Sei sulla pagina 1di 430

Materials Research and Engineering

Edited by B. llschner and N. 1. Grant

Hermann Riedel

Fracture
at High Temperatures
With 109 Figures

Springer-Verlag
Berlin Heidelberg GmbH

Dr. HERMANN RIEDEL


Max-Planck-Institut fiir Eisenforschung
4000 Dusseldorf, FR Germany
New Address, Starting October 1986:
Fraunhofer-Institut fUr Werkstoffmechanik
7800 Freiburg, FR Germany

Dr. rer. nat. BERNHARD ILSCHNER


o. Professor, Laboratoire de Metallurgie Mecanique,
Departement des Materiaux, EPFL, Lausanne

Prof. NICHOLAS J. GRANT


Department of Materials Science and Engineering, Cambridge

ISBN 978-3-642-82963-5
ISBN 978-3-642-82961-1 (eBook)
DOI 10.1007/978-3-642-82961-1
Library of Congress Cataloging in Publication Data.
Riedel, Hermann
Fracture at high temperatures.
(Materials research and engineering)
Bibliography: p.
Includes index.
1. Fracture mechanics. 2. Materials at high temperatures. 3. Materials--Creep.
I. Title. D. Series.
TA409.R54 1987 620.1'126 86-31444

This work is subject to copyright. All rights are reserved, whether the whole or part of the
material is concerned, specifically those of translation, reprinting, re-use of illustrations,
broadcasting, reproduction by photocopying machine or similar means, and storage in data
banks. Under 54 ofthe German Copyright Law where copies are madeforotherthan private
use, a fee is payable to "Verwertl.!,ngsgesellschaft Wort", Munich.
Springer-Verlag Berlin Heidelberg 1987
Softcover reprint of the hardcover 1st edition 1987
The use of registered names, trademarks, etc. in this publication does not imply, even in the
absence of a specific statement, that such names are exempt from the relevant protective laws
and regulations and therefore free for general use.

2161/3020-543210

Editors' Preface

This book covers the fundamentals of fracture at high temperature. It


has been written by an author whose outstanding competence in this field has
gained worldwide appreciation. He has contributed substantially to the understanding of a phenomenon that demands a scientific treatment which judiciously integrates the concepts of mechanical deformation, thermodynamic
eqiulibria, and kinetic rate laws. This challenge has been taken up by Hermann
Riedel with both skill and energy.
For the first time, a comprehensive and coherent interpretation of all
the phenomena related to creep rupture is presented. A thorough analysis of
the relevant observations and a "close-packed" presentation of the associated
phys i ca I concepts introduce the reader to each one of the prob 1em areas
addressed in this book. They provide the material and the tools to construct
models for nucleation and growth of intergranular cavities, for the interaction of particles with sliding ~Boundaries, for stress concentration and
stress relief in triple junctions, for rupture criteria in creep and fatigue,
etc. The quantitative aspect is a permanent guideline, and so is the consideration of qualitative changes due to varying stress, temerature, or grain
size. The numerous references given constitute a complete review of the
re 1evant 1i terature, and a wea Ith of data, sk ill fu Ily extracted from recent
publications, renders the text practical for application in realistic case
studies. The obvious technical relevance of the subject matter is underlined
in the author's preface and becomes visible throughout the book.
Physical reasoning is backed up by mathematical calculus reflecting
the modern approach to process modelling and at the same time the author's
individuality. It is analytical wherever possible, translating reality into
formulas in an easy and elegant way, always keeping the balance between
necessary rigidity and acceptable simplification.
The editors and the publishers are proud to add, with this book,
another important volume to the series "Materials Research and Engineering".
They express the i r confi dence that it wi 11 recei ve the resonance wh i ch it
merits within the community devoted to the science of advanced engineering
materials.

B.

Iischner

(N.J. Grant) - by permission

Author's Preface

Within the framework of fracture research, this book covers the area
temperature

fracture

with

of

high-

a particular emphasis on theoretical modeling. The

book is intended to evaluate the state of the art critically, primarily from
scientific

point

of

view.

Of course, scientific work in a subject area like

high-temperature fracture must be related to practical questions. For


it

should

example,

contribute to the development of better design codes and inspection

procedures in
plants

or

high-temperature

applications

such

as

electricity-generating

gas turbines. To some extent, the book also aims at the improvement

of high-temperature alloys by identifying the

microstructural

features

which

influence the fracture properties beneficially or adversely. Thus, the book has
been written mainly for researchers in material science, metallurgy and
nical

engineering,

but

mecha-

also for design engineers and persons responsible for

the safe operation of high-temperature equipment. In an attempt to give a


plete

com-

and coherent picture of what is known about high-temperature fracture, I

may have sacrificed the degree of simplicity which would be required in


graduate

courses.

On

the

under-

other hand, postgraduate readers will prefer to be

taken in large steps from the underlying assumptions of a model

to

the

final

results, as I have chosen to do.


The book is organized in three parts and 28 chapters. Part I summarizes the deformation and fracture behavior of materials in a qualitative manner and introduces the equations of solid mechanics and of
most

difficult

growth

and

nucleation

part

to

coalescence
process

write
of

was

grain

withstood

stress-directed

diffusion.

boundary

cavities.

In

particular,

the

all attempts to link theories quantitatively to

the observed behavior. I have tried to fill gaps in the existing knowledge
to

utilize

information

from

intergranular fracture, stress


embrittlement.

neighboring
relief

subject

cracking,

areas

hydrogen

such
attack

as
or

and

brittle
helium

Several sections owe their existence to these efforts. Finally,

it turned out to be necessary to retreat to an


nucleation

The

Part II, which describes the nucleation,

stage.

Fortunately,

cavity

growth

empirical

description

of

the

was a less difficult subject.

VII

Among the many models proposed in the literature, some could be


applicable

identified

as

to a given material under given testing conditions. Measured cavity

growth rates, for example in commercial

materials,

could

then

be

explained

quantitatively.
The following example represents a possible, though somewhat ambitious application

of

cavitation

models.

Many

parts of coal-fired electricity-generating

plants operate at high temperatures at which failure by cavitation is a


tial

poten-

danger. To avoid catastrophic failures, as have occurred occasionally, it

is increasingly becoming industrial practice to take replicas from

etched

and

polished surfaces of critical parts during the regular inspections. If cavities


have developed they can be detected by a microscopic examination of the
cas.

At

present,

however,

there

is

no

common

agreement

as

repli-

to what the

occurrence of cavities means for the remaining life of the part. With a
understanding

of

cavity

better

nucleation and growth one should be able to evaluate

the degree of damage more quantitatively and to make more accurate estimates of
the remaining lifetime.
In Part III, theories and
fatigue

crack

growth

experimen~s

relating to creep crack growth and creep-

are comprehensively described. Cracks, as distinct from

cavities, are understood to be larger than the typical microstructural

lengths

of the material such as the grain size. Cracks may be present in an engineering
structure from the beginning in the form of fabrication defects,

or

they

may

develop during service as a consequence of fatigue loading, thermoshock, corrosion or some other problem. Depending on the circumstances, a crack may
the

load-carrying

capacity or the lifetime of a structure conSiderably, or it

may be relatively harmless. Under ambient-temperature conditions, the


having

cracks

in

risk

of

a structure can be successfully assessed using the fracture

mechanics approach. For relatively


factor

reduce

brittle

materials,

the

stress

intensity

characterizes the behavior of cracks within the limits laid down in the

ASTM-E 399 rule of the American Society for Testing and Materials. Under a wide
range

of

conditions, the stress intensity factor also describes fatigue crack

growth and stress corrosion cracking. The use of


testing

of

ductile

the

J-integral

in

fracture

materials is regulated by the ASTM-E 813 rule. So far, no

comparable standards exist for cracks in high-temperature components. In,


the

Boiler

and

Pressure

Vessel

Code

Engineers does not permit any crack-like defects in ferritic


under

fact,

of the American Society of Mechanical


steels

operating

creep conditions. This is unnecessarily restrictive, so that a realistic

basis for making decisions about repair, replacement or continued operation

of

VIII

a cracked part is lacking.


The aim of Part III is therefore to extend the applicability of
anics

methods

to

fracture-mech-

high temperatures. The complicating factor compared to room

temperature is the time dependence of the material response. A significant contribution

to the development of time-dependent fracture mechanics must be made

by theory. Theoretical reasoning plays a more central role


modeling

of

cavitation.

here

than

in

the

Without a careful solid-mechanics analysis one could

hardly identify the load parameters which allow a

macroscopic

description

of

creep crack growth. The presentation in Part III is grouped around" the C*-integral. Its meaning and its practical use are described and its
discussed

thoroughly.

Its

range

limitations

are

of validity, as well as those of other load

parameters, are displayed on load parameter maps.


Finally, I wish to express my gratitude to all my colleagues who supported this
work by discussions, typing the text, drawing the figures, or reading the chapters. The most important point, however, was that I had the freedom to invest a
great

fraction

of

my

working time at Max-Planck-Institut into this project.

Otherwise it would have been impossible to write

the

book

in

(nearly) satisfies the writer and, I hope, pleases the reader.

Dusseldorf, September 1986


Hermann Riedel

form

which

Contents

PART I. INTRODUCTORY CHAPTERS


ON DEFORMATION AND FAILURE UNDER CREEP CONDITIONS
SUMMARY OF THE DEFORMATION BEHAVIOR UNDER CREEP CONDITIONS
1 .1
1.2
1.3
1.4
1.5

1.6
2

The Creep Curve


A Few Facts on the Micromechanisms Underlying the Creep Curve
Diffusion Creep
Inhibition of Diffusion Creep
Grain Boundary Sliding
1.5.1 The infinite grain boundary (an intrinsic sliding model)
1.5.2 Grain boundary sliding in polycrystals (extrinsic models)
Deformation-Mechanism Maps

3
4
6
7
8
8
10
11

INTRODUCTION TO CREEP FRACTURE AND OTHER FRACTURE MODES

14

2.1
2.2

14
15
16
18
18
21
21
22
23
23
23
24
24
26

2.3

The Nature of Creep Damage


Fracture-Mechanism Maps
2.2.1 Cleavage and brittle intergranular fracture
2.2.2 Ductile trans granular fracture by plastic hole growth
2.2.3 Necking and superplasticity
2.2.4 Intergranular creep fracture
2.2.5 Rupture by dynamic recrystallization
2.2.6 Fracture at very high temperature
Empirical Formulas for the Rupture Time in the Creep Regime
2.3.1 The Monkman-Grant rule
2.3.2 The Sherby-Dorn parameter
2.3.3 The Larson-Miller parameter
2.3.4 The Kachanov equations
2.3.5 The a-projection concept

3 THE CONTINUUM-MECHANICAL EQUATIONS


3.1
3.2
3.3
3.4

3.5
4

The Equations for Equilibrium and Compatibility


The Material Law
The Equations for Antiplane Shear, Plane Stress and Plane Strain
General Features of the Continuum-Mechanical Fields
3.4.1 The elastic-viscous analogy (Hoff, 1954)
3.4.2 Scaling properties for power-law materials (Ilyushin, 1946)
3.4.3 Path-independent integrals: J and C*
3.4.4 The HRR crack-tip fields in power-law materials
Numerical Techniques in Solid Mechanics

27
27
28
29
33
33
33
34
36
39

STRESS-DIRECTED DIFFUSION AND SURFACE DIFFUSION

40

4.1
4.2
4.3
4.4
4.5

40
41
43
45
46

The Role of Vacancy Sources in Stress-Directed Diffusion


Stress-Directed Diffusion Along Grain Boundaries
Stress-Directed Diffusion Through the Grains
Surface Diffusion
Grain-Boundary Diffusion Combined with Power-Law Creep

x
Part II. CREEP CAVITIES
5

INTRODUCTION TO PART II

51

5.1
5.2
5.3
5.4

52
53
55
56
56
56
57
57
57
59
59
61
62
62
65
66

5.5

5.6
5.7
5.8
6

NUCLEATION OF CREEP CAVITIES I BASIC THEORIES

67

6.1
6.2

67
69
69
70
72
75
77
78
80

6.3

Experimental Techniques
Materials which Exhibit Intergranular Cavitation
Diffusion as the General Cause for Intergranular Cavitation
The Role of Grain Boundary Sliding
5.4.1 Experiments on bicrystals
5.4.2 The orientation of cavitating boundaries in poly crystals
Cavity Nucleation Sites
5.5.1 Slip bands
5.5.2 Grain-boundary ledges
5.5.3 Triple grain junctions
5.5.4 Grain boundary particles
Wedge Cra.cks
Some Observations on the Kinetics of Cavity Nucleation
5.7.1 The observed nucleation kinetics
5.7.2 Is there a critical stress for cavity nucleation?
Pre-Existing Cavities

Cavity Nucleation by the Rupturing of Atomic Bonds


Cavity Nucleation by Vacancy Condensation
6.2.1 Historical remarks and related subject areas
6.2.2 Cavity shapes
6.2.3 The free energy of a cavity
6.2.4 The nucleation rate according to Raj and Ashby
6.2.5 The Fokker-Planck equation
6.2.6 The steady-state nucleation rate and the nucleation stress
6.2.7 Transient solutions of the Fokker-Planck equation
and incubation times
Discussion of Cavity Nucleation Theories
6.3.1 A theoretical remark
6.3.2 On possible causes for the discrepancy between theoretical
and experimental nucleation stresses
6.3.3 The problem of continuous cavity nucleation

82
83
83
84

CAVITY NUCLEATION BY STRESS CONCENTRATIONS DURING CREEP

85

7.1

86
86
89
90

7.2

An Isolated Sliding Grain Boundary Facet (Shear-Crack Model)


7.1.1 ElastiC analysis of a sliding facet
7.1.2 A sliding boundary facet (shear crack) in creeping material
7.1.3 Relaxation of elastic stress concentrations at a shear
crack by power-law creep
7.1.4 The time to build up elastic stress concentrations
The Triple Grain Junction in Polycrystals
7.2.1 The triple junction in elastic material
7.2.2 The triple junction in power-law creeping material
7.2.3 Stresses during Coble creep (rigid grains)
7.2.4 A cQmbination of power-law creep and grain-boundary diffusion
7.2.5 Relaxation of elastic stress concentrations at triple
junctions by creep
7.2.6 Relaxation of elastic stress concentrations at triple

91
92
93
94
96
98
99
99

XI

7.3

7.4
7.5
8

Concentrations at Particles on Sliding Grain Boundaries


Elastic stress concentrations at two-dimensional particles
Elastic stress concentrations at three-dimensional particles
Stresses at two-dimensional particles during power-law creep
Stresses at three-dimensional particles during power-law
creep
7.3.5 Diffusion and creep around particles during power-law
creep of the grains
7.3.6 Stresses at particles during (free and inhibited) Coble
creep
7.3.7 Relaxation of elastic stress concentrations at particles
by creep
7.3.8 Relaxation of elastic stress concentrations at particles
by diffusion
Stresses at Grain-Boundary ~edges
Summary of Stress Concentrations

THE ROLE OF IMPURITY SEGREGATION IN CAVITY NUCLEATION


8.1

8.2

Stress
7.3.1
7.3.2
7.3.3
7.3.4

11

108
110
112
113
114
115
i

16

Qualitative Observations
116
8.1.1 Grain-boundary brittleness at room temperature
116
(temper embrittlement)
8.1.2 Embrittlement by impurity segregation under creep conditions 117
8.1.3 Stress relief cracking or reheat cracking
119
Theories Related to Segregation and Cohesion
121
8.2.1 Segregation equilibria
121
8.2.2 Segregation kinetics
123
8.2.3 Calculation of interface energies from adsorption data
124
8.2.4 The relevance of segregation for decohesion
127
8.2.5 The effect of segregation on cavity nucleation by
129
vacancy condensation

CAVITY NUCLEATION ASSISTED BY INTERNAL GAS PRESSURE

131

9.1

131
132
135
136
138
139

9.2
9.3
9.4
10

102
102
103
105
107

Oxygen Attack and Related Phenomena


9.1.1 The equilibrium carbon-dioxide pressure in nickel
9.1.2 Carbon~oxides in nickel-chromium alloys
Hydrogen Attack
Helium Embrittlement
Kinetic Aspects

INTERNAL STRESSES DUE TO THE PRECIPITATION OF SOLID PHASES


AND THERMAL EXPANSION

140

10.1
10.2
10.3
10.4
10.5
10.6

140
142
144
145
146
147

The Flux of Carbon to the Carbide


Elastic Accommodation
Accommodation by Power-Law Creep
Accommodation by Grain Boundary Diffusion
Decohesion of Particles by Thermal Expansion
Grain-Boundary Decohesion by Thermal-Expansion Anisotropy

DIFFUSIVE CAVITY GROWTH

148

11.1

149
150

Diffusional Growth of Lens-Shaped (Equilibrium) Cavities


11.1.1 The stress distribution between the cavities
the cavity growth rate

XII

11.1.2

11.2

12

Rupture times by diffusive cavity growth neglecting


nucleation
11.1.3 The effect of the sintering stress on the rupture time
11.1.4 Removal of cavities by compressive loads or by surface
tension forces
11.1.5 The effect of impurity segregation on diffusive
cavity growth
11.1.6 The effect of gas pressure on the diffusive cavity
growth rate
Diffusional Growth of Non-Equilibrium Cavities
11.2.1 The procedure to solve the coupled problem of surface
diffusion and grain boundary diffusion
11.2.2 Re-formulation of the surface diffusion problem
11.2.3 A steady-state solution of the surface diffusion problem
in the crack-like limit
11 .2.4 Similarity solutions for the surface diffusion problem
11.2.5 The relation between growth rate and stress in the
crack-like limit
11 .2.6 Rupture times for non-equilibrium growth
11.2.7 Experiments on copper and silver containing water
vapor bubbles
11.2.8 Void-shape instability/finger-like cavity growth

154
155
156
158
159
160
161
161
163
164
165
167
169
170

CONSTRAINED DIFFUSIVE CAVITATION OF GRAIN BOUNDARIES

172

12.1

173

12.2
12.3
12.4

12.5

12.6

12.7

12.8

Cavity Growth Rates for Constrained Cavitation of an


Isolated Facet
12.1.1 A tensile-crac~< model for the calculation of
constrained growth rates
12.1.2 Comparison with measured cavity growth rates
12.1.3 Additional remarks on constrained cavity growth rates
The Time to Cavity Coalescence on an Isolated Boundary Facet
On the Irrelevance of Constrained Cavity Growth for Rupture
Lifetimes
Comparison of Calculated Times to Cavity Coalescence on Isolated
Facets with Measured Rupture Lifetimes of Pre-Cavitated Materials
12.4.1 Rupture lifetime of prestrained Nimonic 80A
12.4.2 Rupture lifetime of prestrained Inconel alloy X-750
12.4.3 Rupture time of a-brass with implanted water vapor
bubbles
Constitutive Behavior of Creeping Materials Containing Widely
Spaced Cavitating Grain Boundary Facets
12.5.1 The constrained limit (Hutchinson's model)
12.5.2 The unconstrained limit
12.5.3 The effect of cavitation on diffusion creep
Interaction Between Closely Spaced Cavitating Boundary Facets
12.6.1 Self-consistent models for constrained cavitation
12.6.2 The penny-shaped crack in a finite cylinder
12.6.3 Interactions between closely spaced facets in the
presence of grain boundary sliding
Time to Rupture for Interacting Facets
12.7.1 Failure by large strains
12.7.2 Rupture lifetimes for continuous nucleation of
cavitating facets
12.7.3 The combined effect of necking and continuous nucleation
Conclusions on Constrained Cavitation

173
175
179
181
181
182
182
183
184
185
185
186
187
188
188
190
191
193
193
194
196
197

XIII

13

INHIBITED CAVITY GROWTH

13.1
13.2
14

14.2

201
202
204
206
209
210
212
215

15.1

215
215
218
220
220

Cavity Growth by a Coupling of Diffusion and Power-Law Creep


15.1.1 Models for the interactive growth mechanism
15.1.2 Comparison with experiments
Diffusive Cavity Growth with Elastic Accommodation
15.2.1 Elasticity effects in the growth of equilibrium-shaped
cavities
15.2.2 Crack-like cavity growth with elastic accommodation

