly-bonded interface becomes larger during the early stage
of the environmental degradation. This is the result of the weaker strength of the
upper interface due to either the increasing disbonded area or larger individual
disbonded island size. As discussed in the previous section, this weaker interface
creates additional deformation on the overlap of the single-lap joint which directly
result in the higher strain or stress near the end of the interface. This action leads
121

to the maximum peel stresses of the upper surface to begin to slowly increase when
the interface begins to weaken, shown in the upper corner of the Fig. 5.7. Following
this early stage, the maximum peel stresses begin to drop while the disbonded area
or the disbonded island size keeps increasing. This can be named as the second
stage, shown in the lower corner of the Fig. 5.7. In this stage, the upper interface
becomes much weaker and the de ection of the interface is so large that the original
mechanics of the bonded joint has been changed, and the overlap of the single-lap
joint no longer can be treated under the assumptions of linear geometry. Now,
because of the large de ection e ect involved in this peel stress distribution, it
leads the peel stress of the interface to rearrange from both ends of the overlap,
where the high stress concentration regions were located, to other portions of the
interface. Simplifying this activity, it can be said that the peel stresses or strains
begin to shift away from the ends of the overlap. This caused the decreasing of the
maximum peel stress, but the increasing of the minimum peel stress at the same
time. Observing Fig. 5.7, it clearly recon rms this assumption that the maximum
peel stress increase slightly about 7% in the beginning, when disbonded area is
small, then drop tremendously after the disbonded area passes 10% with 1 inch
diameter disbonded island or 78% with 0:1 inches diameter disbonded island.
However, it also leads to another issue in this research that disbonded island
size plays a major role in the strength of the imperfectly-bonded interface. In
Fig. 5.7, the maximum peel stress changes little under 50% disbonded area with
the average 0:01 inch diameter disbonded island size, but the maximum peel stress
becomes only one half of its original value under the same disbonded area with the
average 1 inch diameter disbonded size. It appears that the interface under the
122

same disbonded area containing the larger size of the disbonded islands and fewer
numbers of disbonded islands has much weaker strength, compared to the small
size of the disbonded island or large numbers of the disbonded islands. Therefore,
future inspection criteria for bonded joints needs to consider not only the amount
of disbonded area, but also the size of each individual disbonded island.
In order to further investigate the in uence of the disbonded island size, the
maximum peel stresses located at the middle and bottom levels of the adhesive
bonded layer containing an upper imperfectly-bonded layer in the single-lap joint
are also investigated in this section. The Fig. 5.8 shows the maximum peel stresses
at the middle level of the interface (y = 0). In this gure, similar to the upper level,
the maximum peel stresses initially increase when the interface becomes weaker,
then later drop when the interface keeps growing weaker. Unlike the upper level
peel stresses, the maximum peel stresses do not begin to drop until the disbonded
area reaches 50% with 0:1 inches diameter disbonded islands or 18% with 1 inch
diameter disbonded islands. This can be explained since the maximum peel stresses
of the middle level did not reach as high as those at the upper interface during the
early stage of the imperfectly-bonded interface degradation process. As mentioned
in the previous section, the maximum peel stress of the adhesive is either in the
upper right or the lower left corner of the adhesive bond of the single-lap joint. So,
in this case, the peel stress at the upper right corner is the rst one to reach a
point which the deformation is so large that the peel stress has been redistributed
to other areas within the interface, which includes the middle level of the interface.
Therefore, the peel stresses in the middle level is still increasing at this stage. The
middle level even needs to resist the load coming from the weakened upper interface.
123

Later in the higher percentage of disbonded area or larger disbonded size, 21% with
1 inch diameter island or 87% with 0:1 inches diameter island, the deformation in
the middle level also reaches the point that the right end of the middle level adhesive
layer is no longer able to resist the loading, and begin to transfer to the other regions
of the middle level and lower interface. It also appears that the imperfectly-bonded
interface has more in uence in the stress distribution at the upper layer than the
middle layer which is away from the imperfectly-bonded interface.
In order to further con rm this theory, the lower level of the adhesive bonded
layer has been presented in Fig. 5.9. Similar to the middle level case, the maximum
peel stresses at the lower level of interface showed the maximum peel stresses do not
drop until reaching 40% disbonded area with the 1 inch diameter disbonded island
or 90% with the 0:1 inches diameter disbonded island. Therefore, the maximum
peel stresses of the lower level of the overlap do not have much in uence from the
weaker upper interface, compared to the previous upper and middle level cases. This
is because the bottom interface is even further away from the imperfectly-bonded
interface. Even with the 1 inch diameter disbonded island size, the maximum peel
stresses at the bottom interface increase only 3:8% at the 46% disbonded area,
compared to the stresses at the perfectly-bonded interface, then slowly decrease
2:2% before reaching 80% disbonded area.
In addition to the maximum peel stresses, the maximum shear stresses in the
interface also have been presented in Figs. 5.105.12. Similar to the previous peel
stress cases, the maximum shear stresses become higher while the disbonded area
or disbonded island size is increasing. Then, they begin to drop after reaching an
ultimate point like the previous peel stress case. Once again, the disbonded island
124

size still heavily in uences on the results of the maximum shear stress. Fig. 5.10
displays the maximum shear stresses at the upper level of the interface. This gure
clearly shows that the maximum shear stress initially increases then slightly drops
when the disbonded island size is small (0:01 in). But unlike the peel stress case, the
maximum shear stresses almost immediately decreases shortly after the disbonded
interface occurs when the disbonded island size is larger than 0:05in. This is caused
by the imperfectly-bonded interface having an even weaker bond to resist the shear
stress, since the shear sti ness is smaller than the normal sti ness. This can be
seen in Eq. 5.11 where Kt is lower than Kn. Also, the tension applied at the left
end of the lap causes much higher shear stress than peel stress in the adhesive
bonded layer. Therefore, the extra shear deformation already changes the shear
stress distribution, even in the small disbonded area or disbonded island size cases.
In this gure, the maximum shear stresses drop 36% before the disbonded area
reaches 55% with 0:1 inch diameter disbonded island or 25% with 1 inch diameter
disbonded island size.
The maximum shear stresses at the middle level of the interface have not changed
too much during the upper interface degradation process, shown in Fig. 5.11. The
change of the maximum shear stresses at the middle level of the interface is similar
to the peel stress results, but on a smaller scale. The shear stress increases 2%
when the disbonded area reaches 35% with 1 inch diameter disbonded island or
80% with 0:1 inches diameter disbonded island. Then, it decreases back to almost
its original amount when the imperfectly-bonded interface with 80% disbonded
area increases its island size to 1 inch diameter. This indicates that the in uence
from the imperfectly-bonded interface is not so obvious for those areas which are
125