221

THE CAVITY SIZE DISTRIBUTION FUNCTION FOR CONTINUOUS CAVITY


NUCLEATION. RUPTURE LIFETIMES AND DENSITY CHANGES

225

16.1
16.2

225

16.3

16.4
17

Hole Growth by Creep Flow of the Grains


14.1.1 The growth of isolated holes in linearly viscous
materials
14.1.2 An isolated circular-cylindrical void in nonlinear
viscous material
14.1.3 Spherical voids in nonlinear material under axisymmetric loading. Comparison with penny-shaped cracks
14.1.4 Strain to failure neglecting void interaction effects
14.1.5 Void interaction effects
Cavity Growth by Grain Boundary Sliding

CREEP-ENHANCED DIFFUSIVE CAVITY GROWTH AND ELASTIC ACCOMMODATION

15.2

16

198
200

CAVITY GROWTH BY CREEP FLOW OF THE GRAINS OR BY GRAIN BOUNDARY SLIDING 201
14.1

15

Inhibited Cavity Growth Rates


Time to Cavity Coalescence and Time to Rupture for
Inhibited Growth

198

The Cavity Size Distribution Function


The Cavitated Area Fraction and the Rupture Lifetime
16.2.1 Lifetimes for diffusive cavity growth and continuous
nucleation
16.2.2 Crack-like diffusive growth and continuous nucleation
16.2.3 Constrained diffusive growth and continuous nucleation
16.2.4 Inhibited cavity growth and continuous nucleation
16.2.5 Plastic hole growth and continuous nucleation
Comparison of Calculated Rupture Times with Experiments
Involving Continuous Nucleation
16.3.1 Rupture lifetimes of ferritic steels
16.3.2 Lifetimes of austenitic steels
16.3.3 Rupture lifetimes of astroloy
Density Changes During Cavitation

227

228
231
231
233
233
234
234
237
239
240

SUMMARY OF RESULTS ON CAVITY NUCLEATION AND GROWTH

242

17.1
17.2

242
243

17.3

Nucleation
Cavity Growth Rates and Rupture Lifetimes for Instantaneous
Nucleation
Rupture Lifetimes for Continuous Nucleation

246

XIV

18

GRAIN BOUNDARY CAVITATION UNDER CREEP-FATIGUE CONDITIONS

247

18.1
18.2

2"47
248
249

Micromechanisms of Creep-Fatigue Failure


Theories of Cavitational Failure for Slow-Fast Fatigue Loading
18.2.1 Cycles to failure for unconstrained diffusive cavity
growth
18.2.2 Cycles to failure for plastic hole growth
18.2.3 Cycles to failure for unconstrained growth
18.2.4 Summary of fatigue lifetimes for different cavity growth
mechanisms
18.3 Comparison with Results of Slow-Fast Tests
18.3.1' Low-cycle fatigue tests on Al-5:Mg
18.3.2 Low-cycle fatigue tests on nickel
18.3.3 Low-cycle fatigue tests on copper
18.3.4 Low-cycle fatigue tests on austenitic steel
18.4 Why Do Cavities Grow under Balanced Cyclic Loading?
18.5 Discussion

251
252
253
254
254
255
256
257
258
259

PART III. CREEP CRACK GROWTH AND CREEP-FATIGUE CRACK GROWTH


19

20

21

22

INTRODUCTION TO PART III

263

19.1
19.2
19.3

263
264
265
265
266

The Relevance of Cracks


The First Aspect: Deformation Fields in Cracked Bodies
The Second Aspect: Mic~omechanisms
19.3.1 Grain boundary cavitation ahead of the crack tip
19.3.2 Corrosive processes at the crack tip

NONLINEAR VISCOUS MATERIALS AND THE USE OF C*

267

20.1
20.2

267
268
269
271

Definition of the C*-Integral


Stress Fields and the C*-Integral in Power-Law Viscous Materials
20.2.1 The C*-integral in power-law viscous materials
20.2.2 Crack-tip fields in power-law viscous materials

C*-CONTROLLED CREEP CRACK GROWTH BY GRAIN-BOUNDARY CAVITATION

272

21.1
21.2
21.3
21.4
21.5

273
277
279
280
281
281
283
285

Creep Crack Growth Based on a Local Critical-Strain Criterion


Strain-Controlled Cavity Growth and Stress-Controlled Nucleation
Diffusive Growth of a Constant Number of Cavities
Diffusive Cavity Growth and Stress-Controlled Nucleation
Comparison with Experiments
21.5.1 Tests on a lCr-1/2Mo steel
21.5.2 Comparison of the data with models
21.5.3 Conclusions

SPECIMEN SIZEREQUlREHENTS FOR C*-TESTING CAUSED BY CRACK-TIP BLUNTING 286


AND BY 3-D EFFECTS
22.1
22.2

Limitations to C* Set by Blunting


The Third Dimension in Fracture Mechanics and its Practical
Consequences
22.2.1 The C*-integral in three dimensions

286
288
289

xv
22.2.2
22.2.3
22.2.4
22.2.5
22.2.6
22.2.7
22.2.8
22.2.9
23

292
293
295
296
297
298
301

23.1

301
302

23.3

23.4

25

290
290

ELASTIC/NONLINEAR VISCOUS MATERIALS. APPLICABILITY OF KI AND OF C*

23.2

24

Crack-tip fields in specimens of finite thickness


The singularity at the intersection of the crack front
with the surface
Ranges of validity of singular fields in parallel-sided
specimens with straight crack fronts
Conditions for plane strain near the crack tip
Thumbnail-shaped crack fronts
Shear lips
Crack-tip fields in side-grooved specimens
The compliance and C* in parallel-sided and side-grooved
specimens

Stationary Crack under Step Loading


23.1.1 Similarity solutions in the small-scale creep, or
short-time, limit
23.1.2 The crack-tip field in the short-time limit
23.1.3 The complete stress field in the short-time limit
23.1.4 The creep zone
23.1.5 A characterisitc transition time
23.1.6 Interpolation formulas for the transient regime
23.1.7 Possible generalizations and related work
Stress Fields at Growing Cracks in Elastic/Nonlinear Viscous
Material
23.2.1 Derivation of the singularity at growing cracks for
Mode III
23.2.2 The growing crack~ingularity: results for Mode I
23.2.3 Fields for steady-state crack growth under small-scale
creep conditions
23.2.4 Steady-state crack growth during extensive creep of the
whole specimen
23.2.5 The evolution of the asymptotic field under non-steadystate conditions
Crack Growth in Elastic/Nonlinear Viscous Material Subject to
a Critical-Strain Criterion
23.3.1
rHR < Xc and a-a o < rcr
23.3.2 ~~:~~S~O~~ht~~b~:~~ to a critical-strain criterion for
small-scale creep
Application to Experiments
23.4.1 The appropriate load parameter
23.4.2 A 1Cr-1/2Mo steel
23.4.3 Nimdnic 80A

304
305
306
308
309
311
312
312
314
315
316
317
319
319
321
324
324
324
325

INSTANTANEOUS PLASTICITY

327

24.1
24.2
24.3
24.4

328
329
330
331

Deformation Fields in Elastic/Plastic Material


Growth of a Creep Zone in an Initially Fully-Plastic Body
The Special Case N = 1/n
An Experimental Example for J-Controlled Creep Crack Growth

PRIMARY-CREEP EFFECTS
25.1

Strain-Hardening Model for Primary Creep


25.1.1 Primary creep of the whole specimen
25.1.2 Growth of a primary-creep zone in an elastic field

332
332
333
334

XVI

25.1.3
25.2

25.3
26

336
338
338
340
341
342
346

26.1
26.2

346
346

Constitutive Law
The Effect of Diffusion Creep on the Deformation
Fields in Cracked Bodies
Crack Growth Rates Assuming a Critical-Strain Criterion

A DAMAGE MECHANICS APPROACH TO CREEP CRACK GROWTH


21.1

21.2

21.3

21.4
21.5

21.6
21.1
28

335

DIFFUSION CREEP

26.3
21

Growth of a secondary-creep zone in a primary-creep


field
25.1.4 Summary and introduction of a load parameter map
Hardening/Recovery Model for Primary Creep
25.2.1 The constitutive equations
25.2.2 Solutions for crack geometries
25.2.3 Elasticity effects and load parameter map
Analysis of an Experiment in the Transition Range Between
J, CI\. and C*

Introduction
21.1.1 The constitutive model
21.1.2 The relation between fracture mechanics and damage
mechanics
Small-Scale Damage in Extensively Creeping Specimens
21.2.1 Similarity solutions
21.2.2 Crack growth rates
21.2.3 Approximate and numerical methods in small-scale damage
21.2.4 The process zoq~
The Range of Validity of the Small-Scale Damage Approximation
in Extensively Creeping Specimens
The Evolution of Damage and Crack Growth for Small-Scale Creep
21.4.1 Crack grows faster than creep zone
21.4.2 Creep zone grows faster than process zone
Primary-Creep Effects
21.5.1 Small-scale damage in a specimen which creeps in the
primary stage
21.5.2 The transient from elasticity over primary to secondary
creep
The Evolution of the Crack Length and the Lifetime
Discussion

348
349
349
349
350
352
352
352
353
354
355
356
351
358
359
359
359
359
362

CREEP-FATIGUE CRACK GROWTH

364

28.1

365
365

28.2
28.3

28.4

Micromechanisms of Fatigue Crack Growth


28.1.1 The alternating slip model (also called the crack-tip
blunting model
28.1.2 Fatigue crack growth by grain boundary cavitation
28.1.3 Corrosive effects in creep-fatigue crack growth
Fatigue Cracks in Viscous Materials
28.2.1 Growth rates by the alternating slip mechanism
28.2.2 Growth by cavitation in viscous materials
Fatigue Cracks in Elastic-Plastic Materials
28.3.1 Elastic-plastic deformation fields
28.3.2
The cyclic J-integral, Z
28.3.3 Z-controlled crack growth rates by alternating slip
Fatigue Cracks in Elastic/Nonlinear Viscous Materials
28.4.1 Stress fields in elastic/nonline~r viscous material
after a load step

361
368
310
310
310
311
312
312
312
313
313

XVII

28.4.2

28.5

28.6
28.7

Gradual load variations in elastic/nonlinear


viscous material
28.4.3 Stress fields for rapid cyclic loading
28.4.4 Crack growth rates by the alternating slip mechanism
28.4.5 Fatigue crack growth by cavitation ahead of the crack
The Combined Effects of Elastic, Plastic and Creep Deformation
on Fatigue Crack Growth Rates
28.5.1 An approximate general expression for the crack growth
rate by alternating slip
28.5.2 Creep-fatigue crack growth rates in fracture mechanics
specimens
28.5.3 Fatigue lifetimes of initially smooth specimens by
microcrack growth
Discussion
Summary

374
376
377
378
379
379
380
381
384
385

APPENDICES
APPENDIX A: MATERIAL PARAMETERS

389

APPENDIX B: ELASTIC STRESS FIELDS AT NOTCHES, CRACKS AND GRAIN


BOUNDARY TRIPLE POINTS

391

B.l
B.2

Stress Fields at Sharp Notches and Cracks


B.1.1 The eigenvalue eqUation for sharp notches
B.l.2 Crack-tip fields
The Stress Singularity at a Triple Junction of Sliding
Grain Boundaries

392
392
393
395

APPENDIX C: CALCULATION OF C* FOR TEST SPECIMEN CONFIGURATIONS

396

REFERENCES

401

INDEX

417

Part I

Introductory Chapters
on Deformation and Fail
Under Creep Conditions

1 Summary of the Deformation Behavior Under


Creep Conditions

Deformation and fracture of materials under elevated-temperature creep


ions

are

time-dependent

condit-

processes. At temperatures below some 30 per cent of

the absolute melting temperature it is a reasonable and widely used

idealizat-

ion to consider the elastic-plastic behavior of metals as time-independent. The


strain developed instantaneously in response to a load is large compared to the
additional

strain which is accumulated within any practically interesting hold

time. Many technical applications, however, require temperatures far beyond the
time-independent

regime,

which

ends

at

some 400 0 C for ferritic steels, for

example. Then the continuing plastic deformation (creep) under sustained


which

load,

eventually leads to creep fracture, often becomes the determining factor

for the design of a structure.


The subject of the present

monograph

is

creep

fracture

rather

than

creep

deformation. However, the fracture mechanisms are usually intimately related to


the preceding deformation processes. Hence a brief account of
behavior

of

the

deformation

materials at elevated temperatures must be given. There are other

more comprehensive treatises of creep deformation available, for example


by

Ilschner

those

(1973), Gittus (1975), Langdon (1981), Lagneborg (1981) and Frost

and Ashby (1982).

1.1 The Creep Curve


The elementary test to study creep deformation
creep

test.
t.

failure

is

the

uniaxial

A smooth tensile bar is subjected to a time-independent load (or

stress a), and the elongation (or strain


time

and

Figure

1.1

schematically

E)

shows

is measured as a

function

of

a creep curve as it is obtained at

around half the melting temperature and at stress levels which are typical
creep

tests

the
for

in the laboratory. Immediately upon load application, there is an

elastic (plus, possibly, an instantaneous plastic) strain. Then, in the primary

1.

Creep Deformation

t
tertiary
secondary

time - -

Fig. 1.1. The creep curve (schematic).

stage of the creep curve, the creep strain rate (that is the slope of the creep
curve)

is

initially

large

secondary stage, which is


tertiary

stage,

the

and

decreases

alternatively

creep

until

called

it becomes constant in the

steady-state

creep.

In

the

rate accelerates until final fracture occurs. The

shape of the creep curve varies from material to material.

Pure

metals

often

have a pronounced primary stage, whereas in many structural alloys the tertiary
stage predominates. Several phenomenological descriptions of

the

creep

curve

will be introduced later in this book.

1.2 A Few Facts on the Micromechanisms


Underlying the Creep Curve
The primary and secondary stages of the
combined

action

of

strain

dislocation structure (Bailey, 1926,


strain

hardening

by

the

creep

curve

formation

Orowan,
of

1946).

balance

between

~train

requires

determined

by

the

In

the

primary

stage,

dislocation tangles or a dislocation

substructure predominates, whilst the secondary stage


dislocation creep is the

are

hardening and thermally activated recovery of the

is

characterized

by

hardening

and recovery. The rate-controlling step in

climb

edge

of

dislocations

over

obstacles.

This

the diffusion of vacancies and therefore takes place only at elevated

temperatures at a practically interesting rate. Having surmounted an obstacle a


dislocation can glide freely until it is impeded by the next obstacle.
Specific models of steady-state, secondary creep based on the

idea

of

climb-

1.2

Micromechanisms Underlying the Creep Curve

controlled dislocation motion have been proposed by Weertman (1955) and Friedel
(1967). Their common result is a power-law dependence of the strain rate

on

the stress, which is often called Norton's (1929) creep law:

B an ,

( 1.1)

where Band n are material parameters. For the stress exponent


predict

values

of

the

models

(Friedel, 1967) or 4 (Weertman, 1955). If the vacancies

diffuse predominantly along dislocation


through

n,

lines

(pipe

diffusion)

rather

the undisturbed lattice, the value of n is raised by 2, i.e., n

6 is predicted (Frost and Ashby, 1977). Further, the models for

than
5 or

diffusion-con-

trolled creep predict that the coefficient B should scale in the following way:
B ~ A*G 1- n (bD IkT),

(1.2)

where G is
constant,

the

elastic

shear

modulus,

k ~ 1.38.10- 23 J/K

is

Boltzmann's

T is absolute temperature, b is the magnitude of the Burgers vector,

A* is an empirical constant which accounts for details not properly included in


the models and

(1.3)

is

the

bulk

diffusion

coefficient

(or

the

pipe

coefficient,

diffusion

respectively) with the pre-exponential factor Dvo and the activation energy Qv;
R ~ 8.315 J/(molK) iA the gas constant. Numerical values for the parameters
have been reported by Frost and Ashby (1977, 1982) for a number of materials. A
selection of material parameters is also compiled in Table A.1 in Appendix A.
Several other relationships between strain rate and stress have
in

the

literature.
by

stress

in

eq.

(1.1)

is

o-oi. In precipitation-hardened alloys, the description by a back

stress is particularly successful (Peterseim and Sauthoff,


necessarily

proposed

The most widespread idea is to introduce an internal back

stress, ai' developed in the material such that the


replaced

been

1984).

It

is

not

assumed that 0i is a constant threshold stress, but it may vary as

a function of the dislocation structure and the precipitate structure. Gibeling


and

Nix (1980) have prepared a review on observations and models pertaining to

the concept of internal stress. The idea of a variable internal stress

is

em-

ployed also in the description of primary creep by Robinson (1978) and Pugh and
Robinson (1978) (Section 25.2), and in the constitutive model of Hart (1976).

1.

The acceleration of creep in the tertiary stage of the creep


fracture

are

generally

Creep Deformation

curve

and

final

ascribed to the progressive accumulation of damage in

the material during creep. Several possible mechanisms are briefly described in
Chapter 2, while the special mechanism of grain boundary cavitation is examined
in greater detail in Part II.

1.3 Diffusion Creep


So far we have only mentioned creep deformation by climb-controlled dislocation
motion.

Creep

can occur also by diffusional flow of atoms from parts on grain

boundaries (or other interfaces) where they


which

are

under

tension.

This

are

under

stress-directed

compression

flow

of

to

specimen longer in the tensile direction. Grain boundaries play a central


for

diffusion

creep

parts

atoms renders the


role

since they can accommodate or release atoms or, in other

words, generate or annihilate atomic vacancies. Within a perfect crystal

latt-

ice, the generation of a vacancy-interstitial pair requires far higher energies


than the generation of a vacancy at a grain boundary, and

such

high

energies

are usually not available from thermal lattice vibrations.


If grain boundaries
vacancies,

the

are

the

only

possible

sources

and

sinks

for

atomic

rate-controlling step of diffusion creep is the diffusion over

distances of the order of


represents

the

driving

the
force.

grain

size

d,

while

the

applied

stress

The fundamentals of stress-directed diffusion

will be outlined later. At this stage only the final result of

Nabarro

(1948)

and Herring (1950) for the diffusional creep rate is given:


(1 .4)

Here, Dv is the diffusion coefficient in the grains, g is atomic volume,


grain

is

size

and a v is a dimensionless numerical factor whose magnitude depends


on the shape of the grains. For equi-axed, hexagonal grains is av = 24. If the
atoms diffuse along grain boundaries rather than through the grains, then
(1. 5)

(Coble, 1963). The factor ~ has a value of around 50, and 6Db is the grain
boundary diffusion coefficient, which has the physical dimension m3 /s and has
the

usual

temperature

dependence

of

thermally

activated

process,

1.3

Diffusion Creep

6Db = 6Dboexp(-Qb/RT). Numerical values for the activation energy and the
pre-exponential factor are given in Appendix A for a few materials.
As eqs. (1.4) and (1.5) show, diffusion creep
relationship

between

strain

rate

and

is

characterized

stress.

Coble

by

linear

creep predominates at

intermediate temperatures and small grain sizes, while Nabarro-Herring creep is


preferred

at

higher

temperatures and larger grain sizes. The ranges in which

different deformation mechanisms dominate will be shown systematically

on

the

deformation-mechanism maps introduced in Section 1.6.

1.4 Inhibition of Diffusion Creep


Experimentally, the predictions of eqs. (1.4) and
creep

rates

(1.5)

for

the

diffusional

have been confirmed for a great number of metals and other mater-

ials. At very low stresses, however, there seems to

be

threshold

behavior

such that a in eqs. (1.4) and (1.5) must be replaced by O-Oth where 0th is a
threshold stress. In pure metals at around half the melting temperature, the
threshold
0th

role

stress

is

generally

well

below

1 MPa

except

for

silver where

1.25 MPa (Towle and Jones, 1976). These small threshold stresses play
in

no

usual creep tests and their measurement requires a special technique,

namely, the creep of helical spring specimens.


in metals which contain high-melting particles on
threshold

stress

may

be

larger

and

the

grain

boundaries,

diffusion creep may even be suppressed

altogether, since at higher stresses dislocation creep intervenes. The


has

been

reviewed

by

the

subject

Burton (1977), by Gibeling and Nix (1980) and by Arzt,

Ashby and Verrall (1983). Sritharan and Jones (1980) and several other

authors

observe an increase in threshold stress in commercial stainless steels as carbide precipitation on grain boundaries progresses. However, the threshold stress
never

exceeds

2.6 MPa.

(The

value of 40 MPa quoted by Arzt et aI, 1983, for

stainless steel is not a threshold stress for diffusion creep according to

the

paper by Evans and Knowles, 1980, which Arzt et al refer to). For oxide-dispersion strengthened superalloys, Arzt et al report values up to 0th
,The inhibition

of

difficulty

accommodate

to

diffusion

creep
atoms

has

been

ascribed

to

1976).

Each

of

52 MPa.
conjectured

in grain boundaries (Ashby, 1969) or in the

interface between hard grain boundary particles and the matrix


Harris,

the

(Burton,

1973,

these authors suggests a mechanism how the inhibited

1.

Creep Deformation

plating of atoms can be enforced by stresses exceeding the threshold stress. An


alternative

model,

in which the inhibition is bypassed by power-law creep, is

described as a by-product of the analysis in

Section

7.3.6.

The

subject

of

inhibition is taken up in that section since it is important in relation to the


nucleation and growth of
cannot

creep

cavities.

If

the

particle/matrix

interface

accommodate atoms, high stresses on the particles may result, which can

possibly promote cavity nucleation.

On

the

other

hand,

the

inhibition

of

temperatures

is

diffusion may impede cavity growth (Chapter 13).

1.5 Grain Boundary Sliding


A typical degree of freedom which becomes active at
grain

boundary

elevated

sliding. The process has at least three different aspects each

of which is related to a different size scale. On an atomic scale, the


ance

resist-

against sliding is determined by the mobility of grain-boundary dislocat-

ions. It is generally believed that in engineering high-temperature


this

materials,

intrinsic sliding resistance is negligible compared to the effect of hard

second-phase particles in the boundary. If the particles are effectively rigid,


grain boundary sliding cannot progress unless the particles are circumvented by
diffusional accommodation processes or by dislocation creep as described below.
Models

which

refer to either the dislocation level or the particle level were

jointly termed 'intrinsic sliding models' (Langdon


trinsic

models',

in

the

same

terminology,

are

and

Vastava,

those

individual grain boundary facets is constrained by the

1982).

'Ex-

in which sliding of

surrounding

polycryst-

alline array (Section 1.5.2).


1.5.1

The infinite grain boundary (an intrinsic sliding model)

An atomically planar grain boundary would slide


motion

of

actual rate of sliding is controlled by some


necessary

comparatively

easily

by

the

grain boundary dislocations. According to Raj and Ashby (1971), the


in

order

to

avoid

separation

accommodation

process

which

is

or overlapping of material, if the

sliding boundary contains hard particles or is irregular in shape. Probably the


most

important mechanism by which the accommodation is achieved is diffusional

flow of matter. Atoms are removed (or vacancies are depoSited)


tends

where

material

to overlap, and atoms are deposited where the motion of the grains tends

to open up gaps (Fig. 1.2). Rate-controlling is the diffusion of


irregularities

or

particles

atoms

around

in the grain boundary. Under these premises, Raj

1.5

Grain Boundary Sliding

and Ashby (1971) show that the sliding rate, u b '

is

related

to

the

applied

shear stress Tb in a linear viscous manner,


( 1 .6)

The viscosity, or friction coefficient, n, depends on the shape


icles

or

of

the

part-

of the irregularities. For a grain boundary having a sinusoidal form

with wave-length A and amplitude h/2, for example, n takes the form
(1.7)

If grain boundary sliding is obstructed by rigid, cube-shaped particles of size


p and spacing Ap in the boundary, as shown in Fig. 1.2b, then the result is
( 1 .8)

oD i is the diffusion coefficLent


which is, however, rarely known.

where

(a)

/ r?
Tb. Ub

along

the

particle/matrix

interface,

(b)

diffusive flux

(d) /

~
A

Tb b

shear crack

in an infinite body

Fig. 1.2. Sliding of grain boundaries having a sinusoidal form (a) or


containing hard particles (b). Sliding in a polycrystal (c) and
idealization by a shear crack model (d).

10

Creep Deformation

1.

It should be mentioned that measurements compiled by Evans


stronger
which

dependence

follows

sliding

will

from
be

of

the

sliding

diffusive

(1984)

indicate

rate on stress than the linear relation

accommodation.

Nevertheless,

grain

boundary

described as a linear viscous process below. For the present

purposes, the intrinsic sliding properties of

grain

boundaries

need

not

be

in great detail, since most frequently the limiting case of free slid-

modeled

ing will be considered anyway.


1.5.2

Grain boundary sliding in polycrystals (extrinsic sliding models)

It is clear that a polycrystalline array can generally not be deformed by grain


boundary

sliding alone without deformation of the grains, i.e. sliding is con-

strained by the surrounding material. The possibility to accommodate sliding by


crack formation in grain boundaries is excluded in this section.
If the grains deform by diffusional flow, the effect of grain boundary
is

sliding

merely a slight modification of the numerical factors in the rate equations

for Nabarro-Herring creep and Coble creep (Raj and Ashby, 1971).
The situation is more complex if grain
power-law

creep

of

the

grains.

The

boundary

sliding

is

accommodated

by

rate of sliding may then be controlled

either by the intrinsic viscosity of the grain boundary or by the accommodation


process.

At high stresses, accommodation by power-law creep occurs readily and

the boundaries behave comparatively rigidly. At low stresses, however, the rate
of

sliding

is

controlled

by creep accommodation, while the boundaries slide

effectively freely.
The weakening effect of grain boundary sliding on

creeping

polycrystal

is

conveniently expressed by the stress-enhancement factor, f, defined as follows.


If ~ = Bon is the creep rate for a material with non-sliding grain

boundaries,

the same material with freely sliding grain boundaries creeps at a greater rate
( 1 .9)

Crossmann and Ashby (1975) and Gharemani (1980) analyze a two-dimensional array
of

hexagonal grains as in Fig. 1.2c using the finite element method. Gharemani

finds that the stress-enhancement factor, f, ranges from 1.16 to 1.19 for n = 1
to

4. Later in this book, a shear-crack model for grain boundary sliding (Fig.

1.2d; Riedel, 1984b) is described which, for hexagonal grains, leads to

1.5

11

Grain Boundary Sliding

[1 + (n/8)(n/3)1/2]1/n

or

(1.10)
[1 - (n/8)(n/3)1/2]-1/n.

The first form is obtained if the material surrounding the shear crack has
creep

E=

properties of the grains,

Ban, whereas the second is obtained if, in

the spirit of self-consistent methods, the


properties

of

the

material

is

given

the

effective

polycrystal with sliding grain boundaries described by eq.

(1.9). Equation (1.10) approximates the finite


well.

the

element

results

of

Gharemani

Beere (1982) summarizes approximate results of previous workers, e.g. of

Chen and Argon (1979) who find somewhat greater


1.43 for n

1 to

values,

namely,

1.33

to

factor

is

m.

For three-dimensional

arrays

of

grains,

the

stress-enhancement

greater. The shear crack model, in its self-consistent form, leads to


(1 _

8 sin(2a)
)-l/n
3n (1+3/n)1/2
'

(1.11)

where a is the angle of the grain boundary defined in Fig. 1.2c. For n = 1
a

30 0 ,

eq.

(1.11)

yields

= 1.58,

a value which compares well with other

estimates for n = 1: Chen and Argon (1979) calculate f


and

and

1.67,

while

Anderson

Rice (1985) find.f = 1.63 and 2.9 for two different geometrical grain con-

figurations (for details of the latter paper, see Section 12.6.3).


The characteristic stress to distinguish between high stresses where the
boundaries

behave

slide effectively

effectively
fre~ly

grain

rigidly, and low stresses where the boundaries

is given by
=

(B

n d)-l/(n-l),

(1. 12)

where n is the grain boundary viscosity coefficient and d is grain size.

1.6 Deformation-Mechanism Maps


Deformation maps are diagrams in the stress-vs.-temperature plane in which
areas

are

indicated

where

the

a particular creep mechanism predominates (Ashby,

1972). The idea is that different

mechanisms

operate

independently

and

the

1.

12

fastest

one

determines

the

deformation behavior and occupies the respective

regime in the stress-temperature plane. The boundaries


regimes

are

Creep Deformation

between

the

different

usually calculated (rather than measured) by comparing the strain

rate equations for the different mechanism. Figure 1.3 schematically shows such
a

deformation

map.

The

boundary

between Coble and Nabarro-Herring creep is

calculated by equating the creep rates given in eqs. (1.4) and (1.5) for

these

two mechanisms, which leads to


( 1.13)
Because of the same (linear) stress dependence of
mechanisms,

eq.
at

two

diffusional

creep

(1.13) is represented by a vertical line in Fig. 1.3, and the

mechanism with the greater


dominates

the

higher

activation

temperatures.

energy,

i.e.,

Similarly,

yields the boundary between Nabarro-Herring

creep

Nabarro-Herring

creep,

equating eqs. (1.1) and (1.4)


and

the

dislocation-climb

controlled regime, viz.


(1.14)

(JIG

Since now the temperature dependencies of both mechanisms are the same (that of
lattice

self-diffusion),

eq.

(1.14)

is

represented

by

a horizontal line.

Finally, Coble creep is separated from the dislocation-climb regime by the line

10 -/ - - - -ideal strength- - dislocation glide


l----~----~--~--~

yield
stress

elastic strain
predominates

dynamic
recrystallisation

NabarroHerring
~-....,...--t creep

for long times


10- 6 ' - -_ _ _ _ _ _ _---'-__________.....

0.5

homologous temperature TITm-

Fig. 1.3. Deformation-mechanism map (schematic).

1.6

13

Deformation-Mechanism Maps

defined by equating eqs. (1.1) and (1.5) which leads to


(1.15)
The regime of dislocation climb is subdivided by
which

runs

more

or

strain

rate,

dashed

lines.

The

one

less horizontal, represents eq. (1.12). It separates the

low-stress regime where grain boundary


total

two

from

behave as if they were

the

rigid.

sliding

high-stress
The

vertical

contributes

markedly

to

the

regime where the grain boundaries


dashed

line

separates

vacancy

diffusion along dislocation lines (low temperatures) from diffusion through the
lattice (high temperatures).
Towards high stresses, the range of validity of the creep mechanisms is cut off
by

the onset of extensive dislocation glide. The plastic strain in this regime

is attained

very

quickly

after

load

application

and

can

practically

be

described by instantaneous, rate-independent plastic yielding in many cases.


The greatest practical advantage of the deformation maps probably lies
guidance

which

they

provide

in

in

the

the extrapolation of creep data. Most creep

tests in the laboratory are done within a year or less, sometimes within a

few

years, whereas engineering structures like electric power-generating plants are


designed to last for at least 25 years. The laboratory data can be extrapolated
to

the in-service conditions with some confidence only if both lie in the same

regime of the deformation map. The maps also are useful tools for the

designer

to choose the proper ,constitutive law for the stress analysis of a given part.
The most comprehensive collection of material data that are
construction

relevant

the

of deformation maps has recently been compiled by Frost and Ashby

(1982). They give deformation maps for over 40 materials covering pure
commercial

for

alloys

and

metals,

ceramics. Material parameters have also been compiled,

e.g., by Frost and Ashby (1977), Needleman and Rice


Ashby (1981). Table A.1 in Appendix
convenient use in thi$ monograph.

(1980)

and

Swinkels

and

A gives a selection of these data for

2 Introduction to Creep Fracture


and Other Fracture Modes

2.1 The Nature of Creep Damage


The acceleration of creep in the tertiary stage of the
caused

by

the

formation

and

creep

curve

is

often

joining of micro-cavities on grain boundaries.

Creep fracture is therefore generally intergranular. The cavities may be nucleated early in the creep life, possibly even during the primary stage. Initially
their effect on the creep rate is negligible but, as their number and size
crease,

in-

they weaken the material progressively and finally induce failure. The

nucleation and growth of cavities is studied extensively in


book.

Part

II

of

this

Accelerating creep rates can also be caused by a degradation of the microstructure

of

the material. Many engineering alloys contain second-phase particles,

which serve as obstacles against dislocation motion and therefore provide creep
strength

to

the

material.

In

long-time service, the possible growth of the

larger of these particles and the disappearance of the smaller ones leads to
gradual loss of

cr~ep

In practice, this may be as important as grain-boundary cavitation, but,


it

is

resistance or, in other words, to increasing creep rates.


since

not the subject of this book, only a few pertinent references are given

here. Particle coarsening and concommittant softening of the material

are

ob-

served in creep-resistant CrMo steels (Williams and Wilshire, 1977, 1981; Williams and Cane, 1979; Pizzo and Mandurrago, 1981), in

many

Y'-hardened

nickel

base superalloys (Burt, Dennison and Wilshire, 1979; Stevens and Flewitt, 1981;
Dyson and McLean, 1983), and in aluminum alloys (Claeys and Jones, 1984). Dyson
and

McLean (1983), point out that the pronounced tertiary stage of superalloys

cannot be explained by particle coarsening alone. At least part of the accelerating

creep rate must be ascribed to grain-boundary cavities and cracks and to

changes in the dislocation microstructure. Henderson and


propose

that

McLean

(1984,

1985)

the dislocation structure developing in the interface between Y'

particles and the matrix is of primary importance.

2.1

The Nature of Damage

15

Another possible cause for accelerating creep rates in the

tertiary

stage

is

corrosive damage at or below the surface, such as internal oxidation, sometimes


selectively along grain
initiation

by

boundaries,

and

subsequent

crack

formation.

Crack

corrosion has been observed, for example, by Schnaas and Grabke

(1978) in an austenitic steel. Also, some superalloys fail

by

the

growth

of

oxidized surface cracks (Burt, Elliott and Wilshire, 1981). The role of chemical environment in high temperature fracture is an important subject area
cannot

be

which

covered systematically within the scope of this book. Only effects,

which are closely related to grain-boundary cavitation, like hydrogen attack or


oxygen

attack are described in Chapter 9. A brief overview on other effects of

high-temperature corrosion is given by Ashby and Dyson (1984).

2.2 Fracture-Mechanism Maps


Crystalline solids can fracture by one
subject

of

this

of

several

mechanisms.

good to remember which alternative mechanisms exist and


temperature

regime they

predomina~e.

Ashby,

1977).

In

analogy

to

in

which

stress

fracture-mechanism

map

and

(Wray,

the deformation maps, fracture maps are

diagrams with tensile stress on one axis and temperature on the other.
stress-temperature

the

In a somewhat simplified picture, many of

the important mechanisms can be displayed on a


1969,

Although

book is high-temperature fracture at low strain rates, it is

In

the

plane, the regions are indicated within which the different

fracture modes are found to prevail.


Figure 2.1 schematically shows such a fracture map. The location of the boundaries

between the different mechanisms depends on the material, its heat treat-

ment, the chemical environment and on the other stress components


tensile

besides

the

stress. The applied load is assumed to be constant, i.e. fatigue fail-

ure under cyclic loads is not considered. Actual fracture maps

have

been

put

together by Ashby et al (1979) and by Gandhi and Ashby (1979a) for face-centered cubic metals and alloys using observations of the fracture modes reported in
the literature. Gandhi and Ashby (1979b) give maps for a great number of materials which are capable of cleavage, i.e.,
metals

body-cenetred

cubic

and

hexagonal

as well as covalent ceramics and ionically-bonded materials like ice or

rock salt. Fracture maps have been constructed for iron and steels by Fields et
al

(1980), and for titanium alloys by Krishnamohanrao et al (1986). The mecha-

nisms appearing on the fracture map will be described briefly below.

2.

16

slip - induced
'0-2 cleavage
"
orB/F
Tt

insittntaneous fracture

....

T
cleavage or

10-3

BIFfrom
largest crack

lJ..J

Introduction to Creep Fracture

intergranular
creep fracture

10-'

no fracture

10- 5
0

--

0.5
TITm

Fig. 2.1. Fracture-mechanism map (schematic).

2.2.1

Cleavage and brittle intergranular fracture

Many materials such as the body-centered cubic and


well

as

most

ceramics

the

hexagonal

metals,

as

and ionically bonded materials, fracture in a brittle

manner at low temperatures and high stresses. Brittle fracture in this range is
brought

about

by

the formation of cracKs and their propagation along certain

crystallographic planes. This mode of fracture is called cleavage.


for

fracture

An

example

surface with cleaved grains is shown in Fig. 2.2a. Most face-

centered cubic metals do not cleave, for reasons which are plausible from
siderations

con-

of the balance between cleavage and dislocation formation at crack

tips (Rice and Thomson, 1974).


Cleavage fracture of ferritic steels has been the subject of extensive investigations

(see Chapter 7 of Knott's textbook, 1973; for a more recent paper, see

Riedel and Kochend6rfer, 1979). It was found that cleavage fracture is


initiated

a coarse carbide particle. At very low temperatures, cleavage


ready

formed,

cracks

are

al-

when plasticity is still confined to single, favorably oriented

grains. At somewhat higher temperatures


plastic

usually

by plastic slipping or twinning, often where a slip band impinges on

cleavage

occurs

only

after

general

yielding of the whole polycrystal. This means that the strain to fail-

2.2

Fracture-Mechanism Maps

17

Fig. 2.2. a) Cleavage fracture surface of a coarse-grained carbon steel


fractured at -100 0 C. b) Brittle intergranular fracture of a low-alloy
NiCrMoV turbine rotor steel containing 0.048%P, tempered 816 h at 400 0 C,
fractured at -100 0 C (courtesy Moller, Erhart and Grabke).

ure increases from below 1% to sometimes over 10%. Because


low

ductilities

of

the

relatively

associated with cleavage fracture, it is usually necessary in

engineering applications to avoid temperatures where cleavage predominates.


In ceramic materials it is common that cleavage fracture starts from pre-existing

microcracks before the stress for plastic yielding is attained. The almost

inevitable brittleness of ceramics must be taken into account in the design

of

structural parts.
Instead of fracturing by transgranular cleavage, many metals and ceramics
by

fail

brittle intergranular fracture (BIF) at low temperatures. The balance betw-

een these two modes seems to be a delicate one, so that both can occur simultaneously

in

the same specimen. Small impurity additions may shift the fracture

path from transgranular to intergranular. In ferri t ic steels,


is

known
at

phenomenon

as temper embrittlement (cf. Section 8.1.1). A typical fracture sur-

face is shown in Fig. 2.2b. Ordered intermetallic compounds


BIF

this

high

temperatures

sometimes

exhibit

such as Ni3Al at 600 to 800 0 C (Liu, White and Lee,

1985). This has been ascribed to an environmental interaction with oxygen.

18

2.

Introduction to Creep Fracture

Fig. 2.3. Dimpled fracture of a structural steel at room temperature.

2.2.2

Ductile transgranular fracture by plastic hole growth

Above the brittle-to-ductile


changes

from

tran~ition
temperature, Tt , the fracture mode
cleavage, or brittle intergranular fracture, to a ductile trans-

granular mode. Here, separation of the specimen halves is brought about by

the

growth and coalescence of holes, which are nucleated at second phase particles,
such as MnS-particles in steels, and which grow by plastic flow of
ial.

This

fracture

mode

is

dimpled fracture surface is shown


engineering

alloys

the

mater-

called dimpled, or fibrous, fracture. A typical


in

Fig.

2.3.