away from the imperfectly-bonded interface within the adhesive bonded layer. The
next gure, Fig. 5.12, presents the maximum shear stresses at the lower level of the
interface. The maximum shear stresses of this level increase like the previous results
when the disbonded area or disbonded island size is increasing, but do not decrease
while the disbonded area or island size keeps increasing. Actually, the results show
the maximum shear stresses keep increasing up to 16% at small disbonded island
size (0:01in) or up to 11% at large disbonded island size (1:0in) before reaching
80% disbonded area. The reason is that the lower level of the interface has little
additional deformation caused by the imperfectly-bonded interface, as discussed
previously. Because it is away from the imperfectly-bonded interface, the maximum
shear stresses at the lower level are not a ected by the environmental degradation
of the upper interface.
According to these maximum stress results, the strength of the imperfectlybonded interface is very sensitive to each individual disbonded island size. Therefore, the imperfectly-bonded interface with smaller disbonded island size can sustain
a higher percentage of the disbonded area than those with larger disbonded islands.
Using this conclusion, a possible design or inspection criteria of imperfectly-bonded
interfaces can be proposed by considering both its disbonded area and each individual disbonded island size. For this criteria, the author assumes that when the additional deformation in a particular portion of structure caused by the imperfectlybonded interface begins to change the original stress distribution of the structure,
it means the object is not longer reliable and safe. In this research, this is considered to happen when the maximum peel and shear stresses begin to drop. In
other words, when the peel or shear stress of the single-lap joint begins shifting.
126

This point can be located in the three-dimensional graphic meshes when the slope
of the mesh is equal to zero. Using the previous assumptions, it is believed that the
joints' reliability and safety is questionable when all three levels (y = t ; 0; t ) of
either maximum peel or shear stresses begin to decrease or the slope of the mesh is
zero or becomes negative. However, a more detailed de nition of this criteria can
be made after the tension tests of the various imperfectly-bonded joints have been
conducted in the future. The ultimate goal of this criteria is to provide the eld
engineers a procedure to acknowledge the reliability of imperfectly-bonded singlelap joints by evaluating the average disbonded island size and the percentage of the
disbonded area. Furthermore, based on the material properties and geometry of
the structures, more detailed and accurate three-dimensional graphical information
can be customized for each individual adhesive material as a design or inspection
criteria of the bonded structures in the future.
2

127

Stress Ratio y/P

4
3
2
1
0
0
0

0.2
0.4

0.2
0.6

0.4
0.8

0.6
0.8

Disbonded Island Size

Disbond %

Figure 5.7: Maximum Peel Stress Distributions at the Upper Level (y = t ) of the
Bonded Interface along the Percentage of Disbonded Area and Disbonded Island
Size (inches)
2

128

1.2

Stress Ratio y/P

1
0.8
0.6
0.4
0.2
0

0.1
0.2

0.2
0.3

0.4

0.4
0.6

0.5
0.6

0.8

0.7
0.8

Disbonded Island Size

Disbond %

Figure 5.8: Maximum Peel Stress Distributions at the Middle Level (y = 0) of the
Bonded Interface along the Percentage of Disbonded Area and Disbonded Island
Size (inches)

129

Stress Ratio y/P

3.8

3.6

3.4
0

3.2
0

0.2
0.2

0.4
0.4

0.6
0.6

0.8
0.8

Disbonded Island Size

Disbond %

Figure 5.9: Maximum Peel Stress Distributions at the Lower Level (y = t ) of the
Bonded Interface along the Percentage of Disbonded Area and Disbonded Island
Size (inches)
2

130

1.4
1.2

Stress Ratio

xy

/P

1
0.8
0.6
0.4
0.2
0
0

0
0.2

0.2

0.4
0.4

0.6
0.6

0.8
0.8

Disbonded Island Size

Disbond %

Figure 5.10: Maximum Shear Stress Distributions at the Upper Level (y = t ) of the
Bonded Interface along the Percentage of Disbonded Area and Disbonded Island
Size (inches)
2

131

0.61
0.6

Stress Ratio

xy

/P

0.59
0.58
0.57
0.56
0.55
0.54

0.53
0

0.2
0.4

0.2
0.6

0.4
0.8

0.6
0.8

Disbonded Island Size

Disbond %

Figure 5.11: Maximum Shear Stress Distributions at the Middle Level (y = 0) of


the Bonded Interface along the Percentage of Disbonded Area and Disbonded Island
Size (inches)

132

0.95

Stress Ratio

xy

/P

0.9
0.85
0.8
0.75
0
0.7
0

0.2
0.4

0.2
0.6

0.4
0.8

0.6
0.8

Disbonded Island Size

Disbond %

Figure 5.12: Maximum Shear Stress Distributions at the Lower Level (y = t ) of


the Bonded Interface along the Percentage of Disbonded Area and Disbonded Island
Size (inches)
2