At

room

temperature,

most

fail by this fracture mode, and an extensive literature on

the subject exists. For reviews see, for example, Goods

and

Brown

(1979)

or

applied

stress

is

Chapter 8 of Knott's textbook (1973).


Fracture by hole growth may occur instantaneously, if
high

enough,

or

slowly

by

creep

flow

if

temperature is relatively high. Hole growth by

the

the stress is lower, and if the


time-dependent

creep

flow

is

discussed also in Chapter 14 of this book.


2.2.3

Necking and superplasticity

The strain to failure permitted by the plastic hole growth mechanism is


ally

gener-

high. Theoretically, strains of well above 100% are expected (cf. Chapter

14). In tensile tests, however, there is the possibility of a necking instabil-

2.2

Fracture-Mechanism Maps

19

QJ

bJ

Fig. 2.4. (a) Partial and (b) complete necking of a round bar
in a tensile test.

ity which may develop at smaller strains and which may terminate the test prior
to void coalescence in the uniformly strained parts of the specimen (Fig. 2.4).
Necking is not a microscopic fracture mechanism like the
the

fracture

map.

mechanisms

shown

on

Rather it is a macroscopic instability of the tensile test

which may intervene, whenever the microscopic mechanism allows for sufficiently
high

strains.

This

is

the case in the plastic hole growth regime and in the

regime of dynamic recrystallization (Section 2.2.5). A necking


commonly

observed

when

which the material response


temperature.

instability

is

ductile metals are tested in the temperature range in


is

essentially

rate-independent,

e.g.

at

room

The onset of necking is then described by the classical Considere

criterion (see, e.g., Hart, 1967).


At higher temperatures, the material response is
idealized

as

nonlinear

viscous

creep,

as

time-dependent,

and

can

be

described by eq. (1.1). For such

power-law viscous materials, Hart's (1967) analysis predicts that

the

necking

instability should commence directly at the start of the test if the stress
exponent is n > 1. In reality, however, one does not observe necking at small
strains

under

creep conditions. Hutchinson and Obrecht (1977) point out where

this comes from; the time constant for the neck to develop

is

so

large

that

strains of the order 100% can be achieved before necking leads to failure.
If the stress exponent n is not very much larger

than

unity

(practically

it

20

2.

suffices

that

2),

the

material can even become sijperplastic. This means

that in a slow tensile test at elevated


hundred

Introduction to Creep Fracture

temperature,

elongations

of

several

per cent can be reached without necking. The vast literature on super-

plasticity has been reviewed, for example, by Padmanabhan and Davies (1980) and
by Gittus (1982). If a viscous material having a stress exponent n
develop voids, it can still sustain strains of more than 100%

2 tends to

(Langdon,

1982;

Stowell, Livesey and Ridley, 1984).


Following Hoff (1953) and Hutchinson and Obrecht (1977), we now
time

after

which

calculate

necking leads to failure of a tensile specimen under a con-

stant load. The material is assumed to be nonlinear viscous and to


internal

damage

such

develop

to

no

as voids. The specimen is cylindrical, small deviations

from the ideal cylindrical shape being admitted if they have a long
compared

the

wavelength

the diameter of the specimen. Such nonuniformities may be present

from the start of the test or may develop during straining. As long as the nonuniformities

have

a long wavelength, the stress state is locally uniaxial and

uniform over the section of the specimen.

It

is

clear,

then,

that

rupture

occurs when the minimum initial cross section has been strained to zero area.
If A(t) is the current cross sectional area of the

specimen

at

its

thinnest

location, and if P is the applied load, the stress in the minimum cross section
is

0 =

P/A(t), and the strain rate, from Norton's law, is

(2.1)

In the second form, the initial minimum area, Ao' and the- initial strain
ES =

B (P/Ao)n,

the steady-state, or secondary, strain


sectional

area

rate,

in the minimum cross section are used; the subscript s denotes


becomes

marked.

rate

before

the

reduction

Another relation between A and

E is

of

cross

obtained

from the requirement that the material is incompressible:


E =

- AlA.

Equating from eqs. (2.1) and (2.2) gives a

(2.2)
differential

equation

for

A(t)

which can be solved by separation of the variables:


(2.3)
Fracture occurs when A(t)

0, i.e. when

2.2

Fracture-Mechanism Maps

21

(2.4)

lin,

where t f is the time to rupture, and Estf is called the Monkman-Grant product
(see Section 2.3.1). The strain to failure at the minimum cross section is
infinite

in

elongation

the
of

no-damage

the

material

considered

here,

while

overall

the

specimen depends on the initial nonuniformity of the cross

section.
2.2.4

Intergranular creep fracture

At lower stresses and elevated temperatures, hole growth by


or

by

creep

flow

plastic

straining

becomes so slow that fracture by grain-boundary cavitation

intervenes. The strain to fracture is then relatively low. This mechanism


be

described

will

extensively in Part II. On the fracture map, the field of inter-

granular cavitation is bounded on the left and below by a regime

in

which

no

fracture occurs within any practically interesting time. This may be so because
cavity growth, which is often diffusion-controlled, becomes exceedingly slow at
lower temperatures, or because no cavities are nucleated at very low stresses.
2.2.5

Rupture by dynamic recrystallization

At very high temperatures, diffusion becomes so rapid that


dynamic

recrystallization

commences.

In

at

high

stresses,

this regime it is favorable for the

metal to resolve dislocation tangles by forming new grains in heavily


regions

rather

than

deformed

by the recovery processes operating in the creep regime.

Since grain boundaries can now migrate, cavities, if they are nucleated at all,
are

continuously

cut off from the supply of vacancies arriving by grain boun-

dary diffusion. In addition, dynamic recrystallization is strongly


pendent

and

depends

de-

tensile specimen is therefore prone to necking before cavities

can coalesce. The boundary between rupture by recrystallization and


failure

stress

cavitation

strongly on the cleanliness of the materials. Precipitates on

grain boundaries, which are usually present in commercial materials to minimize


grain boundary sliding, also prevent grain boundary migration. In addition,
precipitates tend to favor cavitation by nucleating cavities.
2.2.6

Fracture modes at very high temperature

The regime of rupture by recrystallization may be interrupted once


times

by

ductility-dips.

or

several

In steels, for example, aluminum nitride or niobium

2.

22

Introduction to Creep Fracture

carbo-nitride particles can be precipitated at high temperatures. The particles


pin

the

grain

boundaries

thus

increased strength, but also to


ductility-dip

near

1000 0 C

inhibiting
brittleness.

recrystallization. This leads to


Another

possible

between iron sulfide and iron (Melford, 1980). This phenomenon


shortness

and

may

play

cause

for

is the formation of molten phases like an eutectic

role

is

called

hot

in fabrication processes such as continuous

casting.
As the solidus temperature of the alloy

is

approached,

grain

boundaries

or

interdentritic material start to melt. This bounds the region of recrystallization rupture towards high temperatures. Welding defects may sometimes
liquidation

be

such

cracks caused additionally by thermal stresses which accompany the

welding process (Robinson and Scott, 1980).


It should be kept in mind that the fracture map shown
include

effects

in

Fig.

2.1

does

of the chemical environment such as stress corrosion cracking

and hydrogen embrittlement at lower temperatures, hydrogen attack at


iate

temperature

not

intermed-

and effects like carburization or internal oxidation at high

temperatures. Further, Fig. 2.1 is valid for static load. Cyclic loading

leads

to new failure modes. In the following, interest will be focussed on the regime
of intergranular creep fracture with a few excursions to environmental

effects

and to fatigue loading.

2.3 Empirical Formulas for the Rupture Time in the Creep Regime
There are several empirical formulas in
sophistication,

describe

the

use

which,

on

different

levels

of

dependence of the rupture time on stress and on

temperature.
2.3.1

The Monkman-Grant rule

Monkman and Grant (1956) observed that the product of

the

steady-state

creep

rate, ~s' and the rupture time, t f , is approximately independent of the applied
stress and of temperature:

(2.5)
Later work has confirmed that the

Monkman-Grant

'constant'

is

indeed

inde-

2.3

23

Empirical Formulas for the Rupture Lifetime

pendent

of stress and temperature over a wide range of conditions. Feltham and

Meakin (1959), for example, demonstrate its approximate


between

~OOoC and

of data by Evans

750 0 C

constancy

for

copper

and lifetimes from 1 minute to 10 hours. A compilation

shows that for different materials CMG typically lies


in the range 0.03 to 0.3. In the light of the many possible mechanisms of creep
(198~)

rupture, the approximate constancy of the Monkman-Grant product is


ingly

simple

surpris-

observation, which may serve as a guideline to select the models

having most practical relevance.


The practical advantage of the Monkman-Grant
creep

rule

is

that

the

steady-state

rate can be measured early in the creep life. Provided that the value of

CMG has been determined in relatively short-term tests, the lifetime in a longterm test can be predicted from t f = CMG/ES once the steady-state creep rate
has been established. This means a substantial reduction of testing time. Of
course, to utilize this possibility, the range of validity of the Monkman-Grant
rule must be known. The modeling of the mechanisms should

contribute

to

that

knowledge.
The Monkman-Grant rule is closely

r~lated

to a critical-strain criterion. Often

the

strain to failure is some fixed multiple of CMG The Monkman-Grant product


represents the strain which would be accumulated if steady-state creep predominated

during

the whole creep life, while the total strain to failure also in-

cludes primary creep and tertiary creep, as well as the effect of necking if it
OQcurs.

In

engineering

alloys

the

greatest

accumulated in the tertiary stage. It should be

amount of strain is frequently


noted

that

also

failure

by

complete necking leads to a constant Monkman-Grant product having the value lin
[eq.

In a tensile test, this represents an upper bound for CMG , whereas


damage such as grain boundary cavitation reduces that value. A com-

(2.~)J.

internal

parison of observed values of CMG and lin (Evans, 198~) shows that in materials
developing cavities is indeed CMG < lin, whereas in others is CMG = lin.

2.3.2

The Sherby-Dorn parameter

Orr, Sherby and Dorn (1954) suggest that the rupture


temperature T are related by

time

tf,

stress

-Q/RT
a(a) ~ t f e
,
where Q is a fittable activation energy, R is the. gas constant and the

a and

(2.6)
Sherby-

24

2.

Dorn

parameter,

Introduction to Creep Fracture

9(0), is an empirical function of stress but not of temperat-

ure. Once 9(0) and Q have been determined, eq. (2.6) can be used to interpolate
and extrapolate to other stresses and temperatures.
2.3.3

The Larson-Miller parameter

Similar in principle but different in its functional form is


Larson

and

Miller

the

approach

of

(1952) who assume the following relation between lifetime,

stress and temperature:


P(o) = (20

log t f ) T/1000.

(2.7)

Here P(o) is the Larson-Miller parameter whose dependence on stress is


ined

empirically,

determ-

t f is in hours and T in Kelvin. The Larson-Miller parameter

is widely used to characterize the rupture behavior of engineering materials.


There are other related approaches, most of which are also useful to
measured

data

in

concise

form.

temperatures than those covered experimentally, however, are


is

easy

to

imagine

examples

represent

Extrapolations to longer times and lower

where

questionable.

It

different approaches lead to completely

different lifetime predictions.


2.3.4

The Kachanov equations

Another type of phenomenological description of the rupture behavior is due

to

Kachanov (1960) and Rabotnov (1969). Here, the evolution of a damage parameter,
w, is conSidered, which is not meant to be a physically identifyable

although

it

is

somehow

related

to

the

cavitated

boundary. The damage parameter varies from w = 0 for


w= 1

at

failure.

area
the

quantity,

fraction of grain

virgin

material

to

It is an internal variable (i.e. not directly measurable),

which is assumed to obey the kinetic law

D aX

(2.8)

(l+~)(l-w)~ ,

and which affects the relation between the

measurable

quantities

stress

and

strain rate according to


(2.9)

2.3

Empirical Formulas for the Rupture Lifetime

25

The superposed dot denotes the time derivative, and B, D, n, X and


justable

parameters

which

depend

are

ad-

on temperature. Numerical examples will be

shown in Table 27.1 in Chapter 27.


For time-independent stress, eq.
resulting

(2.8)

can

readily

be

integrated.

If

the

w is inserted into eq. (2.9), the strain rate can also be integrated

with the result that the strain increases in time according to


(2.10)
which is plotted in Fig. 2.5. Here, the time

to

failure

and

the

strain

to

failure are given by


(2.11)
(2.12)
The initial, or steady-state, strain rate is s
Multiplying

with

the

rupture

time

from

Ban, since

at

O.

eq. (2.11) gives the Monkman-Grant

product
(2.13)

o.

0.5
t/t, --Fig. 2.5. Normalized creep curves from Kachanov model, eq. (2.10).

26

2. Introduction to Creep Fracture

Thus the Kachanov equations imply a


n

X and

Band

D have

Monkman-Grant product is

the

the

constant

same

Monkman-Grant

product

only

if

temperature dependence. In any case the

following

stress-independent

fraction

of

the

strain to failure
(2.14)
This fraction, as well as the shape of the normalized creep curve shown in Fig.
2.5,

depend

only

on the ratio n/($+1). Obviously, the Kachanov equations are

designed to model secondary and tertiary creep, whereas primary

creep

is

not

described.
In Chapter 27, a generalization of the Kachanov equations to
states

of

stress

will

be

described.

Together

with

three-dimensional

the

equilibrium

and

compatibility conditions they form a basis for what is called damage mechanics.
These equations will be applied to model creep crack growth.
2.3.5

The a-projection concept

The a-projection concept by Evans, Parker and Wilshire (1982) is an attempt


represent

creep

curves

mathematically

shape of the creep curve is described by


empirical

functions

over
an

a wide range of conditions. The


expression

which

contains

rupture,

another

four

parameters

are

parameters, the deformation and rupture behavior can be


and

it

is

four

of stress and temperature each of which can reasonably be

described using four fit table parameters (Evans, Beden and Wilshire, 1984).
incorporate

to

required.
described

To

With these 20
accurately,

claimed that the range of validity includes very long times, where

measurements are rare and expensive.

3 The Continuum-Mechanical Equations

Many of the quantitative models


fracture

mechanics

equations

of

equation,

the

concepts

solid

for
to

mechanics.

compatibility

the

be

creep

presented

These

condition

equations
and

failure

mechanisms

and

in this book are based on the


comprise

the

equilibrium

the material law for multi-axial

states of stress. The presentation of the basic equations in this chapter


be

concise,

and

the

reader

1966)

or

will

who is not familiar with solid mechanics at all

might wish to consult one of the textbooks on


Lifshitz,

the

the

subject

(e.g.

Landau

and

on elasticity theory (Timoshenko and Goodier, 1951; Love,

1952; Leipholz, 1968; Muskhelishvili, '977) or on plasticity theory (Prager and


Hodge, 1951; Hill, 1956a).

3.1 The Equations for Equilibrium and Compatibility


In the context of this book, body forces and

mass

inertia

can

generally

be

neglected. Then equilibrium of the volume elements of a stressed solid requires


that the stress

tenso~

obeys the partial differential equation

(i,j

1 to 3),

where 0ij are the components of the (symmetric) stress tensor

in

Cartesian

coordinate system x"x 2 ,x 3 and Vj = o/OXj is the gradient operator. The


Cartesian tensor notation is employed including the summation convention for
repeated indices. On the boundary of the body, the stress field must satisfy
the boundary condition nj 0ij = Ti' where nj is the outward normal unit vector
on the boundary and T. is the surface traction vector, which may of course
1
vanish on part of the surface. The analysis in this book is based on the
assumption

of

infinitesimally

small

strain and displacement. This implies a

linear relationship between the strain Eij and the displacement field uj ' viz.,

3.

28

The Continuum-Mechanical Equations

(3.2)
A strain field that is derivable from a displacement

field

according

the

displacement

to

eq.

(3.2) obeys the compatibility relation

:he same relations hold for the strain rate, ij'

and

rate,

uj .

3.2 The Material Law


The set of continuum-mechanical equations is completed
Different

forms

will

be

considered

with

by

the

material

law.

the common feature that the total

strain rate is a sum of the elastic strain rate and the nonelastic strain rate
+

(ne)
ij .

(3.4)

The superposed dot denotes the time-derivative. The elastic strain

is

related

to the stress by Hooke's law, which, for isotropic material, takes the form

(3.5)

Here, E is Young's modulus, v is Poisson's ratio, 0 .. is the unit tensor, and


lJ
0kk is the trace of the stress tensor; the summation convention for repeated
indices implies 0kk = 0" + 022 + 033. The elastic strain rate is related
the stress rate in the same way as elastic strain is related to stress.
The form of the nonelastic strain rate depends on which

range

temperature

where climb-controlled

is

considered.

dislocation motion
description
form, i(ne)

Under

predominates,

creep
the

conditions,

simplest,

commonly

of

used

stress

to

and

constitutive

of the nonelastic strain rate is Norton's creep law. The uni-axial


=

B on, in which Band n are material parameters, is generalized to

multi-axial states of stress using von Mises' J 2 -flOW theory. To account for
the incompressibility of creep flow, this theory assumes the strain rate to be
proportional to the deviator of the stress tensor, ij = ij - ijOkk /3 , and to
depend otherwise only on the second invariant of the
stress
tensor,

3.3

29

The Equations for 2-D Problems

Ge = (3GljGlj/2)1/2,
which is
Norton's law takes the form:

called

the

equivalent

tensile

stress. Then

e: (ne)

(3.6)

ij

This is a special case of nonlinear viscous material laws, which are characterized

by

the

fact that the nonelastic strain rate is a unique function of the

stress tensor.
More complex material laws will be introduced and applied
including

the

effects

in

later

the effect of damage on tertiary creep. Damage mechanics equations


be

introduced

chapters,

of strain hardening and recovery on primary creep, and

empirically

as,

for

example,

can

either

in the Kachanov model (Section

2.3.4), or one can try to derive them from models

for

cavity

nucleation

and

growth. The reader who is especially interested in the constitutive description


of the deformation behavior of materials is referred to the

book

Constitutive

Equations in Plasticity, edited by Argon (1975), (especially to the articles by


Rice, Hart et aI, Ilschner, and

Git~ys),

and to the papers by Pugh and Robinson

(1978), Schmidt and Miller (1981), and Estrin and Mecking (1984).

3.3 The Equations for Antiplane Shear, Plane Stress and Plane Strain
The constitutive equations presented in the
partial
(even

differential
numerically)

specialize

the

equations
for

general

preceding

are

nonlinear

three-dimensional

problems.

Therefore

we

equations for the two-dimensional problems illustrated in Fig.

3.1 using the crack geometry as an example. The


all

section

for stress and strain which are hard to solve

characterized

by

vanishing

derivatives

two-dimensional
of

stress

and

problems
strain

x3 -direction, i.e., parallel to the crack front. Three-dimensional

are

in the

aspects

in

fracture mechanics will be discussed in Chapter 22.


a) In antiplane shear the only non-zero components of displacement, stress
are

u3 '

G13 ,

and

e: 13 , e: 23 , respectively. Such a field develops if the


applied surface tractions, have x3-components only (Mode-III loading) and are

strain

G23 ,

independent of x 3 The dimension of


X"3-direction. Then the equations for
material law take the form

the

body

must

equilibrium,

be

infinite

compatibility

in the
and

the

3.

30

The Continuum-Mechanical Equations

Fig. 3.1. Mode I and Mode II can occur in plane strain and plane stress.
Mode III corresponds to antiplane shear.

(3.7)

(3.8)
2 )(n-1 )/2

0 23

where G = E/[2(1+v)] is the shear modulus and


law

has

been

body is ann!
0 23 = O.

B = /3 n + 1

(3.9)

1,2)

i3'

B. Here, Norton's power

inserted for the nonelastic strain rate. On the boundary of the


T3 . On

the

traction-free

crack

faces

this

specializes

to

It is sometimes convenient to replace the set of equations (3.7) to (3.9) by

single

equation for a stress function. In antiplane shear, a stress function

is defined such that the stress components are given by the derivatives
(3.10)
Inserting eq; (3.10) into eq. (3.7) shows that

stress

components

derived

in

this way from a stress function automatically satisfy the equilibrium equation.
If now the material law, eq. (3.9), is inserted into the compatibility
condition

for

the strain rate, eq. (3.8), and stress components are expressed

by the stress function through eq.

(3.10),

the

resulting

equation

for

the

stress function takes the form

o.

(3.11)

3.3

The Equations for 2-D Problems

31

This equation will be applied to crack problems in Section 23.2.1.


b) Plane stress is characterized by zero x3-components of stress, 03i = 0,
further

13 = 23 = O.

Such

stress

and

and

strain field develops in a plane

specimen that is infinitely thin in x3 -direction and is loaded only in the x1and x2 -directions, i.e. T3 = O. The compatibility relation is then Simpler than
in the general three-dimensional case, viz.,
(3.12)
The equilibrium condition retains its general form,
indices

i,j

take

eq.

(3.1),

however,

the

the values 1 and 2 only. In the material law, the vanishing

stress and strain components can be deleted. The equivalent tensile stess is
(3.13)
As in Mode III, the equilibrium equation can be satisfied identically,

if

the

stress components are derivatives of a stress function, which is now called


Airy's stress function and is denot~d by ~. In plane stress (as well as in
plane strain to be introduced shortly) is:
(i,j
The governing equation for
(3.4)

1,2).

is obtained by inserting the

(3.14 )
material

law,

eqs.

to (3.6), into the compatibility condition, eq. (3.12), using the stress

function representation for the stress. If Norton's creep law is

employed

for

the nonelastic strain rate, there results


(3.15)
For plane stress, the equivalent tensile stress is given by
(3.16)
For purely elastic problems, where the term multiplying B can be

neglected

in

eq. (3.15), the governing equation simplifies to


(3.17)

32

3.

The Continuum-Mechanical Equations

Solutions for both, linear elastic and nonlinear problems will be developed

in

later chapters.
c) A plane deformation field with 3i = 0 and 0'3 = 023 = 0 is called a plane
strain field. Such a field prevails in a specimen which is very thick in the
x 3-direction and which is subjected to boundary tractions (T"

T2 ), which are
of x3 The compatibility and equilibrium equations are the same as
those for plane stress. The equivalent tensile stress now takes the form
independent

= (

M) [(

'IY2

2 J' 12 .
0,,-022 ) 2 + ( 0"-033 ) 2 + ( 022-033 ) 2 + 6 '2

As in plane stress, the in-plane stress components are


the

derivatives

of

Airy stress function according to eq. (3.'4). If again the material law is

inserted

into

representation
~

second

(3.'8)

the

compatibility

condition

using

the

stress

function

of stress, the resulting equation contains two unknowns, namely

and the deviatoric stress component 033' A second

inserting the plane-strain condition, 33

equation

is

obtained

by

0, into the material law. Thus, for

plane-strain deformation of an elastic-nonlinear viscous material

one

obtains

the coupled equations:

(3.' gal

(3.,gb)
where
(3.20)

For linear elastic problems, this reduces to the


stress,

v4~

O.

Hence,

same

equation

as

in

plane

in the linear elastic case the stress fields are the

same for plane stress and plane strain if the boundary tractions are the same.
The situation where 33 is independent of all co-ordinates Xi but
called generalized plane strain.

non-zero

In both, plane stress and plane

and

loading
plane

(Fig.
(X 2

= 0)

3.').

If

and

the

the

strain,

we

distinguish

Mode-I

is

Mode-II

specimen is symmetric with respect to the crack

boundary

tractions

obey

the

symmetry

relations

3.3

The Equations for 2-D Problems

33

T1 (x 1 ,x 2 ) = T1 (x 1 ,-x 2 ) and T2 (x 1 ,x 2 ) = -T 2 (x 1 ,-x 2 ), this is called Mode-I


and
loading. For this tensile mode of loading, the stress components all'
a33 are symmetric with respect to the crack plane whereas a12 is antisymmetric.
Mode II (in-plane

shear)

is

characterized

by

symmetric

T2

and

a12 , and
antisymmetric all' a22 , a33 and T1 A linear superposition of Mode-I, Mode-II,
and Mode-III fields is only possible for linear material.

3.4 General Features of the Continuum-Mechanical Fields


Throughout the following chapters, a few arguments of general validity will

be

referred to repeatedly. They are listed below.


3.4.1

The elastic-viscous analogy (Hoff, 1954)

Comparing a (generally nonlinear) elastic material described by ij = f(a ij )


with a nonlinear' viscous material obeying the same functional relationship for
the strain rate, ~ij s f(a ij ), one recognizes that the constitutive equations
for the two materials are identical if strain rate in the viscous case is
identified with strain in the elastic case. Therefore the stress fields in
two

materials

the

must be the same if the applied loading is the same. Strain and

displacement in elastic material correspond to

their

time

rates

in

viscous

material.
3.4.2

Scaling properties for power-law materials (Ilyushin, 1946)

If strain rate (or strain) is a power-law

function

of

stress,

Ban,

the

stress field scales with the applied load, P, according to


(3.21 )
where W is a length representing the

specimen

dimensions,

represents

the

spatial coordinates and a ij is a dimensionless function, which depends on the


specimen shape but not on its size nor on the magnitude of the load. The
independence

of the specimen size holds for all material laws which contain no

characteristic length, but the independence of the load is unique for power-law
materials.
To prove the validity of eq. (3.21), insert it into the equilibrium

condition,

3.

34

and into the boundary condition for the tractions. Obviously, P/W 2

(3.1),

eq.

The Continuum-Mechanical Equations

is a common factor to all terms and can therefore be deleted. Analogously, if

2 n the strain rate field, Eij = B (P/W) Eij(r/W), which follows from eq. (3.21),
is inserted into the compatibility condition, the common factor B (P/W 2 )n can
be deleted. There remains a system of equations for 0ij and Eij which is indeed
independent of the load and of the absolute specimen size.
The scaling properties defined by eq. (3.21)
elastic

power-law

and

viscous

are

materials.

not

confined

Rather,

they

incremental, power-law plasticity as long as the loading is


no

unloading

(i.e.,

the

applied

loads

increase

parameter). Even under cyclic loading, the


that

the

hysteresis

loops

obey

as

measured

from

the

upper

also

apply

proportional

properties

to
with

hold

provided

power-law relation, OE

B(oo)n,

independent of the loading amplitude, where OE and 00


differences

power-law

in proportion to a scalar

scaling

unique

to

or

are

lower

strain

reversal

and

stress

pOints of the

hysteresis loop. Other classes of power-law materials exhibit the same

scaling

properties, too, as will be shown in later chapters.


For all classes of
distribution

is

materials
the

same

having
as

that

these
in

scaling

properties,

the

stress

power-law elastic material since the

governing equations reduce to the same form. Advantage can then be taken of all
the

favorable

problems

properties

there

are

of

nonlinear elastic fields. For example, in crack

path-independent

integrals,

the

crack-tip

fields

are

available, and tabulated solutions for many specimen geometries exist.


3.4.3

Path-indepelluent integrals: the J-integral and the C*-integral

The stress and strain fields in elastic materials (not


power-law

elastic

materials)

obey

necessarily

linear

or

certain conservation laws. In relation to

two-dimensional crack problems (Fig. 3.2), the most important of these laws can
be expressed as follows. The contour integral defined by
(3.22)

is independent of the path r as long as r starts at an arbitrary point


lower

crack

and no other singularities besides the crack tip


1968b).

In

on

the

surface and ends at an arbitrary point on the upper crack surface


are

enclosed

eq. (3.22), ds is the differential arc length of

ward normal unit vector on

r,

(Rice,

1968a,

r, n i is the out-

3.4

35

General Features of Fields

p
Xl

J=jF((J.E,U)ds

C=jF(o,iiJ)ds

1-----

W -----to!

Fig. 3.2. Integration path for the evaluation of J and C*

(3.23)
is the strain energy density, and
fields

around

the

crack

0 ..
lJ

and u. are the


1

stress

and

displacement

solving the continuum-mechanical equations for non-

linear elastic material. The value

~f

J can be measured at the load points of a

pair of specimens which have incrementally different crack lengths, a and a+da,
but are otherwise identical. If P1 denotes the load per
ness, J is given by

unit

specimen

thick-

(3.24)
where d is the load line deflection.
As a consequence of the elastic-viscous analogy, the
repeated

for

above

If in eqs. (3.22) to (3.24), ij' u i and d are replaced by


the

viscous

arguments

can

analogue

their

time
Some

From eq. (3.24) it is clear that J can be measured at

pOints

At

the

the

load

authors

of

same time, due to its path-independence, it also determines

the severity of the crack tip fields as will be shown explicitly for
materials

rates,

of the J-integral is obtained, which is often called C*

(Landes and Begley, 1976; Ohji et aI, 1976; Nikbin et aI, 1976).
use the notations J', j or J* instead of C*.

specimen.

be

nonlinear viscous material (not necessarily power-law material).

power-law

shortly. This property of J forms the basis of fracture mechanics in

rate-independent elastic-plastic materials. Analogously,

C*

plays

central

3.

36

role

in

the

fracture

mechanics

of

The Continuum-Mechanical Equations

creeping

bodies

and

it

will be used

extensively in the fracture mechanics part of this book. Alternative methods of


measuring J, or C*, will also be described there.
3.4.4

The HRR crack-tip fields in power-law materials

Rice and Rosengren (1968) and Hutchinson (1968a,b) (often jointly


as

HRR)

have

analyzed

the

referred

to

crack-tip fields in power-law elastic materials.

Here, the analogous power-law viscous results will be presented. The derivation
starts

from

shear,

eq.

Considering

the

equation

(3.15)

for

for the stress function [eq. (3.11) for anti-plane

plane

stress

and

eq.

(3.19)

plane

strain].

nonlinear viscous material means that the elastic terms must be

deleted in these equations. (In fact, they would be


near

for

negligible

asymptotically

the crack tip if they were taken into account). Now it can be verified by

insertion that the equation for the stress function has solutions of
(power

the

form

of r) (function of 6) where rand 6 are polar coordinates centered at

the crack tip as in Fig. 3.2. Working out these factorized solutions

leads

the following so-called HRR fields, which are valid near a crack tip as r

C*
n

B r

)1/(n+1)

aij (6)

dimensionless

factor

parameter
same and

0:

(3.25)

For power-law elastic materials, characterized by


the J-integral and the material
quantities in eq. (3.25) remain the

to

= Ban,
C* is replaced by
o

B is replaced by Bo' The other


are defined as follows: the

a..

is chosen such that the dimensionless function


(6)
_ _
1/2
lJ
is normalized, the maximum of a = (1.5a i' .a i'.)
being equal to 1. For Mode
e
J J
III (anti-plane shear) , closed-form solutions are available from Rice (1967a)
n

or, more explicitly, from Riedel (1978). The expressions for


are

complicated

In

~(n+1)/n.

and

will

not

For plane strain

be
and

reported
plane

here,

stress,

while

alJ..(6)

in Mode III

In takes the form

numerical

solutions

are

available from the papers of HRR, and Shih (1974, 1983). Table 3.1 and Fig. 3.3
contain a few examples.
The HRR-field was originally derived for Mode-I loading, but the Mode-II fields
have the same general form with different values for In and
and Fig. 3.3 show results which were taken from Shih (1974).

aij (6).

Table 3.1

3.4

37

General Features of Fields

Values of I n for:

n =

Plane strain,

Mode I

Plane stress,

Mode I

Plane strain,

Mode II

Plane stress,

Mode II

I In

1I(n+1)/n, Mode III

--- --- --- --- --- --2

5.94

5.51

5.02

4.71

4.19

3.77

==

4.77

10

...

4.54

3.72

3.46

3.14

------- --- - - - --4.22


2.98
2.49
3.86
3.41
3.17
--- --- ----- --- --0.61
0.84
0.74
1.12
0.78
0.97
--- --- --- --- - - - --0.80
1.47
1.26
1.01
0.95
1.09
--- --- --- --- --- --3.59

Table 3.1. Values of the factor In appearing in the HRR-fields.

......
ct>
......

,ti:::0

t 05

......
ct>

......

.ti:::-

-05
l5

......

120

180

8 in degrees __

n=3

t 05

ct>
......

60

plane f1, Mode II

- 1

60

120

60

120

180

8 in degrees - - 8 in degrees - Fig. 3.3. Dimensionless functions characterizing the angular distribution
of stress in the HRR-field.

3.

38

The Continuum-Mechanical Equations

It should also be noted that for n=1, the HRR-field reduces to


linear

elastic

crack-tip

field

with

the

well-known

the characteristic inverse square root

singularity:
KI

0iJ.

(3.26)

= - - f .. (8).

/211l'

IJ

Note that the stress fields for plane strain and plane stress are identical

in

the linear elastic case. The angular distribution of stress is expressed by the
function f ij (8), which is conventionally not normalized in the linear elastic
case, and which is given in Appendix B in closed analytic form for Mode I and
Mode II. For linear elastic material, the asymptotic crack-tip stress field
usually

expressed

in

terms

of

the

subscript indicates Mode-I loading.

For

stress

intensity

Mode-II

and

is

factor KI , where the


loading, the

Mode-III

respective stress intensity factors are KII and KIll' In linear elastic material, the stress intensity factor is related to the J-integral through

for plane strain; for plane stress, the factor (1-v2 ) is deleted.
The stress intensity factor can be measured, in principle, by measuring J using
eqs. (3.24) and (3.27). In
tabulated handbook solutions
Cartwright,

1974 r.

practice, however, KI is usually taken from


Rooke
(Tada, Paris and Irwin, 1973 ;
and

For purely dimens ional reasons the stress intensity factor

in an arbitrarily shaped body must be proportional to the

applied

stress

and

the square root of the crack length, a:


KI = g(a/W) a

;mao

(3.28)

The dimensionless function g(a/W), or a similarly defined quantity, is tabulated

in

the above references for a great number of specimen geometries. In most

cases the results are based on numerical solutions. For


plane-stress
uniform stress

crack
0,

of

length

is g(a/W) = 1.

plane-strain

or

2a in an infinite body subjected to a remotely

3.5

Numerical Techniques

39

3.5 Numerical Techniques in Solid Mechanics


Most of the results in this book are based on analytic solutions, but
questions

such

specific

as the determination of the geometrical function g(a/W) in the

preceding section must be dealt with numerically.


By far the most advanced numerical method in nonlinear solid mechanics

in

two

and three dimensions is the finite element method. There are numerous textbooks
available, for example those by Zienkiewicz
Wilson

(1976),

(1971),

Oden

(1972),

differential
of

Owen

(1984).

equations
the

discretization

body
is

In

the

finite

conferences

on

of

element

method,

the

partial

of solid mechanics are discretized based on a discretinto

preceded

finite
by

elements.

In

nonlinear

problems

the

a linearization of the equations, so that the

full nonlinear problem is solved iteratively


sequence

the

Methods in Fracture Mechanics edited by Owen and Luxmoore (1980) and

by Luxmoore and
ization

and

Segerlind (1976) and Owen and Hinton (1980). Special questions

of fracture mechanics are addressed in the proceedings of


Numerical

Bathe

or

incrementally

by

solving

linear problems. This procedure is generally carried out on large

computers. Several program systems are available commercially, some of them

at

low cost for non-profit institutions.


A method which may offer computational advantages in certain
boundary

finite element method, and its advantages are


with

problems

is

the

element method. However, the method is less well established than the
substantial

specimen. For such

linear
a

elastic

problem

Banthia

probably

contribution
and

confined

to

to

problems

the deformation of the

Mukherjee

(1985)

have

recently

compared the computational efficiency of the two methods.


Another method which is sometimes a convenient tool to solve nonlinear problems
is

the

Rayleigh-Ritz

method, in which the solid mechanics equations are cast

into a variational prinCiple and a functional is minimized with

respect

to

finite set of parameters. This method is best suited for simple geometries such
as spherical holes or cracks in infinite bodies (see,

e.g.

Budiansky

et

aI,

1982, and He and Hutchinson, 1981).


Finally, the finite difference methods, which compete with the

finite

element

method but seem to be less flexible, and the boundary collocation method, which
is confined to linear elastic problems, should be mentioned.

4 Stress-Directed Diffusion and Surface Diffusion

Stress-directed diffusion of

atoms

is

the

relevant

process

not

only

for

diffusional creep and grain boundary sliding, which have already been described
in Chapter 1, but also for various aspects of void nucleation and growth.

4.1 The Role of Vacancy Sources in Stress-Directed Diffusion


In crystalline solids, the microscopic mechanism of diffusion
migration

is

usually

of atomic vacancies. Interstitials are exceedingly scarce in thermal

equilibrium because of their high formation energy. This is true not


diffusion

the

in

the

grains,

but

also

only

for

diffusion along grain boundaries can be

explained most satisfactorily by a vacancy mechanism (Peterson, 1980).


Phenomenologically, the diffusion properties of a material are described by the
self-diffusion

Dv ' and for the grain


boundary, oD b . These coefficients are measurable accurately by observing how
quickly

coefficients

radioactive

tracer

for

the

grain

interior,

atoms deposited on the specimen surface penetrate

into the substrate (Peterson, 1980). Measured values are given in Appendix A.
There is a remarkable difference between diffusion in these tracer

experiments

and stress-directed diffusion. In tracer experiments, there is only an exchange


of tracer and host atoms with no net influx of

atoms

specimen.

however,

During

regions on grain
vacancies

stress-directed
boundaries

preferentially

diffusion,

which

leave

are

these

experiencing
regions.

This

depletion of vacancies and a corresponding drop of the

into

any

part

of

the

atoms diffuse towards


tensile

stress,

possibly

leads

diffusivity

of

while
to

atoms,

unless new vacancies are generated fast enough thermally or by other processes.
Now it is often claimed that grain boundaries contain effective vacancy sources
such

as

jogs

in

grain-boundary dislocations (Balluffi, 1980), but it is not

quantitatively clear under which

circumstances

these

sources

are

effective

4.1

41

The Role of Vacancy Sources

enough

compared

to

the

drain of vacancies by stress-directed diffusion. The

observed threshold stress for diffusional creep, for example, (Section 1.3)
a

strong

indication

that

is

thermal vacancy production can be rate limiting at

very low stresses. A related phenomenon

is

the

inhibition

of

sintering

by

high-melting particles (Ashby et aI, 1980).


The rate at which vacancies are produced
boundary

thermally

per

unit

area

of

grain

is given by the area denSity of sources No times the Arrhenius factor

times a frequency factor,

\Ill. e

-Q/kT

( 4.0

where \Ill. is the lattice vibration frequency (- 10 14 s-1) and Q is the


activation energy for vacancy generation. It will depend strongly on the type
of source which is available.
prohibitively

example,

in

the

perfect

lattice,

Q is

large under creep rupture conditions since a vacancy can only be

produced together with an


(typically

For

6 eV).

interstitial

which

has

high

formation

energy

At a very favorable vacancy source, the activation energy Q

might become as low as the format i orr energy of a vacancy (typically 1 to 1.5 eV
in the lattice and less in the grain boundary). For quantitative estimates, eq.
(4.1) is only of limited value because of the uncertainty in Q.
Therefore the procedure in this book will be to
assumption

that

start

from

the

conventional

grain boundaries are good vacancy sources and sinks providing

the vacancies for

cavity

nucleation

and

growth

easily.

Inhibited

vacancy

production and its effect on cavity growth will be considered in Chapter 13.

4.2 Stress-Directed Diffusion Along Grain Boundaries


We now turn to the formulation of the equations
diffusion

assuming

granted by

their

transport

along

which

govern

stress-directed

that the thermal equilibrium concentration of vacancies is


sufficiently

grain

rapid

boundaries

is

thermal

generation.

considered.

In

First,

fact,

diffusion often dominates compared to volume diffusion under

diffusive

grain boundary

conditions

which

are typical for cavity growth processes.


The driving force for a diffusive flux of atoms is the gradient of the chemical
potential.