133

CHAPTER 6
CONCLUSIONS

A new passive damping concept based on magneto-elastic e ects from magnetomechanical coating materials is proposed for vibratory stress and amplitude
reduction on vibrating beams and blades. A model for the vibration of beams
and blades containing a magnetic coating layer is presented. This is based on the
magneto-elastic theory [30, 31] developed to formulate the mechanical energy dissipation from the irreversible movement of magnetic domain boundaries of the coating
material.
The steady state responses of coated beams and blades are computed under a
harmonic excitation force. The e ect of the magnetic coating layer has been successfully examined and compared in this forced vibration analysis. The closed-form
solutions and the nite element results show a signi cant vibratory stress and amplitude reduction of the vibrating beams and blades with one coating layer. The
e ects of the coating layer on the vibratory energy dissipation are found to be sensitive to the level of the strain or stress in the coating layer and independent of
the vibration frequency. Additionally, several case studies have been presented to
demonstrate the ability of this magnetomechanical coating with various thicknesses
to suppress vibration in the high rotating speed environment. The results proved
134

that the maximum von Mises stress can be reduced 8  36% by adding one coating
layer. The present model could be extended to real turbine blades with twisted
non-uniform cross-sections and to also account for the e ects of aerodynamic, and
thermal loading. Furthermore, the in uence of the di erent host materials, such
as nonisotropic materials, on the stress-dependent damping capacity of the magnetomechanical coating materials will need to be investigated in order to improve
the accuracy of the simulation results. Also, the thermal coecient of expansion
and the creep e ect of this magnetomechanical coating material need to be further
studied to insure the reliability of the coating material during the high temperature
and rotating speed operations in the turbine engine.
Like the blade vibration, another structural problem, the imperfectly-bonded
adhesive joint, that seriously a ects aircraft safety and reliability has been further
investigated. A two-dimensional nite element procedure for double-strap joints
and single-lap joints was developed by using the Quad-8 isoparametric shell element.
Due to the stress singularities near the ends of the overlap, the number of elements
was tripled in order to have more accurate solutions. The comparison between this
two-dimensional nite element procedure and Goland and Reissner's, Hart-Smith's,
and Oplinger's analytical solutions for the single-lap joint and Oplinger's analytical
solution for the double-strap joint are presented and proved the accuracy of this
new nite element procedure.
The further development of a new two-dimensional nite element procedure to
simulate the environmental degradation of the bonded layer has been successful.
Based on Baik and Thompson [3], Margetan, Thompson, and Gray [23], and Laventyev and Rokhlin's [21] previous research, a nite element procedure, using the
135

distributed spring elements, has been created and shows the peel and shear stress
reduction at both ends of the overlap of the single-lap joint after the environmental degradation occurred. The spring model is later improved and replaced by a
newly developed one-dimensional line element. These line elements successfully
improve the simulation of the imperfectly-bonded interface between the adherend
and the adhesive. This element is also able to predict the shear stress as well
as the peel stress due to the environmental degradation in a more accurate manner. Also, the line element was extended to predict the partial imperfectly-bonded
interface. From these results, newly developed three-dimensional graphic design diagrams have been created as a possible design or inspection criteria for the adhesive
bonded joints. The results show the strength of the imperfectly-bonded adhesive
joint is highly dependent on not only the percentage of disbonded area it contains,
but also each individual disbonded island size. In the future, after conducting more
tension tests of the adhesive joints, the information for the stress propagation from
the imperfectly-bonded interface can be determined to create a more reliable design
and inspection criteria. This information can be further used as a foundation to
detect and avoid the premature failure of bonded joints and improve the reliability
and quality of bonded joints. Also, similar procedures can be repeated to create
the design or inspection criteria for the di erent adhesive joint structures.

136

APPENDIX A
Collocation method

The method of collocation is a procedure that approximates the solution of a


di erential equation, like the equation of motion in Eq. 2.23 of this research, as a
function, such as a polynomial or sine function, by reducing the di erences between
the solution and the approximated function.
In the simply supported beam case, a sine function has been chosen as our
approximation as:
w(x) = a0 sin(

nx
nx
nx
)
+
a1 cos(
)
+
a2 sinh(
2l
2l
2l )
+a3cosh( nx
2l )

(A.1)

Also, the equation of motion for the simply-supported beam, Eq. 2.23, can be
written as:
f (x; w(x)) = EI (1 + is )wiv +

Kh 4(h + 2d)d2 00 iv


000
[
(
w w +w
2
2
i

+ 8(h + 2d)d w00 w000 + 8hd



3

0002

! 2 b(2d + h)w

137

+ 4hd wiv ]e
i

2dw

00

i

(A.2)

Suppose that there is a polynomial function as P (x) which can be represented


as:
P (x) = a0 + a1 x + ::: + an xn

(A.3)

Or actually using Eq A.1 as the P (x). In order to determine P (x), the expression
is considered as:
e(x) = f (x; p(x)) F (x)

(A.4)

where F(x) is the exciting forcing function.


The object of the error function e(x) is to nd the coecients ai of P (x) to make
this function as small as possible. One easy method to achieve this goal is to use
the least-squares method to minimize the expressions as:
Z

minf

l
0

[e(x)] dxg
2

(A.5)

Due to the nonlinear properties of Eq. 2.23, Eq. A.2 is too complicated to nd
its global minimum position. So, rather than using Eq. A.2, the method of collocation proceeds by asking that p(x) be chosen so that e(ti ) = 0 at certain selected
"collocation points", t , t , t , ... , tn . Instead of minimizing Eq. A.2, the solution
can be computed by solving n + 1 equations as follows:
0

e(t0 )

= 0
e(t ) = 0
1

::::::

138

e(tn )

= 0

(A.6)

By solving this nonlinear system of n + 1 equations in n + 1 unknowns, the


approximated solution of the equation of motion can be represented as the steady
state response of the coated system. In practice, the initial condition of P (x) is
very important to the convergence speed of the collocation method. If P (x) has the
correct starting value, Eq. A.6 will be solved in a very short period. Otherwise, it
will take long time to reach a convergence point, and even sometimes fail.