The chemical potential of an atom in a grain boundary is lowered by

II.

112

Diffusion

a normal stress, an' acting on the boundary such that


(11.2)
where ~o is the potential of the atom in the absence of the stress, and n is
atomic volume. The contribution of the stress to the chemical potential is
given by the negative work, -ann, which the normal stress does when an atom is
added to a boundary. (It is easier to visualize this if a whole plane of atoms
is plated on the grain boundary).
According to the principles of thermodynamics, the diffusive

flux

density

in

the grain boundary, j. (in atoms per unit time and unit length) is equal to the
1

negative gradient of the chemical potential times the atomic mobility. Thus, in
the

absence

of other driving forces besides stress, the flux density is given

by

(11.3)

where the subscript i

denotes

the

vector

component

and

the

first

factor

represents the atomic mobility.


Mass conservation requires that the negative divergency of this current density
is equal to the number of atoms plated per unit time and unit area on the grain
boundary. Hence
un

the

normal

displacement

rate

of

the

adjacent

grains

-n Viji' or with eq. (11.3):


n6D b
kT

- - - Va

(11.11)

Equation (11.11) above is our basic relation for stress-directed diffusion


grain

is

boundaries,

which

relates

the displacement rate due to the plating of

atoms to the second derivative of the normal stress


Equation

acting

on

the

boundary.

(11.11) is particularly simple to solve if the grains can be considered

as being effectively rigid, so that


connection

along

Un

is constant. Examples will be treated in

with cavity nucleation and growth. If the grains must be considered

as being deformable, eq. (11.11) represents a boundary condition for the


ation of the grains (cf. Section 11.5).

deform-

4.3

Stress-Directed Diffusion Through the Grains

43

4.3 Stress-Directed Diffusion Through the Grains


Stress-directed diffusion through the volume is more

complicated

to

describe

than grain boundary diffusion. The thermal equilibrium concentration of vacancies in the immediate neighborhood of a stressed grain boundary (which is

again

considered as an ideal vacancy source) is


(4.5)

where Co is the concentration in the unstressed solid and an is the normal


component of stress acting on the respective grain boundary (Nabarro, 1948). In
most cases, the second form of eq. (4.5) can be used, in which the

exponential

function is expanded up to linear order of ann/kT.


In a given stress field, differently oriented grain boundaries experience diff-

an. The resulting difference in vacancy concentration


These fluxes
obey the usual diffusion equation (see, e.g., Crank, 1956)
erent

normal

stresses,

near these boundaries causes diffusive fluxes through the grain.

(4.6)
The diffusion coefficient

for

vacancies,

Dvac'

is

related

to

the

volume

self-diffusion coefficient by Dvac = Dv/(cvfc)' where the dimensionless correlation factor fc has values around 0.5, the precise value depending on the type
of

the

crystal

lattice.

The

flux of vacancies from and to grain boundaries

implies a displacement rate between adjacent grains:

(4.7)

IJn

Here, n i is the unit normal vector on the boundary pointing into the upper
grain, Vic~ is the concentration gradient above the boundary and Vic~ is the
gradient taken from below the boundary. (In
symmetric
situations
is
+

Vic v = -ViCv )
Equation (4.7) couples the diffusion problem with
grain

deformation.

It

the

sets

the

problem

of

can be considered as a boundary condition for the dis-

placement rate field in the grains. If the grains are


(4.7)

mechanical

boundary

conditions

effectively

rigid,

eq.

for the diffusion problem with un being

given by the rigid-body motion of the grains. The analysis

of

this

diffusion

4.

44

problem
acv/at

Diffusion

is still not simple (Herring, 1950), but in the steady state, i.e. for
0, the displacement rate must take the form

(4.8)

simply for dimensional consistency. Here L is

the

characteristic

geometrical

length over which diffusion takes place (for example the grain size in NabarroHerring creep), and av is a factor which depends on the geometrical details
the diffusion problem (for example the grain shape).

of

Similarly the displacement rate by grain boundary diffusion can be written as


(4.9)

The contributions to un by volume diffusion and by grain boundary diffusion are


additive owing to the linearity of the problem provided that the linearized
form of eq. (4.5) is applied. Although eq. (4.8) holds for rigid grains and
steady-state

diffusion

grain-bo~~dary

diffusion to hold

equation,

only,

we
in

assume
other

the
cases

additivity
as

well.

of

volume

Then

the

and
basic

eq. (4.4), for stress-direction grain boundary diffusion is modified

to include volume diffusion such that

(4.10)

Within this approximation volume diffusion can be treated as a (not necessarily


small)

correction

to boundary diffusion. In practical cases, the contribution

of volume diffusion is in fact often found to be small. Raj


give

and

Ashby

(1971 )

approximate

values for the ratio av/ab in the range 0.2 to 0.3 for a few
configurations including diffusional creep of a polycrystal as well as sliding
of a grain boundary containing hard particles. For a ferritic steel, Frost and
Ashby (1977) repor.t measured values of Dv/6Db ~ 2.10 8 m- 1 exp(-8,OOO KIT). This
implies that the contribution of volume diffusion to eq. (4.10) reaches that of
boundary diffusion not before the a-Y
L > 70 ~m. For T = 5300 C and L

transition

temperature

(723 0 C)

unless

10 ~m, volume diffusion contributes only 2 per

cent to the displacement rate in eq. (4.10).

4.

Surface Diffusion

45

4.4 Surface Diffusion


Surface diffusion determines the shape

of

growing

cavities

and,

sometimes,

affects the cavity growth rate markedly (Section 11.2).


The driving force for surface diffusion is the tendency to
surface

energy.

minimize

the

free

For isotropic materials this means that free surfaces tend to

assume spherical shapes. The chemical potential of an atom on a curved

surface

is given by
(4.11)

(Herring, 1951). Here,

is the chemical

~o

potential

of

an

atom

on

flat

surface, Ys is the specific surface energy and K1 and K2 are the principal
curvatures of the surface measured positive for the surface of a round cavity.
The

atomic

current

density (in atoms per unit arc length of free surface and

unit time) is given by the negative gradient of

times

the

atomic

mobility,

i.e. ,
( 4.12)

where 60 s is the surface diffusion coefficient having the physical dimension


m3 /sec. For conservation of mass the divergency of ji must equal the number of
atoms plated per unit time and unit area on the surface. Therefore the normal
displacement

rate

of

the

surface caused by the deposition of atoms from the

surface diffusion flux is given by

with Vn being

positive

(4.13)

for

growing

void.

Combining

the

preceding

two

equations gives the governing equation for surface diffusion,


!l 60 s Y
2
--k"';T"--'::'s V (K1

(4.14)

2)

fn analogy to the equation for boundary diffusion,

eq.

void growth problems will be worked out in Section 11.2.

(4.4).

Solutions

for

4.

46

Diffusion

4.5 Grain-Boundary Diffusion Combined with Power-Law Creep


The combined processes of grain-boundary diffusion and power-law creep will
considered

in

assumed to be nonlinear viscous as described by

Ban

be

problems involving cavity nucleation and growth. The grains are


eq. (3.6),

which

reduces

to

in uniaxial tension. Normal displacement rates at the grain boundaries

occur by diffusive transport according to eq. (4.4).


In such a model there are no time-dependent transients

for

constant

external

load since both, the strain rate in the grains and the displacement rate across
the boundaries are functions of the current stress only.

However,

the

Simple

scaling properties of the stress fields in power-law materials are not preserved in the presence of grain boundary

diffusion.

This

is

because

creep

and

diffusion have different stress dependences, so that at high stresses power-law


creep determines the stress
transport

becomes

distribution,

while

at

low

stresses

diffusive

predominant. The interaction between the two mechanisms can

be expressed in terms of a characteristic length,

~,

to be introduced next.

The governing equations, eqs. (.1.6) and (4.4), can be written in the dimensionless form
(3/2) l:n-1

(4.15)

l:!.
IJ

(4.16)
with the characteristic diffusion length,

kT

Un

e:i/(Ba )

(4.18)

un/(Ban L)

(4.19)

l: ..
IJ

is

(4.17)

E
n

EIJ
..

and the normalized quantities:

{l6D b a )1/3

Here,

~,

characteristic

a . ./a
IJ

(4.20)

x/L.

(4.21)

geometrical

length

of

the

problem

under

4.5

47

Grain-Boundary Diffusion and Power-Law Creep

consideration

(for

example,

grain

size

average stress acting on that size

scale

strain

the

rate.

parameters,

The

~/L

equations

for

or

cavity

and

size), a represents the

Ban

is

the

corresponding

normalized quantities contain only two

and n, the latter being less important. Therefore the character

of the stress and strain distribution is determined primarily by


If the diffusive length

~/L.

defined by eq. (4.17) is large compared to the length

L, the displacement rate u by diffusion is large and creep deformation can be


n
neglected. The grains behave effectively rigidly. For ~/L 1, on the other
hand, the displacement rate by diffusion goes to zero, and the deformation
field is determined by creep of the
points

grains

except

possibly

near

particular

where V2 an is large, for example, at a triple junction of sliding grain

boundaries. At such

singularities,

grain

boundary

diffusion

smoothens

out

stress concentrations.
Numerical values for
Inspection
stresses,

of
~

a = 6 MPa),

their

have

been

tabulated

by

Needleman

has usually values of many micrometers


so

and

Rice

(1980).

table shows that at half the melting temperature and low

that

diffusion

is

the

dominant

34

(~=

~m

for

a-iron

lengths as large as typical grain sizes. For ten times greater stresses at
same

temperature,

drops to the order of a micrometer

(~ =

0.4

~m

so that diffusion is important for stress relaxation at second-phase


which

have

typically

if

deformation mechanism over


the

for a-iron)
particles

such a size or for cavity growth. Higher temperature as

well as higher stress reduce the effect of grain boundary diffusion compared to
dislocation

creep.

The

temperature

dependence

occurs

since the activation

energy for grain boundary diffusion is usually smaller than that for
creep.

power-law

Part II
Creep Cavities

5 Introduction to Part II

It has long been known that at elevated temperatures metals and alloys fracture
with

relatively

low

ductility when loaded to moderate stress levels for pro-

longed times. The strain to failure drops from typically 50% to 10% or less
the

test

duration increases from a few minutes to a few months or years. This

loss in ductility is usually caused by


shows

as

var i ous

intergranular

cavitation .

Figure

5.1

stages of creep cavitation. Single isolated cavities can be de-

tected in the secondary stage of the creep curve, sometimes even in the primary
stage . Theoretically, the cavity size at nucleation should range from less than
10 to 100 nm (Section 6.2). Figure 5.1a shows a later stage, in which
start

cavities

to coalesce on grain boundary facets to form grain-size microcracks . The

stage of well-developed microcracks -is shown in Fig. 5.1b.

(These

two

micro-

graphs were taken from areas near a macroscopic crack, but in uniaxial tension,
creep damage exhibits a similar appearance). Finally, coalescence of the cavities leads to a dimpled, intergranular fracture surface as shown in Fig . 5.1c.

Fig. 5.1 . a) and b) Grain boundary cavitation in 1 12Cr-112Mo-1/4V steel


tested at 540 0 c (courtesy V. Detampel) . b) Intergranular fracture surface
of a 1%CrMoV steel tested at 550 0 C (courtesy D. Horstmann) .

5.

52

Introduction to Part II

Qualitative observations on cavity nucleation and growth are summarized in

the

present chapter, while the remainder of Part II is devoted to detailed analyses


of models and comparisons with experiments. Chapters 6 to 10 refer to the nucleation stage describing, in this order, the basic theories, the role of stress
concentrations, the effects of impurity segregation to grain boundaries and gas
bubble

formation. These investigations aid the understanding of the nucleation

process, but do not have a predictive capability in a quantitative


chapters

on

nucleation

are

sense.

cavity growth and coalescence mechanisms. The calculation of rupture


in

The

followed by Chapters 11 to 15 describing various


lifetimes

engineering alloys is complicated by the fact that cavity nucleation occurs

continuously. Therefore a theory for the evolution of the cavity size distribution function is developed in Chapter 16. Section 16.3 contains a detailed comparison of calculated and measured rupture lifetimes. Chapter 17 summarizes the
results

of

the

preceding

chapters. Finally, a chapter on cavitation failure

under cyclic loading conditions concludes Part II.

5.1 Experimental Techniques


Several experimental techniques have been
cavitation.

Accurate

density

employed

measurements

allow

total cavity volume (e.g., Hanson and Wheeler, 1931;


1961;

Johannesson

to

study

Boettner

total

cavity

boundary

and

Robertson,

and Tholen, 1969; Dunlop, Twigg and Taplin, 1978; Hanna and

Greenwood, 1982). Small-angle neutron scattering is another


the

grain

for a determination of the

volume

and,

under

favorable

means

to

measure

conditions, the cavity size

distribution (e.g., Saegusa et aI, 1978; Nilsson and Roth, 1981; Page, Weertman
and

Roth,

1982;

Yoo et aI, 1982). The method is sensitive to small cavities,

but the signal arising from precipitates, which usually also grow during
of

structural

materials,

cannot be resolved. Similar problems are expected to obstruct


of

other

the

to

distinguish

Further, high voltage electron microscopy has been used to study the
and

application

non-destructive testing techniques, like ultrasonic or micromagnetic

methods, whose response to cavities is difficult


response to changes in the precipitate structure.

size

creep

may be dominant, so that the scattering by cavities

from

the

location,

shape of very small cavities (e.g., Fleck, Taplin and Beevers, 1975;

Svensson and Dunlop, 1979) or of gas bubbles (Braski et

aI,

1979).

Also

the

scanning electron microscope is a widely used tool to observe cavities and, together with energy dispersion analysis, to identify particles at which cavities

5.1

53

Experimental Techniques

may have nucleated. Chen and Argon (1981a) combine scanning electron microscopy
with a two stage creep technique; by a
strain

after

polishing

small

additional

increment

by

ion

beam

small

the

polished

cavities

etching of the polished sections. A method which is now

well established for the field inspection of power plants is to


from

creep

the specimens, they open up small cavities, which may

have been closed by polishing. Needham and Gladman (1980) make


visible

of

take

replicas

and etched surface of critical parts such as welds or pipe

bends and to inspect the replicas for creep damage in the optical

or

scanning

electron microscope (Neubauer, 1981; Neubauer and Arens-Fischer, 1982).


Another class of methods consists in interrupting creep tests at various stages
and

fracturing the crept specimens in a brittle manner along the grain bounda-

ries, so that the cavities are exposed and can be observed. Brittle intergranular

fracture

can be achieved in magnesium alloys, ferritic steels and even in

certain austenitic steels by impact loading at


Wingrove,

1967;

Cane

and

Greenwood,

low

temperatures

(Taplin

and

1975; Pilkington, Miller and Worswick,

1981; Chen and Argon, 1981 a). Other mater ials can be made to cleave along their
grain

boundaries

by

liquid

metal

embrittlement (Reiley, 1981) or by stress

corrosion cracking (Elliott and Wilshire, 1979).

5.2 Materials which Exhibit Intergranular Cavitation


The first systematic study on cavitation, which is known to the author, is that
of

Hanson

and Wheeler (1931) on aluminum alloys, which was probably motivated

by the need of understanding the long-term performance of aircraft engines.


Large engineering structures such as electricity-generating plants are
built

usually

of the relatively cheap ferritic steels. Around the turn of the century,

steam engines operated at steam temperatures of 200 0 C. The turbines of the late
1920's

used

superheated

steam of around 400 0 C, temperatures which were still

achievable with the conventional carbon steels. Pushing the temperature

beyond

400 0 C required new concepts in alloy design. This led to the development of the
bainitic low-alloy chromium-molybdenum steels in the 1930's, in particular
applications

based on the same ideas (Khare, 1983). The maximum operating


the

low-alloy

for

in power plants. Many of today's creep-resistant steels are still


steels now range from 530 to

565 0

temperatures

for

C. Steels with higher chromium

content, for example the 12%CrMoV steels with tempered martensite

microstruct-

ures, retain a sufficient creep resistance (and oxidation resistance) up to 600

5.

54

Introduction to Part II

to 650 0 C. Under service conditions, many of these steels develop


creep

cavities

interest in cavitation in creep-resistant steels. Examples are


Johannesson

intergranular

which can lead to failures. Hence, there has been a continuous

and

Tholen

(1969),

the

papers

of

Tipler and Hopkins (1976), Cane (1976, 1979,

1981), Lonsdale and Flewitt (1981), Sklenicka et al (1981), Needham and Gladman
(1982,

1984),

and

Needham

(1983). In parallel, grain boundary cavitation in

pure a-iron has been studied by Taplin and Wingrove (1967), Cane and
(1975)

Greenwood

and Player and Brinson (1975). For an evaluation of creep rupture tests

on ferrous alloys see Sections 12.1.2, 16.3.1 and 16.3.2.


Austenitic steels can be applied at higher temperatures than
up

to

800 0

some

c,

ferritic

for example in the chemical industries. Also the pressure

vessels of fast breeder reactors are made of austenitic steels. Some


not

cavitate

steels,
heats

readily (Abou Zahra and Schroder, 1982) while others do (Dunlop,

Twigg and Taplin, 1978; Needham and Gladman, 1980; Chen and Argon, 1981;
deman

et

do

Swin-

aI, 1981; Gandhi and Raj, 1982). The pronounced heat-to-heat variat-

ions are ascribed primarily to differences in the carbide morphology, which, in


turn, is affected by the presence of trace elements (Swindeman et aI, 1983).
Another impetus to study intergranular cavitation came from failures of nuclear
reactor fuel rods which, in the gas-cooled Magnox reactors, are made from magnesium alloys. Greenwood, Miller and Suiter (1954) studied cavitation in magnesium

(and

also

in

copper and a-brass). They were among the fjrst who clearly

established that the loss of ductility at elevated temperatures and slow strain
rates was a consequence of grain boundary cavitation. Subsequent work on magnesium and its alloys was reviewed by Perry (1974).
More recently, the development of nickel-base superalloys added
of

commercial

conditions.

materials

Superalloys

applications

and

for

which
are

jet

may

used

another

class

fail by cavitation under typical service


in

gas

turbines,

both

for

stationary

engines, and also for pipes and sheet materials in

high temperature reactors. Perry (1974) has thoroughly reviewed published


on

cavitation

of

superalloys.

work

More recent papers are, for example, those by

Dyson and McLean (1977) and by Shiozawa and Weertman (1983).

Tests

on

nickel

and its alloys are evaluated in Sections 12.4.2, 12.4.3 and 16.3.3.
Copper and its alloys have been convenient materials to study
cavitation

the

physics

of

starting with the work of Jenkins et al (1943) on a Cu-Ni-Si alloy.

Other measurements on copper and brass will be described

in

Sections

11.1.2,

5.2

Materials Exhibiting Cavitation

55

11.2.7 and 12.4.3. Observations of cavitation in titanium alloys are summarized


by Krishnamohanrao et al (1986). Among the less common metals,
1966),

silver

(Price,

tungsten (Stiegler et al, 1967) and zirconium (Snowden et aI, 1981) are

known to cavitate under appropriate conditions, whereas lead does

not,

unless

specific impurities are present (Greenwood, 1978).


Other classes of materials some of which fail by

inter granular

cavitation

at

high temperatures are the ordered intermetallic compounds like C0 3Ti (Takasugi
and Izumi, 1985) and the structural ceramics (Davidge, 1979; Porter et aI,
1981; Hsueh and Evans, 1981; Evans, 1982).
Although cavitation is a widespread failure mechanism, it is by no

means

uni-

versal. Pure aluminum (Chang and Grant, 1953), lead and titanium do not exhibit
cavitation nor do some of the low-alloy creep-resistant steels nor some austenitic

steels.

nickel-base superalloys Nimonic 105, Nimonic 115 and IN 100


fail by the growth of oxidized surface cracks at temperatures around 900 0C
(Dennison

et

The

aI, 1978) and changes in microstructure during creep play an im-

portant role. However, Nimonic 105 cavitates when pre-strained at room


ature

(Burt,

Elliott

and

Wilshi~,

temper-

1981). The occurrence of cavitation also

depends on details of the heat treatment (as in ferritic and austenitic steels)
and

on

minor

changes

in

alloy composition. For example, small additions of

zirconium to a Ni-20%Cr alloy can suppress cavitation (Schneibel et

aI,

1983)

to

inter-

in favor of ductile transgranular failure.

5.3 Diffusion as the General Cause for Intergranular Cavitation


The general cause for the transition from voidage within the grains

granular cavitation is that atomic vacancies become mobile at elevated temperatures. At typical service temperatures
along

of

creep-resistant

alloys,

diffusion

grain boundary predominates. Vacancies diffusing in grain boundaries can

aggregate to form cavity nuclei. Further, grain boundaries are considered as


effective vacancy sources (Balluffi, 1980) so that enough vacancies are supplied to feed cavity growth. Finally, the mobility of atoms leads to the possibility

of grain boundary sliding. Associated with sliding are stress concentra-

tions at obstacles like second-phase particles or triple grain junctions.


may

enhance

cavity

nucleation.

At

This

even higher temperatures, however, rapid

dIffusion enables grain boundary migration and recrystallization to take place,


which terminates the range in which cavitation prevails.

5.

56

Introduction to Part II

5.4 The Role of Grain Bonndary Sliding


5.4.1

Experiments on bicrystals

The importance of grain boundary sliding for cavity nucleation in a pure


has

been

convincingly

demonstrated

by

metal

Chen and Machlin (1956, 1957) and by

Intrater and Machlin (1959, 1959/60). They performed

creep

rupture

tests

on

copper bicrystals and found that no cavitation takes place when the grain boundary is subjected to normal tension only, but that cavitation
if

becomes

profuse

a shear force is applied to the grain boundary prior to tensile loading. On

the other hand, Raj (1975) obtained cavitation of bicrystal grain boundaries in
copper

under tensile loading alone. In this case, nucleation occurred at oxide

particles. Possibly no such particles were present in the


Chen

material

tested

by

and Machlin. Gandhi and Raj (1982) observed cavitation in stainless steel

bicrystals under normal tension. Thus at least in pure metals,


sliding

seems

to

play

grain

boundary

an important role in cavity nucleation, but it is not

always a necessary prerequisite.


5.4.2

The orientation of cavitating boundaries in polycrystals

That grain boundary sliding can hardly be the overriding factor

in

cavitation

is also demonstrated by the common observation that in polycrystals grain boundaries which are oriented normal to the applied tensile stress,
slide

the

least,

and

therefore

cavitate preferentially. This was shown by Davies, Williams

and Wilshire (1968), by Johannesson

and

Tholen

(1969),

by

Chen

and

Argon

(1981a) and by Chen and Weertman (1984). In some cases, the orientation distribution of cavitated boundaries depends on the testing
(1973),

for

conditions.

Rao

et

al

example, observe preferred cavitation of inclined boundaries only

in fine-grained material of an austenitic steel. Further, cavitation on inclined

boundaries

seems

to predominate at high strain rates, while at low strain

rates normal boundaries cavitate (Wingrove and Taplin, 1969a, on iron;

Gittins

and Williams, 1967, on copper). However, this dependence on strain rate was not
observed by Davies, Williams and Wilshire (1968) for pure copper. In summary,
cavitation

appears to be more pronounced on normal boundaries than on inclined

boundaries. Even if this were due


preferred

to

faster

cavity

growth

rather

than

to

nucleation, cavitation of normal boundaries is probably the relevant

process that leads to fracture.

5.5

51

Cavity Nucleation Sites

5.5 Cavity Nucleation Sites


It is very unlikely that cavity nucleation can occur at an

arbitrary

location

in a grain boundary. In pure metals, cavities are often initiated at the intersection of a slip band with a grain boundary or

at

ledges

in

the

boundary.

Commercial materials usually contain second-phase particles on grain boundaries


in order to prevent or to minimIze grain boundary sliding. These particles

are

prone to nucleate cavities, and in this respect they effect the mechanical properties of materials adversely.
5.5.1

Slip bands

Dislocation motion during plastic deformation is concentrated in more


pronounced

slip

bands

within

the

or

less

grains. If such a slip band impinges on a

grain boundary, it causes a stress concentration which might nucleate a cavity.


Indeed,

Watanabe

and Davies (1918) present evidence that cavities are located

where slip bands intersect the grain boundary in a copper bicrystal. Similarly,
Nieh and Nix (1980b) find that in copper polycrystals the cavity spacing is
equal to the slip band spacing. Both-vary in proportion to 1/0 in the stress
range 10 to 100 MPa at 500 oC. Sklenicka et al (1981a,b), on the other hand,
deny such a correlation, although their

testing

conditions

were

similar

to

those of Nieh and Nix.


In nickel-base superalloys the slip character is usually
slip

coarse

intense

bands. Cavities are often found at places where a slip band hits a grain-

boundary carbide, and the cavity spacing is approximately


band

with

spacing

equal

to

the

slip

(Shiozawa and Weertman, 1983). An analysis of the stress concen-

tration at a slip band will be described in Section 1.1.


Instead of nucleating a cavity, the stress concentration at
initiate

slip

in

the

neighboring

slip

band

can

grain. This leads to a ledge in the grain

boundary as indicated in Fig. 5.2. Subsequently, grain

boundary

sliding

con-

centrates stress on such a ledge.


5.5.2

Grain-boundary ledges

In the early papers of Chen and Machlin (1956, 1951) and


idea

of

cavity

initiation

Gifkins

(1956),

the

at ledges in grain boundaries was favored. McLean

(1963) supported the idea on theoretical grounds, but Harris (1965) showed that

5.

58

Introduction to Part II

aJ

compressive

cavities

ledge
tensile ledge
compressive ledge

Fig. 5.2. Cavity nucleation at grown-in and slip-induced ledges

and at a triple junction.

McLean's conclusion would have been different, had he used a correct value

for

the diffusion coefficient.


The ledges can be grown-in features of the boundary, or they
slip

can

result

from

in the adjacent grains, which intersects the boundary (Fig. 5.2). Depend-

ing on the orientation of the ledge with respect to the sliding direction,
ledge

the

experiences tensile or compressive stresses, and only the tensile ledges

are expected to initiate cavities. Incidentally, slip-induced ledges tend to be


of

the compressiye type, for geometrical reasons, which are apparent from Fig.

5.2b. Application of an external compressive load,

however,

leads

to

ledges

which experience tensile stresses in a subsequent tensile creep test. In agreement with this expectation, Davies, Williams and Wilshire (1968) found a higher
incidence

of cavitation on inclined, i.e. sliding, boundaries if the specimens

were compressed before the tensile creep test. Watanabe and Davies
port

(1978)

re-

cavity formation at slip-induced compressive ledges in copper bicrystals,

but these cavities may have been nucleated already by the slip band rather than
by grain boundary sliding focussing stress on the ledge.
Further experimental evidence for cavity nucleation at ledges was presented
Watanabe

(1983)

for an iron/tin alloy. In this case, rather large ledges were

produced by grain boundary migration. The defects generated by


sliding

by

grain

boundary

at these ledges were microcracks rather than typical cavities. Similar

5.5

59

Cavity Nucleation Sites

results were obtained by Presland and Hutchinson (1963/4) on magnesium in which


a

high

percentage

of the cavities initiated on cusps in grain boundaries re-

sulting from intersection of a sub-boundary. In summary, cavity


ledges

seems

probably less important, since


Stress

nucleation

at

to occur in pure metals. In engineering alloy, this mechanism is


there

nucleation

at

particles

predominates.

concentrations at ledges during grain boundary sliding will be analysed

in Section 7.4.
5.5.3

Triple grain junctions

If grain boundaries slide easily, as they do in pure metals, they focus


on

triple

grain

stress

junctions (see Fig.5.2). These stress concentrations will be

described in Section 7.2. In fact, cavities at triple junctions have

been

ob-

served frequently. An early study of triple-point cracking is that of Chang and


Grant (1956). Cavities at triple junctions sometimes have

the

form

of

wedge

cracks (Section 5.6) as distinct from the usual, rounded cavity shapes.
5.5.4

Grain boundary particles

In commercial materials, the most common cavity nucleation


phase

sites

are

second-

particles in grain boundaries. Numerous investigations show that in low-

alloy steels, cavities are often associated with carbide particles (Johannesson
and

1969;

Thole~,

Cane,

1976;

Miller

and

Pilkington,

1978; Lonsdale and

Flewitt, 1979). In a detailed study, Needham (1983) observes cavities at grainboundary carbides, preferably of the type Mo 2C, in two low-alloy bainitic Cr-Mo
steels. In vanadium-bearing steels, VC-particles are known to nucleate cavities.

Besides the carbides, there are other inclusions at which cavities can be

formed. Gooch (1982) reports cavity nucleation at sulfide, silicate


inclusions

and

oxide

in a 12%Cr steel, where carbides act as secondary nucleation sites.

If MnS-particles are precipitated to grain boundaries, which may happen in


heat-affected

zone

of

weld,

they

the

nucleate cavities readily (Cane, 1981;

Middleton, 1981; Cane and Middleton, 1981; Needham, 1983). This may give rise
to stress relief cracking (Section 8.1.3) and to poor creep rupture properties.
In a-iron, cavity formation was found to occur at sulfides, but not at carbides
(George,

1985). Oxides decohere from the iron matrix only if sulfur is present

in the particle/matrix interface. Older papers simply state that oxides act
cavity

nucleation

sites

as

in a-iron containing 200 ppm oxygen (Taplin and Win-

grove, 1967) or 2.8 ppm oxygen (Player and Brinson, 1975) with no reference

to

5.

60

Introduction to Part II

a possible effect of sulfur.


In austenitic steels cavities nucleate at grain-boundary carbides (Argon,
and

Lau,

1981;

Swindeman,

Farrell

Chen

and Yoo, 1981). The same is true for the

formation of helium bubbles during creep of irradiated material (Braski et

aI,

1979; see also Section 9.3). After long-time exposure to high temperature, when
sigma phase has been prectpitated, cracks sometimes nucleate at the sigma-austenite

interface (Williams, Harries and Furnival, 1972; Lai and Wickens, 1979).

Cavity formation at sigma-phase particles was shown


(1979)

to

be

responsible

by

Chasterl

and

Flewitt

for the premature failure of a Type 316 stainless-

steel superheater tube in a power station. Finally Rao, Rao and

Pandey

(1973)

observed cracking along the interface between the matrix and Cr 2N-precipitates
within the grains of a manganese-nitrogen austenitic steel.
In nickel-base superalloys, coarse M23 c6 -carbides have been identified as cavity nucleation sites (Law and Blackburn, 1980), often in conjunction with a slip
band impinging on the particle from the grain interior (Shiozawa and

Weertman,

1983).
In copper, there are often small oxide particles (Si0 2 or CuO) which nucleate
cavities (Raj, 1975; Pavinich and Raj, 1977; Chen and Weertman, 1984). By highvoltage electron microscopy, Fleck, Taplin and Beevers

(1975)

observe

cavity

nucleation in a copper alloy exclusively at (unspecified) particles.


Harris (1965) shows that in commercially pure magnesium and
alloy

in

the

magnesium

Magnox AL80, cavities nucleate at particles, the nature of which was not

identified.
There are several possible reasons why grain-boundary particles act as preferred cavity nucleations sites. First, because they resist grain boundary sliding,
they experience stress concentrations, which will be calculated in Section 7.3.
Second, they may be weakly bonded to the matrix. While this is conceivable for
sulfide and oxide particles, carbides are generally assumed to be well bonded,
since they do not easily decohere during room temperature straining. Therefore,
a third possibility may be important, namely, that cavity nucleation
icle/matrix

at

part-

interfaces may be relatively easy via vacancy condensation even if

the interface has a high strength as will be shown in Section 6.2.

5.6

61

Wedge Cracks

5.6 Wedge Cracks


Intergranular cracks often emanate
shown

in

from

grain-boundary

triple

junctions

as

Fig. 5.3. Their wedge-type shape suggests that they are opened up by

the sliding of inclined grain boundaries. These wedge cracks are preferentially
observed

after creep testing at high stresses. In nickel-base superalloys, for

example, McLean (1956/7) finds that wedge cracks predominate at stresses above
150 MPa in the temperature range 700 0 C to 950 oC. Lai and Wickens (1979) observe
wedge cracking in austenitic steel above 200 to 250 MPa at temperatures between
500 0 C and

675 0 C.

The

transition

from the occurrence of rounded cavities to

wedge cracks has been called Stroh/McLean transition.


sidered

Stroh

(1957)

has

con-

brittle cracking in connection with cleavage fracture, and wedge crack

formation is sometimes envisaged as a brittle decohesion of the grains.


In many cases, however, wedge cracks are obviously a result of cavity
ence

rather

than

coalesc-

one of brittle cracking. Sometimes their origin is apparent

from their serrated edges (Chen, 1956; Chen and Argon, 1981a).

Experimentally,

the distinction between brittle decohesion and cavity coalescence is not simple
as exemplified by the study of Mullendore and Grant (1961) (quoted by Grant,
1971) on an AI-5%Mg alloy. After a certain preparation of the polished sections
of crept specimens, wedge cracks were observed which appeared

to

have

smooth

edges. A further light etching, however, revealed numerous small rounded cavities on boundaries which had not yet cracked. Most probably the
have

linked up to form wedge cracks,

cavities

had the test been continued.

a)

Fig. 5.3. Wedge cracks having smooth or serrated edges.

would

A possible

5.

62

example for brittle decohesion is the brittle mode of


discussed
by

stress

relief

cracking

recently by McMahon (1984) (Section 8.1.3). Brittle cracking is cer-

tainly sensitive to impurity segregation to grain


shown

Introduction to Part II

boundaries

as

was

already

McLean (1956/7) for nickel-base superalloys. An effect of impurities

is plausible since it is known that elements like phosphorus or

sulfur

reduce

the cohesive strength of grain boundaries (Section 8.2.4).


If cavity coalescence is the prevailing mechanism for wedge-crack formation,
separate

discussion of wedge cracks is not necessary. The kinetics of the pro-

cess would then be determined by the nucleation and growth of cavities, and the
difference

to

homogeneous

cavitation on the boundary facet would probably be

inSignificant. Brittle decohesion of the grain boundaries, however,

will

obey

different kinetic laws. Attempts to model wedge-crack growth have been based on
Stroh's (1957) and Cottrell's (1958) elastic analysis of a wedge
and

Williams,

1970,

1971;

Burton

stable, time-dependent crack growth. However, this is


artefact,

for

crack

(Heald

and Heald, 1975). Such models indeed give


probably

theoretical

Smith (1966) has shown that a slip-induced micro-crack together

with the nucleating slip band is not a mechanically stable configuration, i.e.,
the

crack would propagate unstably once it propagates at all. Smith's conclus-

ion is based on the analysis of a configuration


boundary

and

the

microcrack

plane

which

the

sliding

grain

are coplanar. Riedel (1976) generalized Smith's

analYSis by examining a triple junction with an


crack

in

angle

of

135

between

the

and the sliding boundaries. This model gives a very limited range

of crack stablility which, however, cannot account for

the

stable

growth

of

wedge cracks over appreciable fractions of a grain facet. Thus the relevance of
existing models (most of which have
fracture)

is

questionable

in

originally

relation

to

been

developed

wedge-crack

for

growth

cleavage

under creep

conditions. Therefore, they will not be presented here.

5.7 Some Observations on the Kinetics of Cavity Nucleation


5.7.1

The observed nucleation kinetics

Numerous investigations agree on that cavity nucleation generally starts


and

continues

early

over an appreciable fraction of, or over the whole, creep life.

Quantitative measurements of the cavity nucleation rate have been made by counting

the

number

densities of observable cavities at various fractions of the

lifetime. These counts are usually made at cavity sizes of 0.5 to 1

~m,

whereas

5.7

Observations on

cavity

nuclei

Nu~leation

63

Kinetics

are much smaller, say, 20 nm. Therefore the apparent nucleation

kinetics may be distorted by cavity growth. Dyson (1983), however,

points

out

that small cavities tend to grow rapidly and therefore the counting of cavities
having a size of 0.5

reflects the kinetics of cavity nucleation sufficiently

accurately in many practical cases.


As Dyson (1983) further noticed, a common result of many
is

that

the

number

of

cavities,

experimental

studies

N, per unit grain boundary area increases

approximately in proportion to creep strain with a

factor

of

proportionality

which, to a first approximation, is independent of stress. This observation can


be expressed as
a' e:

or

(5.1)

J*

a'

E,

where J* is the cavity nucleation rate per unit grain boundary area, a'
empirical

factor

is

an

of proportionality having the physical dimension [m- 2 ] and

is strain rate. By agreement, the number of cavities is referred to those grain


boundaries

only

whose

normals

deviate by at most 30 0 from the tensile axis,

because these boundaries cavitate preferentially.


Such a linear dependence of cavity density
slight

modifications)

by

Greenwood

on

(1969)

strain

has

been

found

(with

and by Fleck, Taplin and Beevers

(1975) for copper, and by several authors for steels

and

for

superalloys.

selection of measured data for the material parameter a' is shown in Table 5.1,
the values for which are partly taken from a diagram compiled by Dyson
The

table

material.

shows
There

1/2Cr-1/2Mo-1/4V

that
is
and

(1983).

the number of cavities varies greatly from material to


substantial
1Cr-1Mo-1/4V.

difference

even

between

the

steels

Further, the cavity density in overheated

material (austenized at 1300 0 C to simulate the heat affected zone of a weld) is


much

higher

than in material which was austenized at lower temperatures. This

is probably so because the overheated material contains finely dispersed sulfides

on

grain

boundaries,

at which cavities nucleate easily. By reducing the

content in trace impurities (Cu, As, Sb, S, 0 and N), Tipler and Hopkins (1976)
optained very low cavity densities as shown by the data for high purity steels.
Nickel-based alloys exhibit a similar sensitivity to trace impurities,
observed, for example, by Thomas and Gibbons (1984).

as

was

5.

64

Introduction to Part II

Material

Source

a' in m- 2

21/~Cr-1Mo

steel
(austenized at' 1300 0 C)

Cane (1979)
Dyson (1983)

4.10 12

2.7

lCr-1Mo-1/4V steel
(austenized at 1300 0 C)

Tipler and
Hopkins (1976)

1.4.1012

0.2-1

1 12Cr-l 12Mo-1/4V steel

Tipler and
Hopkins (1976)

1.5.1012

2.2

lCr-1Mo-1/4V steel

Tipler and
Hopkins (1976)

4.10 10

high purity steels

Tipler and
Hopkins (1976)

Type 347 stainless


steel

Dyson (1983)
Needham and
Gladman (1980)

8.10 11

NimoniQ 80A (in tension)

Dyson (1983)

4.10 10

25

Ef

(2-10) .10 9

in %

6-19

Table 5.1. Cavity nucleation rate (per unit strain) and strain-to-fracture.

Not surprisingly, the materials with the lowest cavity


highest

ductilities.

For

the

between the strain to failure,


E
f

materials
Ef

densities

exhibit

the

included in Table 5.1, the relation

, and a' can be approximated by

1250 a,-0.4

(5.2)

although the scatter around this relation is considerable.


Needham (1983), by comparing smooth and notched specimens, examined the
of

the

effect

stress state on the nUCleation rate in two Cr-Mo steels. He found that

it is the principal tensile stress, aI'

which

controls

the

nucleation

rate

according to N O~ with m = 4 to 7. The von Mises equivalent stress, 0e' is


usually less important except at high stresses in 21/~Cr-1Mo steel where
Needham

finds

oi

a!. The predominance of the principal tensile stress was

also found by Dyson and McLean (1977) who carried out tests on Nimonic

80A

in

tension and torsion.