139

APPENDIX B
The Abbreviated MSC/NASTRAN Routine for the
Cantilevered Blade with the Magnetomechanical Coating

The following abbreviated routine is written in the popular nite element package MSC/NASTRAN to evaluate the magnetomechanical coating e ect on a cantilevered blade.
SOL 103
TIME = 60
CEND
TITLE = Turbine Blade
SUBTITLE = Magnetomechanical Coating
$ GLOBAL CASE
SPC = 1
DISPLACEMENT = ALL
STRESS = ALL
DLOAD = 10
FREQUENCY = 17
METHOD = 88
BEGIN BULK
PARAM AUTOSPC YES
PARAM POST -2
EIGR, 88, SINV, 17000., 18000.
RLOAD1, 10, 22, 32
DAREA, 22, 6, 2, 200.
TABLED1, 32 +
+ 2., 1., 16., 1., ENDT
GRID, 1, 0, 3.00000, -2.E-17, 7.00000,, 0
GRID, 2, 0, 3.03271, -2.5E-2, 7.00000,, 0
GRID, 3, 0, 3.05428, -6.0E-2, 7.00000,, 0

140

GRID, 4, 0, 3.06180,-0.10000, 7.00000,, 0


GRID, 5, 0, 3.05428,-0.14033, 7.00000,, 0
GRID, 6, 0, 3.03271,-0.17523, 7.00000,, 0
GRID, 7, 0, 3.00000,-0.20000, 7.00000,, 0
GRID, 8, 0, 2.83444, 0.11406, 7.00000,, 0
GRID, 9, 0, 2.86343, 8.32E-2, 7.00000,, 0
GRID, 10, 0, 2.88209, 4.31E-2, 7.00000,, 0
GRID, 11, 0, 2.88758, -2.3E-3, 7.00000,, 0
GRID, 12, 0, 2.87899, -4.8E-2, 7.00000,, 0
..................................
CHEXA, 97,, 250, 1, 2, 9, 8, 114, 138+
+, 306, 282
CHEXA, 98,, 250, 114, 138, 306, 282, 115, 139+
+, 307, 283
CHEXA, 99,, 250, 115, 139, 307, 283, 116, 140+
+, 308, 284
CHEXA, 100,, 250, 116, 140, 308, 284, 117, 141+
+, 309, 285
CHEXA, 101,, 250, 117, 141, 309, 285, 118, 142+
+, 310, 286
CHEXA, 102,, 250, 118, 142, 310, 286, 119, 143+
+, 311, 287
CHEXA, 103,, 250, 119, 143, 311, 287, 120, 144+
+, 312, 288
CHEXA, 104,, 250, 120, 144, 312, 288, 121, 145+
+, 313, 289
CHEXA, 105,, 250, 121, 145, 313, 289, 122, 146+
+, 314, 290
CHEXA, 106,, 250, 122, 146, 314, 290, 123, 147+
+, 315, 291
CHEXA, 107,, 250, 123, 147, 315, 291, 124, 148+
+, 316, 292
CHEXA, 108,, 250, 124, 148, 316, 292, 125, 149+
+, 317, 293
...................................
CPENTA, 2257, 250, 106, 107, 113, 2634, 2658, 2802
CPENTA, 2258, 250, 2634, 2658, 2802, 2635, 2659, 2803
CPENTA, 2259, 250, 2635, 2659, 2803, 2636, 2660, 2804
CPENTA, 2260, 250, 2636, 2660, 2804, 2637, 2661, 2805
CPENTA, 2261, 250, 2637, 2661, 2805, 2638, 2662, 2806
CPENTA, 2262, 250, 2638, 2662, 2806, 2639, 2663, 2807
CPENTA, 2263, 250, 2639, 2663, 2807, 2640, 2664, 2808
CPENTA, 2264, 250, 2640, 2664, 2808, 2641, 2665, 2809
CPENTA, 2265, 250, 2641, 2665, 2809, 2642, 2666, 2810
CPENTA, 2266, 250, 2642, 2666, 2810, 2643, 2667, 2811
CPENTA, 2267, 250, 2643, 2667, 2811, 2644, 2668, 2812
CPENTA, 2268, 250, 2644, 2668, 2812, 2645, 2669, 2813
CPENTA, 2269, 250, 2645, 2669, 2813, 2646, 2670, 2814
..........................................

141

CQUAD4, 2401, 501, 1, 8, 282, 114, 0.00000, -8.00E-3


CQUAD4, 2402, 502, 8, 15, 450, 282, 0.00000, -8.00E-3
CQUAD4, 2403, 503, 15, 22, 618, 450, 0.00000, -8.00E-3
CQUAD4, 2404, 504, 22, 29, 786, 618, 0.00000, -8.00E-3
CQUAD4, 2405, 505, 29, 36, 954, 786, 0.00000, -8.00E-3
CQUAD4, 2406, 506, 36, 43, 1122, 954, 0.00000, -8.00E-3
CQUAD4, 2407, 507, 43, 50, 1290, 1122, 0.00000, -8.00E-3
CQUAD4, 2408, 508, 50, 57, 1458, 1290, 0.00000, -8.00E-3
CQUAD4, 2409, 509, 57, 64, 1626, 1458, 0.00000, -8.00E-3
CQUAD4, 2410, 510, 64, 71, 1794, 1626, 0.00000, -8.00E-3
CQUAD4, 2411, 511, 71, 78, 1962, 1794, 0.00000, -8.00E-3
CQUAD4, 2412, 512, 78, 85, 2130, 1962, 0.00000, -8.00E-3
...........................................
MAT1, 101, 1.05E+07, 4.00E+06, 0.3, 7.33E-04, 6.75E-06, 71.33,
MAT1, 102, 1.05E+07, 4.00E+06, 0.3, 7.33E-04, 6.75E-06, 71.33,
MAT1, 103, 1.05E+07, 4.00E+06, 0.3, 7.33E-04, 6.75E-06, 71.33,
MAT1, 104, 1.05E+07, 4.00E+06, 0.3, 7.33E-04, 6.75E-06, 71.33,
MAT1, 105, 1.05E+07, 4.00E+06, 0.3, 7.33E-04, 6.75E-06, 71.33,
MAT1, 106, 1.05E+07, 4.00E+06, 0.3, 7.33E-04, 6.75E-06, 71.33,
MAT1, 107, 1.05E+07, 4.00E+06, 0.3, 7.33E-04, 6.75E-06, 71.33,
MAT1, 108, 1.05E+07, 4.00E+06, 0.3, 7.33E-04, 6.75E-06, 71.33,
MAT1, 109, 1.05E+07, 4.00E+06, 0.3, 7.33E-04, 6.75E-06, 71.33,
MAT1, 110, 1.05E+07, 4.00E+06, 0.3, 7.33E-04, 6.75E-06, 71.33,
MAT1, 111, 1.05E+07, 4.00E+06, 0.3, 7.33E-04, 6.75E-06, 71.33,
MAT1, 112, 1.05E+07, 4.00E+06, 0.3, 7.33E-04, 6.75E-06, 71.33,
.........................................
MAT1, 2, 1.71E+7, 6.38E+6, 0.34000, 4.14E-4, 6.75E-6, 71.3300,
PSHELL, 501,101,8.00E-03,101,,101
PSHELL, 502, 102, 8.00E-03, 102,, 102
PSHELL, 503, 103, 8.00E-03, 103,, 103
PSHELL, 504, 104, 8.00E-03, 104,, 104
PSHELL, 505, 105, 8.00E-03, 105,, 105
PSHELL, 506, 106, 8.00E-03, 106,, 106
PSHELL, 507, 107, 8.00E-03, 107,, 107
PSHELL, 508, 108, 8.00E-03, 108,, 108
PSHELL, 509, 109, 8.00E-03, 109,, 109
PSHELL, 510, 110, 8.00E-03, 110,, 110
PSHELL, 511, 111, 8.00E-03, 111,, 111
PSHELL, 512, 112, 8.00E-03, 112,, 112
.....................................
PSOLID, 250, 2, 0
SPC,
SPC,
SPC,
SPC,
SPC,

1,
1,
1,
1,
1,

137,
161,
185,
209,
233,

123456,
123456,
123456,
123456,
123456,

0.00000
0.00000
0.00000
0.00000
0.00000

142

7.43E-02
1.35E-02
5.10E-03
2.70E-03
1.70E-03
1.30E-03
1.00E-03
9.00E-04
8.00E-04
8.00E-04
8.00E-04
8.00E-04
0.00500

SPC, 1, 257, 123456, 0.00000


SPC, 1, 281, 123456, 0.00000
SPC, 1, 305, 123456, 0.00000
SPC, 1, 329, 123456, 0.00000
SPC, 1, 353, 123456, 0.00000
SPC, 1, 377, 123456, 0.00000
SPC, 1, 401, 123456, 0.00000
.......................
FREQ, 17, 17508.2
ENDDATA

143

APPENDIX C
Dynamics of the Rotating System

The rotating turbine blade in this research is modeled under the assumption that
the coordinate system used in the description of the blade will rotate at a constant
speed about a xed axis. Because the rotating blade is accelerating relative to a
previous stationary inertial blade, the inertial or mass impedance cannot be directly
calculated in the rotating blade. The inertial impedance in the rotating blade must
rst be determined in the stationary blade, and then transformed to the rotating
blade. Following the procedures in two advanced dynamics books [5] and [17], the
blade coordinates from the stationary system rst need to be transferred to the
rotating system in order to simulate the rotating blade, which can be written as:
8
9 2
9
38
>
< u(t)r >
= 6 cos
t sin
t 0 7 >
< u(t)s >
=
v
(
t
)
sin

t
cos

t
0
v
(
t
)
=
(C.1)
4
5
r
s
>
>
>
: w(t)r >
;
:
;
0
0 1
w(t)s
where
is the rotating speed, u, v, and w are the coordinates from the threedimensional system, r represents the rotating system, and s represents the stationary system.
By using the exponential function to replace the cos
t and sin
t, Eq. C.1 is
rewritten as:
8
9 0
9
2
3
2
3 2
31 8
>
< u(t)r >
= B 1 i 6 1 i 0 7 1 i 6 1 i 0 7 6 0 0 0 7C >
< u(t)s >
=
v
(
t
)
i
1
0
i
1
0
0
0
0
v
(
t
)
=
e
+
e
+
(C.2)
@
4
5
4
5
4
5
A
s
>
>
>
2 000 2
: w(t)rr >
;
:
;
0 00
001
w(t)s
144

where  is
t.
De ning three matrices in Eq. C.2 as the transformation matrix as R , R , and
R , the original vector from the stationary system also can be presented in terms of
the vector from the new rotating system by multiplying the inverse transformation
matrix on both sides of the Eq. C.2, shown as follows:
8
9
8
9
>
>
u
(
t
)
r >
< u(t)s >
= 1 i
<
1 e i [R ] + [R ]) v(t) =
v
(
t
)
=
(
e
[
R
]
+
(C.3)
>
>
2
: w(t)ss >
; 2
: w(t)rr >
;
In order to simplify the problem to show the inertial forces, only the vector
u(t) has been used in the derivative process. First, taking the second derivative of
Eq. C.3 with respect to time we obtain
d [u(t)s ]
1 ei [R ] + 1 e i [R ] + [R ]) d [u(t)r ]
=
(
dt
2
2
dt
+i2
( 12 ei [R ] 12 e i [R ]) d[udt(t)r ]
( 21 ei [R ] + 12 e i [R ])[u(t)r ] (C.4)
1

Applying the transformation matrix to the stationary system acceleration vector


given by Eq. C.4, the inertial system acceleration de ned in the rotating system
coordinates will result as
!
d [u(t)s ]
1 ei [R ] + 1 e i [R ] + [R ])( 1 ei [R ] + 1 e i [R ] + [R ]) d [u(t)r ]
=
(
dt
2
2
2
2
dt
r
+i2
( 21 ei [R ] + 12 e i [R ] + [R ])( 12 ei [R ] 21 e i [R ]) d[udt(t)r ]

( 12 ei [R ] + 21 e i [R ] + [R ])( 12 ei [R ] + 12 e i [R ])[u(t)r ] (C.5)


2

Eq. C.5 can be reduced as:


!
d [u(t)s ]
d [u(t)r ]
d[u(t)r ]
=
4

Re(i[R ])
2
Re([R ])[u(t)r ]
dt
dt
dt
2

(C.6)

This transformation consists of two more terms, which are 4


Re(i[R ]) representing the Coriolis acceleration and 2
Re([R ]) referred to as the centrifugal
145
1

acceleration. This new equation represents the inertial system acceleration in terms
of rotating system coordinates. In addition, the inertial force can be written as:
!
d[u(t)r ]
d [u(t)r ]
F (t) = [M ]
4
Re(i[R ])
2
Re([R ])[u(t) ] (C.7)
2

in

dt

dt

where F (t)in is the inertial force, and M is the mass of the turbine blade.
In Eq. C.7, the rst term obviously represents the blade acceleration. However,
the second and third terms are caused by the rotating movement. In the nite
element method, these additional terms in Eq. C.7 are always present on the rotating
blade as an additional sti ness-like term, so called the spin sti ening e ect, written
as follows:
K = dFduin
(C.8)
The general sti ness matrix can be rewritten as:
Kr = Ks + K

(C.9)

where Kr is the sti ness matrix for the rotating system, and Ks is the sti ness
matrix for the stationary system.
The forced vibration problem is terms of the nodal displacements u is expressed
as
[Kr ]u ! [M ]u = F
(C.10)
2

146

APPENDIX D
DMAP Routine for the Centrifugal Force on the Rotating
System

A DMAP code [6] has been customized for calculating the additional centrifugal
forces for rotating cantilevered blades in a nite element package, MSC/NASTRAN.
echooff
COMPILE SESTATIC
alter 'enddo.*hsflag'
TYPE DB SLIST
TYPE PARM,,I,N,lpflg1a=0,NOUP
TYPE PARM,,I,N,SCNDRY=+1,EXTRN=+1
TYPE PARM,,I,Y,DEPEND= 0
TYPE PARM,,RS,N, OMEGA= 0.
TYPE PARM,,CS,N, SCALE
FILE HISTS=APPEND
MATGEN ,/TMPMGG/1/2
MATGEN ,/TMPBGG/1/2
MATGEN ,/TMPKGG/1/2 RESTART TMPMGG,,/HISTS
RESTART TMPBGG,,/HISTS
RESTART TMPKGG,,/HISTS
SEID=-1
DO WHILE ( lpflg1a >= 0 )
IF ( RSONLY ) THEN
lpflg1a=-1
NOUP=-1
SEID=0
ELSE
SEP2DR SLIST,EMAP//S,N,SEID/S,N,PEID//S,N,lpflg1a/////
S,N,NOUP/S,N,SCNDRY/S,N,EXTRN/S,N,NOMR/'ALL'
ENDIF
MESSAGE //' '/

147

MESSAGE //' USER DMAP INFORMATION MESSAGE - DIFFERENTIAL'/


MESSAGE //' STIFFNESS AND GYROSCOPIC EFFECT GENERATION'/
MESSAGE //' AND ASSEMBLY FOR SUPERELEMENT'/ SEID
MESSAGE //' '/
IF ( SCNDRY = -1 ) THEN
MESSAGE //' '/
MESSAGE //' USER DMAP WARNING MESSAGE - SECONDARY'/
' SUPERELEMENTS'/
MESSAGE //' MUST BE IDENTICAL ELEMENTS (NOT MIRROR) '/
' WITH LOADING'/
MESSAGE //' AND DISPLACEMENTS IDENTICAL TO THE PRIMARY'/
' SUPERELEMENT'/
MESSAGE //' '/
ELSE
CALL SETQ CASES//SEID/PEID/S,MTEMP/S,K2GG/S,M2GG/
S,B2GG/S,MPC/S,SPC/S,LOAD/S,DEFORM/S,TEMPLD/
S,P2G/S,DYRD/S,METH/S,MFLUID
PARAML GPECT//'PRES'////S,N,NOGPECT
IF ( NOGPECT >= 0 ) THEN
CALL CORICEN SLT,BGPDTS,EQEXINS,CSTMS,SILS,MJJ/
KCJJ,BRJJ/S,OMEGA
IF ( OMEGA = 0. ) THEN
MESSAGE //' '/
MESSAGE //' USER WARNING: NO RFORCE SPECIFIED,'/
MESSAGE //' GYROSCOPIC TERMS WERE NOT'/
MESSAGE //' ADDED TO STRUCTURAL MATRICES.'/
MESSAGE //' '/
ENDIF
EMG EST,CSTMS,MPTS,DIT,,UG,ETT,EDT,,,BGPDTS,GPSNTS,,,,/
KDELM,KDDICT,,,,/
1/0/0//1//TEMPLD/DEFORM//////1/
EMA GPECT,KDDICT,KDELM,BGPDTS,SILS,CSTMS,,/KTJJ,/
PARAML KTJJ//'NULL'////S,N,NOKTJJ
IF ( NOKTJJ < 0 ) THEN
MESSAGE //' '/
MESSAGE //' USER WARNING: NO LOADING SPECIFIED,'/
MESSAGE //' DIFFERENTIAL STIFFNESS TERMS'/
MESSAGE //' WERE NOT ADDED TO STRUCTURAL'/
MESSAGE //' MATRICES.'/
MESSAGE //' '/
ENDIF
ADD5 KTJJ,KCJJ,,,/KJJGYRO/
PARAML KJJGYRO//'NULL'////S,N,NOGYRO
IF ( NOGYRO < 0 ) JUMP ENDLOOP
IF ( (DEPEND < 0 ) AND (OMEGA <> 0.) ) THEN
SCALE= CMPLX( -1.0/(OMEGA*OMEGA), 0. )
ADD MJJ,KJJGYRO/MJJTEMP//SCALE
EQUIVX MJJTEMP/MJJ/ALWAYS
EQUIVX MJJ/MJJTEMP/NEVER
SCALE= CMPLX( 1.0/OMEGA )
ADD BJJ,BRJJ/BJJTEMP//SCALE

148

EQUIVX BJJTEMP/BJJ/ALWAYS
EQUIVX BJJ/BJJTEMP/NEVER
MESSAGE //' '/
MESSAGE //' USER INFORMATION: SPEED DEPENDENT TERMS'/
' ADDED'/
MESSAGE //' TO STRUCTURAL MATRICES'/
MESSAGE //' '/
ELSE IF ( (DEPEND < 0) AND (OMEGA = 0.) ) THEN
MESSAGE //' '/
MESSAGE //' USER WARNING: SPEED DEPENDENCE REQUESTED,'/
MESSAGE //' BUT NO RFORCE SPECIFIED.'/
MESSAGE //' DIFFERENTIAL STIFFNESS OR'/
MESSAGE //' GYROSCOPIC TERMS WERE NOT '/
MESSAGE //' ADDED TO STRUCTURAL MATRICES.'/
MESSAGE //' '/
ELSE
ADD KJJ,KJJGYRO/KJJTEMP/
EQUIVX KJJTEMP/KJJ/ALWAYS
EQUIVX KJJ/KJJTEMP/NEVER
ADD BJJ,BRJJ/BJJTEMP/
EQUIVX BJJTEMP/BJJ/ALWAYS
EQUIVX BJJ/BJJTEMP/NEVER
MESSAGE //' '/
MESSAGE //' USER INFORMATION: CONSTANT SPEED TERMS'/
' ADDED'/
MESSAGE //' TO STRUCTURAL MATRICES'/
MESSAGE //' '/
ENDIF
ENDIF
ENDIF
RESTART ,,,HISTS/TLIST/TRUE
ENDDO
lpflg1a>=0 COMPILE CORICEN
SUBDMAP CORICEN
SLT,BGPDT,EQEXIN,CSTM,SIL,MJJ/KCEN,BCOR/OMEGA
TYPE
PARM,,I,N,SLTRECNO T Y P E P ARM; ; RS; N; OM E GA
P ARM; ; C S; N; N 1C; N 2C; N 3C TYPE PARM,,CS,N,OMEGAC
SLTRECNO= 1 DO WHILE ( SLTRECNO > 0 )
PARAML SLT//'DTI'/S,N,SLTRECNO/1//S,N,LTYPE/
IF ( (SLTRECNO <> -1 ) AND ( LTYPE = 10 ) ) THEN
PARAML SLT//'DTI'/SLTRECNO/4//S,N,CID/
PARAML SLT//'DTI'/SLTRECNO/5/S,N,RPT/
PARAML SLT//'DTI'/SLTRECNO/6/S,N,N1/
PARAML SLT//'DTI'/SLTRECNO/7/S,N,N2/
PARAML SLT//'DTI'/SLTRECNO/8/S,N,N3/
SLTRECNO = -2
ENDIF
SLTRECNO = SLTRECNO + 1
ENDDO
IF ( RPT = 0. ) THEN

149

OMEGA= 0.
RETURN
ENDIF
MATGEN ,/I1X1/1/1/
MATGEN ,/I6X6/1/6/
DIAGONAL MJJ/MJJC/'COLUMN'
IF (N1 <> 0.) THEN
MATGEN ,/PVC6/6/6/1/1/4
MATGEN ,/PVR6/6/6/2/1/3
MERGE ,,,,I1X1,PVC6,PVR6/ONE6X6/1/0/1
MPYAD ONE6X6,I6X6,ONE6X6/ROTX/1/-1
DELETE /PVC6,PVR6,ONE6X6,,/
N1C = CMPLX(N1,0.)
ENDIF
MATGEN ,/PVC6/6/6/2/1/3
MATGEN ,/PVR6/6/6/0/1/5
MERGE ,,,,I1X1,PVC6,PVR6/ONE6X6/1/0/1
MPYAD ONE6X6,I6X6,ONE6X6/ROTY/1/-1
DELETE /PVC6,PVR6,ONE6X6,,/
N2C = CMPLX(N2,0.)
ENDIF
MATGEN ,/PVC6/6/6/0/1/5
MATGEN ,/PVR6/6/6/1/1/4
MERGE ,,,,I1X1,PVC6,PVR6/ONE6X6/1/0/1
MPYAD ONE6X6,I6X6,ONE6X6/ROTZ/1/-1
DELETE /PVC6,PVR6,ONE6X6,,/
N3C = CMPLX(N3,0.)
ENDIF
ADD5 ROTX,ROTY,ROTZ,,/ROTTOT/N1C/N2C/N3C
ADD ROTTOT,/OMEGA/OMEGAC
MATMOD OMEGA,SIL,,,,/OMEGAGG,/5
PARTN OMEGAGG,MJJC,/,,,OMEGAXX/
PARAML CSTM//'PRESENCE'////S,N,NOGTOB=0
IF (NOGTOB < 0) THEN
DIAGONAL MJJ/GTOBY/'SQUARE'/0.
ELSE
MATMOD CSTM,SIL,BGPDT,,,/GTOBY,/5//-1
ENDIF
PARTN GTOBY,MJJC,/,,,GTOB/
IF (CID = 0) THEN
EQUIVX GTOB/CTOG/ALWAYS
ELSE
MATMOD CSTM,SIL,BGPDT,,,/CTOBX,/5//CID
PARTN CTOBX,MJJC,/,,,CTOB/
MPYAD CTOB,GTOB,/CTOG/1
ENDIF
SMPYAD CTOG,OMEGAXX,CTOG,,,/OMEGAYY/3////1
MERGE ,,,,OMEGAYY,MJJC,/OMEGAGGX/
FORCE MATRIX
MPYADMJJ,OMEGAGGX,/CJJ1
MPYAD OMEGAGGX,MJJ,/CJJ2

150

ADD CJJ1,CJJ2/BCOR
CETRIPETAL FORCE MATRIX
SMPYAD OMEGAGGX,MJJ,OMEGAGGX,,,/KCEN/3////////6
RETURN
END
echoon

151

BIBLIOGRAPHY

[1] D. L. Anton. \Simultaneous Fretting and Fatigue of = Ti-6Al-4V". Proceedings (CD) of the 4th National Turbine Engine High Cycle Fatigue Conference,
1999.
[2] U. M. Ascher, R. M. M. Mattheij, and R. D. Russell. Numerical Solution of
Boundary Value Problems for Ordinary Di erential Equations. Prentice-Hall
Inc., 1988.
[3] J. M. Baik and R. B. Thompson. \Ultrasonic Scattering from Imperfect Interfaces: A Quasi-Static Model". Journal of Nondestructive Evaluation, 4:177{
195, 1984.
[4] B. M. Barthelemy, M. P. Kamat, and H. F. Brinson. Finite Element Analysis
of Bonded Joints. Oce of Naval Research No. N0014-82-K-0185, 1984.
[5] F. P. Beer and E. R. Johnston, Jr. Vector Mechanics for Engineers - Dynamics.
McGraw-Hill International Group, 1988.
[6] D. Bella, M. Reymond, and Editors. DMAP Module Dictionary. The MacnealSchwendler Corp., 1994.
[7] R. M. Bozorth. Ferromagnetism. D. Van Nostrand Co., 1951.
[8] D. Chen and S. Cheng. \An Analysis of Adhesive-Bonded Single-Lap Joints".
Journal of Applied Mechanics, 50:109{115, 1983.
[9] A. W. Cochardt. \The Origin of Damping in High-Strength Ferromagnetic
Alloys". Journal of Applied Mechanics, 20:196{200, 1953.
[10] K. R. Cross, R. L. Newman W. R. Lull, and J. R. Cavanagh. \Potential of
Graded Coatings in Vibration Damping". An Engineering Note, Journal of
Aircraft, 10:685{687, 1973.
[11] E. J. Doedel, A. R. Champneys, T. F. Fairgrieve, Y. A. Kuznetsov, B. Sandstede, and X. Wang. Continuation and Bifurcation Software for Ordinary Differential Equations. Canada, 1998. (AUTO97).
152

[12] ASTM E756-93.

Standard Test Method for Measuring Vibration-Damping


Properties of Material. American Society for Testing and Materials (ASTM),

[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]

1993.
M. Goland and E. Reissner. \The Stresses in Cemented Joints". Journal of
Applied Mechanics, 11:A17{A27, 1944.
R. W. Gordon and J. J. Hollkamp. \Internal Damping Treatment for Gas
Turbine Blades". AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference, VI:442{451, 1997.
J. H. Grin. \Friction Damping of Resonant Stresses in Gas Turbine Engine
Airfoils". International Journal of Turbo and Jet Engine, 7:297{307, 1990.
L. J. Hart-Smith. \Adhesive-Bonded Single-Lap Joints". NASA Report, CR112236, 1973.
D. N. Herting. MSC/NASTRAN Advanced Dynamic Analysis User's Guide.
The Macneal-Schwendler Corp., 1997.
Jay Ho man. \Magnetic Damping System to Limit Blade Tip Vibrations in
Turbomachines". U. S. Patent, 5,490,759, 1996.
R. Kielb and etal. \Advanced Damping Systems for Fan and Compressor
Blisks". Proceedings (CD) of the 4th National Turbine Engine High Cycle
Fatigue Conference, 1999.
T. P. Lang and P. K. Mallick. \The E ect of Recessing on the Stresses in
Adhesively Bonded Single-Lap Joints". International Journal of Adhesion and
Adhesives, 19(1):257{271, 1999.
A. I. Lavrentyev and S. I. Rokhlin. \Models for Ultrasonic Characterization
of Environmental Degradation of Interfaces in Adhesive Joints". Journal of
Applied Physics, 76(8):4643{4650, 1994.
Kaushik A. Lyer and Shanker Mall. \Fretting Fatigue of Ti-6Al-4V Under
Variable-Amplitude Loading". Proceedings (CD) of the 4th National Turbine
Engine High Cycle Fatigue Conference, 1999.
F. J. Margetan, R. B. Thompson, and T. A. Gray. \Interfacial Spring Model for
Ultrasonic Interactions with Imperfect Interfaces: Theory of Oblique Incidence
and Application to Di usion-Bonded Butt Joints". Journal of Nondestructive
Evaluation, 7:131{152, 1988.
153

[24] D.C. Maxwell and T. Nicholas. \A Rapid Method for Generation of a Haigh
Diagram for High Cycle Fatigue". Fatigue and Fracture Mechanics, 29:ASTM
STP{1321, 1998.
[25] David M. North. Aerospace Daily. McGraw-Hill Co., 1999.
[26] D. W. Oplinger. \E ects of Adherend De ections in Single-Lap Joints". International Journal of Solid Structures, 31(18):2565{2587, 1991.
[27] D. W. Oplinger. TJOINTNL: For Stresses in Bonded Joints with Tapered
Adherends and Nonlinear Adhesive Response, 1995.
[28] D. Ross, E. E. Ungar, and E. M. Kerwin, Jr. \Damping of Plate Flexural
Vibrations by Means of Viscoelastic Laminate". Structural Damping, ASME,
New York, pages 49{88, 1959.
[29] M. H. Herman Shen. \Reliability Assessment of High Cycle Fatigue Design of
Gas Turbine Blades Using the Probabilistic Goodman Diagram". International
Journal of Fatigue, 21:699{708, 1999.
[30] G. W. Smith and J. R. Birchak. \E ect of Internal Stress Distribution on
Magnetomechanical Damping". Journal of Applied Physics, 39(5):2311{2316,
1968.
[31] G. W. Smith and J. R. Birchak. \Internal Stress Distribution Theory of Magnetomechanical Hysteresis-An Extension to Include E ects of Magnetic Field
and Applied Stress". Journal of Applied Physics, 40(13):5174{5178, 1969.
[32] H. Tada, P. Paris, and G. Irwin. The Stress Analysis of Cracks Handbook. Del
Research Corporation, St. Louis, MO., 1973.
[33] L. Tong, A. Sheppard, and D. Kelly. \Relationship Between Surface Displacement and Adhesive Peel Stress in Bonded Double Lap Joints". International
Journal of Adhesion and Adhesives, 5(1):43{48, 1995.
[34] K. Washizu. Variational Methods in Elasticity and Plasticity. Pergamon Press,
1982.

154

You might also like