The proportionality of the cavity density to strain sometimes holds until

very

5.7

Observations on Nucleation Kinetics

close

65

to final failure (see, e.g., Dyson and McLean, 1972). In other cases, in

particular in the overheated Cr-Mo steels studied by Cane (1979, 1981), by Cane
and

Middleton

(1981) and by Middleton (1981), the cavity density saturates at

some fraction of the lifetime with a saturation value, which increases with
stress as Nsat 0 1 3 Apparently, the sulfide particles which are the primary
nucleation sites in the overheated condition tend to nucleate cavities early in
creep

life.

In

Type

304

stainless

steel, the cavity density also tends to

saturate to a stress-dependent maximum value (Chen and Argon, 1981a).


Another deviation from the nucleat10n kinetics given in eq. (5.1) is an
ional

stress

for a-iron and by Dyson and McLean (1972) for Nimonic


may

be

addit-

dependence of a' which was reported by Cane and Greenwood (1975)

critical

cavity

80A.

Especially,

there

nucleation stress, which will be discussed in the

following subsection. Further, a small offset strain is occasionally found such


that N = a'(-)
(Greenwood, 1969). Finally, the experiments of Shiozawa and
o
Weertmann (1983) on the nickel-base superalloy astroloy revealed a slightly
different

kinetic behavior than eq. (5.1) the difference being possibly due to

inaccuracies in the measurements. They found that the inverse spacings of cavities,

A,

and of coarse slip bands both increase in proportion to strain. This

means that the cavity density, which is N = 1/A 2 , increases


strain

squared.

in

proportion

to

Evans (1984) quotes a few other papers containing information

on the increase in cavity number with strain. Rationalizations of the

observed

behavior will be given in the following chapters, especially in Section 6.3.3.


5.7.2

Is there a critical stress for cavity nucleation?

In copper, Fleck, Taplin and Beevers (1975) observed cavities at stresses down
to 8 MPa, but none below 8 MPa. In a-iron at 700 oC, the lowest stress applied
by Cane and Greenwood (1975) was 9 MPa
Dyson

and

McLean

and

they

still obtained cavitation.


0
(1972) observed cavities at 750 C in Nimonic 80A at applied

stresses between 92 and 385 MPa. It is possible, however, that cavitation would
have

occurred at lower stress as well. In ferritic steels, design stresses for

long-time service are typically 50 MPa at 550C. Often cavities


under

these

are

developed

conditions. In the laboratory, Cane (1979) and Cane and Middleton

(1981) found cavities in overheated 2 i /.Cr-1Mo steel at stresses between 62 and


147 MPa at 565 0 C. In the same material, Needham (1983) did tests down to 92 MPa
at 550C. He found cavitation failure in overheated material (austenized at
1300 oC), while material austenized at 950C failed in a mixed intergranularl
transgranular ductile mode at 92 MPa. At 132 MPa the material failed completely

5.

66

by

intergranular

Introduction to Part II

cavitation. This observation might be taken as an indication

that cavity nucleation becomes difficult below 100 MPa in the


finely

absence

of

the

dispersed sulfides which characterize the overheated material. In aust-

enitic steels, the minimum stresses applied by Needham and Gladman


by Chen and Argon (1981) ranged from 277 MPa at

550 0 C

In summary, cavities

metals

are

nucleated

in

pure

to 76 MPa at
below

(1980)

and

700 0 C.

10 MPa,

whereas

engineering alloys usually do not fracture within a reasonable time at stresses


below 50 to 100 MPa. Whether this extended lifetime is due to

tbe

suppression

of cavity nucleation or to excessively slow growth, remains to be clarified.

5.8 Pre-Existing Cavities


Cavities may be present in materials even before a creep test has been started.
For

example,

rapid cooling from the melt may cause decohesion of second-phase

particles if the thermal expansion coefficients of the particle and the

matrix

are sufficiently different and if the material is unable to accommodate the


difference by creep or some other deformation mechanism (Section 10.5). Similarly,

the growth of precipitates during a heat treatment of the material leads

to misfits, which might also cause decohesion (Sections 10.1 to 10.4).


In other cases, pre-existing cavities are introduced intentionally in order
bypass

to

the nucleation stage and to obtain information on growth alone. A first

method is to implant gas bubbles by heat treatments in


spheres.

These

appropriate

gas

atmo-

bubbles drastically reduce the creep life and creep ductility,

which in this case reflect the growth kinetics exclusively. Pertinent references are given in Sections 9, 11.2.7 and 12.4.3.
Further, the nucleation stage has been bypassed by prestraining at room temperature,

and a subsequent annealing treatment at higher temperatures. In nickel-

base superalloys this procedure produces cavities, probably by


boundary

interactions,

which

creep test (Dyson, Loveday


Shiozawa

and

slip-band/grain

cause a great loss in ductility in a subsequent


Rodgers,

1976;

Parker

and

Wilshire,

1980;

and Weertman, 1981, 1983; Burt, Elliott and Wilshire, 1981; Nazmyand

Duerig, 1982; Loveday and Dyson, 1983; and Pandey, Mukherjee and Taplin, 1984).
The results of some of these tests are evaluated in Section 12.4.

6 Nucleation of Creep Cavities/Basic Theories

Two mechanisms for cavity nucleation are described


rupturing of atomic bonds by high local stresses,

below,

namely,

especia~ly

first

the

across grain boun-

daries or other interfaces, and second the condensation of atomic vacancies.

6.1 Cavitiy Nucleation by the Rupturing of Atomic Bonds


The ideal cohesive strength of solids, aid' is determined by the forces between
neighboring planes of atoms. The following analysis relates aid to Young's
modulus of elasticity, E, and to the free surface energy, Ys ' In order to
establish

this

relationship the force per unit area between two atomic planes

is assumed to vary as

a(u) = (Eu/b) F(u/uo )'

(6.1)

when the planes are pulled apart. Here, u is the relative displacement
planes

of

the

(u = 0 corresponds to the equilibrium distance; see Fig. 6.1), b is the

atomic spacing,
the

atomic

Uo is an adjustable parameter which characterizes the range


of
forces, and F is an as yet unspecified dimensionless function with

F(O) = 1. The form of eq. (6.1) has been chosen such that it gives
elastic

behavior

the

linear

for small u (note that u/b is the extensional strain so that

a = EE). Further, the integral over a(u) is the mechanical work expended during
the

separation

of

the specimen halves. It must be equal to twice the surface

energy, since two new surfaces are created. This requirement determines

Uo

once

the form of F has been specified:

J a(u) du

o
where x

(E/b) u 2 J xF(x) dx
o 0

(6.2)

u/u o ' The ideal strength is the maximum of a(u) which is found to be

6.

68

Nucleation of Cavities

01

~
b

as to
u/b

Fig. 6.1. Interatomic forces for two hypothetic functions F(u/U o )

with uo/b

0.316.

where the dimensionless factor m is defined by


~

(2 I f xF(x) dx)1/2 Max[xF(x)].

x2 )-2, the value of m is


respectively. For E = 1.5.10 5 MPa,

If F is chosen as F(x) = exp(-x), exp(-x 2 /2), and (1


obtained

as

(6.4)

m = 0.52,

0.86

and

0.65,

Ys = 1.5 J/m 2 and b = 2 10- 10m, eq. (6.3) gives the ideal strength in the order
of magnitude of E/l0 in accord with results of Orowan (1948) and Kelly (1966).
0

Thus the ideal strength is of the order of 10,000 MPa, whereas typical stresses
applied

under

creep

rupture conditions are lOa-fold lower. Hence it is clear

that the rupturing of atomic bonds in the perfect crystal lattice cannot be expected to occur unless very high stress concentrations are built up locally.
Usually cavities are nucleated at interfaces, which are weaker than the lattice
and,

possibly,

have

other

properties

that favor cavity nucleation. For the

decohesion of a grain boundary, the grain boundary energy Yb is available and


need not be expended by the loading system. Hence, 2Ys in the above formulas
must be replaced by 2Ys -Yb Since typically Yb

= Ys/3,

this

substitution

im-

6.1

Nucleation by the Rupturing of Atomic Bonds

69

plies no great reduction of the ideal strength. A particle/matrix interface has


an interface energy Yi before separation and the energy Ys + Yp afterwards,
where Yp is the surface energy of the particle. Hence, one must replace 2Ys by
Y +Y -Y i in eqs (6.2) and (6.3). Without knowing these specific energies, it is
s p
clear that certain inclusions are weakly bonded, such as oxides or sulfides in
steels. On the other hand, carbides in steels are considered to be well bonded,
so

that

they

do

not

decohere easily by the rupturing of atomic bonds. From

tests on a spheroidized carbon steel and on a maraging steel at room


ure,

combined

the strength of the carbide/matrix interface to be E/100


tests,

the
so

1,500 MPa. In

creep

applied stresses are usually much smaller than in room-temperature

tests, but the stress concentrations by grain


higher,

temperat-

with a continuum-mechanical analysis, Argon and 1m (1975) infer

boundary

sliding

are

possibly

that it cannot be excluded that a local stress of 1,500 MPa can be

achieved under creep conditions, too.


An effect which cannot be neglected in the discussion of decohesion

of

inter-

faces is the segregation of impurity atoms. The role of impurities is described


in Chapter 8. One of the conclusions will

be

that

impurity

segregation

can

hardly reduce the cohesive strength by more than 50%.


In summary, the rupturing of atomic bonds can only occur at weakly bonded

par-

ticles or by stress concentrations which are probably not achievable under lowstress creep conditions as we shall see. An alternative

me"chanism

for

cavity

of

atomic

nucleation is described next.

6.2 Cavity Nucleation by Vacancy Condensation


6.2.1

Historical remarks and related subject areas

As we have seen in the preceding section,


bonds

the

athermal

rupturing

requires rather high stresses. At elevated temperatures, thermally acti-

vated processes might nucleate cavities at lower stresses, a possibility


is

which

now examined. Early in the history of research on creep cavities, Greenwood

(1952) and Greenwood, Miller and Suiter (1954) proposed

that

voids

could

be

nucleated by the condensation of atomic vacancies. The idea was first dismissed
since theoretical arguments seemed to indicate that vacancy precipitation
voids

into

was very unlikely under creep-rupture conditions (Machlin, 1956, McLean,

1958). Therefore, for about 15 years, void nucleation by the rupturing of atom-

70

ic

6.

Nucleation of Cavities

bonds at stress concentrators was favored. The idea of vacancy condensation

has been revived successfully by Raj and Ashby (1975) and Raj (1978b). In their
analysis,

the driving force for vacancy condensation is the mechanical stress,

whereas earlier workers had considered the supersaturation of vacancies

caused

by dislocation reactions to be the driving force for vacancy precipitation.


The work of Raj and Ashby (1975) and Raj (1978b) is based on a
treatment

of

rather

general

thermally activated nucleation phenomena developed by Volmer and

Weber (1926), Becker and Doring (1935) and Zeldovich (1943). It has been adopted

for

the

Anglo-Saxon literature in the English version of Frenkel's (1946)

book, by Fisher (1948), by Turnbull and Fisher (1949) and by Turnbull (1956). A
more

recent

review has been prepared by Russell (1970). Besides the classical

applications of the general theory to the nucleation

of

droplets

in

vapours

(Feder et aI, 1966), to the precipitation of second phases in solids and to the
cavitation of fluids under negative pressure (Volmer, 1939, Fisher,
has

1948),

it

also been applied to cavity formation in solids. Resnick and Seigle (1957)

examined pore formation in bi-metallic diffusion


and

couples

(Kirkendall

effect)

during diffusion in binary alloys. Here, a supersaturation of vacancies is

created by unequal diffusion velocities of the constituents of the alloy.


A similar effect is observed when, for example, nickel-chromium alloys are oxidized

at

high temperatures (Weber and Gilman, 1984). Chromium diffuses to the

surface to form a Cr203-scale, while Ni, being a slower diffuser, cannot fill
the vacant lattide sites fast enough, so that the supersaturation of vacancies
may lead to cavity nucleation on grain boundaries near the oxidized surface.
Another example where cavity nucleation
vacancies

is

the

phenomenon

of

is

driven

by

cavitational

supersaturation

swelling

during

or

of

after

irradiation with fast neutrons. Brailsford and Bullough (1972) have developed a
comprehensive

theory

for

caused by irradiation, while

the

concentrations

Russell

(1971,

of vacancies and interstitials


1978)

considers

the

thermally

activated nucleation of cavities due to the co-precipitation of impurities,


vacancies and self-interstitials which result from the irradiation damage.
6.2.2

Cavity shapes

One of the reasons that cavity nucleation occurs at distinct


in

the

nucleation

sites

grain boundaries, rather than homogeneously in the crystal lattice, is

that special cavity shapes may greatly reduce the

nucleation

barrier,

as

we

6.2

Nucleation by Vacancy Condensation

71

shall see. A variety of possible nucleation sites is now examined.

A cavity on a grain boundary, tends to assume the lenticular form shown in Fig.
6.2a. The angle ~ at the tip of the void is dictated by equilibrium between
surface tension forces, in this case by the balance of

the

specific

energies

per unit area of the free cavity surface, Ys ' and of the grain boundary, Yb
This balance leads to an angle of

Typically, the void tip angle is ~

70 0

Since a sphere minimizes the

surface

area for a given volume, the cavity surface consists of spherical caps provided
that the anisotropy of the surface energy can be neglected.
Similarly, the void shape at a second-phase particle in a grain boundary

(Fig.

6.2b) is characterized by the angles


(6.6)

cos cp

(6.7)

where Yp is the energy of the free surface of the particle and Yi is the energy
of the particle-matrix interface (always per unit area).
For self-similar cavit,y shapes, the cavity volume V, its surface S, the
of

length

its perimeter in the grain boundary, P, and the grain-boundary area that is

replaced by the cavity, B, can be expressed by powers of the radius of

curvat-

ure, p, of the spherical caps:


V = p3 f

(~),

(6.8)

...-/cav~fy~

...-/
grain
-----boUndary

21/1
2J.L

Fig. 6.2. Cavities on grain boundaries and at grain-boundary particles.

6.

72

Nucleation of Cavities

For the lenticular cavity in a grain boundary shown in Fig. 6.2a,

the

dimens-

ionless functions appearing in eq. (6.8) are given by


(2n/3) (2 - 3cosw

cos 3 w)

4n (1 - COSw)

(6.9)

2n sinw

For the cavity nucleus at a particle, of course,


depend

on

the

angles $ and

geometrical problem and find that the


approximated to within a factor of two by
f (W')

W'

with
be

(W

+ $ -

~)/2.

the

dimensionless

functions

as well. Raj and Ashby (1975) have analysed the


volumetric

(4n/3) (2 - 3cosw'

function

fv

cos 3 w')

W'

which

cannot

0, the void surface has a negative

curvature (if the curvature shown in Fig. 6.2 is defined to be positive).


implies

that

void

This

can be formed spontaneously, or, in other words, such a

particle when precipitated on a grain boundary


Similarly,

be

(6.10)

The exact expression involves integrals

solved in closed analytic form. For

can

would

be

called

non-wetting.

Raj and Ashby (1975) find that the curvature of a cavity surface at

symmetric triple,or quadruple grain junctions

vanishes,

function f (w) goes to zero when the void tip angle is


v

i.e.

the

volumetric

W= 30 0 or 36 0 , respect-

ively. This means, as we shall see, that for sufficiently small void tip angles
no

barrier

exists

for

cavity nucleation at grain junctions. Of course, this

cannot be the usual situation, since otherwise real polycrystals would spontaneously disintegrate or even could not be produced at all. However, this extreme
case illustrates the importance of the wetting angles for cavity nucleation.
6.2.3

The free energy of a cavity

The free energy of a cavity plays a key role in the theory


nucleation.
of

thermal

cavity

In the classical theory, the free energy of a cavity is written as

the sum of a surface term and a volume term. When the


energies

of

cavity

is

formed,

the

the surfaces which are freshly exposed must be expended while the

energies of the grain boundary and of the particle-matrix interface are gained.

6.2

73

Nucleation by Vacancy Condensation

Thus the surface-related part of the free energy,

~G,

is given by

where fi is the geometrical function for the newly exposed area of the inclusion, analogous to eq. (6.8). However, this form of the surface-related free
energy can be replaced by a much simpler one. For
clear

in

the

reasons

which

will

become

paragraph following eq. (6.13) the surface-related term must be

equal to 3 p2 Ys fv(w'), although the author found it difficult to


directly from geometrical considerations.

prove

this

The volume term of the free energy is either given by the (negative) work

done

by

the applied stress during void enlargement or it can arise from a supersat-

uration of atomic vacancies. The stress-related case will be discussed first. A


grain

boundary

containing atomic vacancies at the thermal equilibrium concen-

tration is subjected to a normal stress a. Now let a cavity grow by accepting a


vacancy

and

let

the

equilibrium

number of vacancies be restored by thermal

generation at sources such as jogs in grain boundaries. Then the volume of

the

solid increases by the atomic volume 0, and the applied stress does the work an
for each vacancy absorbed, i.e., the stress-related part of the free energy
a

cavity

with

volume

of

V is equal to -aVo Therefore the total free energy

change due to the formation of a cavity is given by


(6.11)

~G

It should be mentioned that also the elastic elongation of

the

specimen

upon

cavity formation contributes to the free energy (Brinkmann, 1955; McLean, 1958;
Hirth and Nix, 1985). However, the elastic

work

can

generally

be

neglected

since it is by a factor of the order alE smaller than aVo


From the functional form of
voids

(subcritical

~G(p)

shown in Fig. 6.3

it

is

clear

that

small

vacancy clusters) tend to shrink since they then gain free

energy, whereas large (supercritical) ones tend to grow


The critical radius of curvature is found by setting

to

a~G/ap =

macroscopic

size.

D, which leads to
( 6.12)

The associated height of the nucleation barrier is

74

6.

Nucleation of Cavities

<.!)
~
........

<.!)

0.5

1.5

0.5

Fig. 6.3. Free energy vs. radius of curvature of the cavity nucleus.

(6.13)

L'.G*

This equation shows that the nucleation barrier decreases with stress

1/02

as

surfaoe energy as y3. In the limiting case of non-wetting


s
particles, is fv(w') = 0, and the nucleation barrier vanishes.

and

increases

with

Now the reason is given why the surface-related terms of the free
the

energy

have

simple form Used in eq. (6.11). If the dependences of the surface term and

the volume term on the wetting angles differed from one another, the result for
the

critical

radius,

p*,

would have to depend on these angles. However, eq.

(6.12) must be generally valid independent of


equilibrium,

thermal

wetting

angles,

since

in

chemical potential of an atom in the stressed grain

the

boundary at the v0id tip,

the

].I

-an

[see eq. (4.2)], must be equal to that on the

cavity surface, ].I = -ys n(K1+K 2 ) = -2Ys n/p [see eq. (4.11)]. I f the
chemical potential were not continuous at the transition from the grain bound-

spherical

ary to the cavity surface, an infinite diffusive flux of matter to or from the
cavity would result. Equating the two chemical potentials gives eq. (6.12),
which

proves

that

the surface-related terms of the free energy have the form

assumed.
We now turn to the case that a supersaturation of vacancies provides the
ing

force

for

vacancy

kT in(cv/co )' where

Cv

precipitation.

The

driv-

free energy of a vacancy is then

and Co are the concentration of vacancies and its

equi-

6.2

Nucleation by Vacancy Condensation

librium

value,

force if

an

75

respectively. Supersaturation and stress give the same driving

kT In(c Ic ). Thus a stress of 100 MPa corresponds to a supersatv

uration of cv /c o -l ~ 10% (for T s 1000K). Such a large supersaturation does


generally not occur except in situations in which a Kirkendall effect occurs or
under

irradiation. Therefore we proceed assuming that under creep-rupture con-

ditions, stress is the primary cause for cavity nucleation.


Incidentally, the free energy changes originating from stress and
saturation

from

super-

are not additive: precipitation of vacancies into a cavity does not

increase the volume of the specimen and is therefore not associated with mechanical

work,

unless

new

vacancies

are formed at vacancy sources. As long as

there is a vacancy supersaturation, thermal equilibrium

does

not

demand

the

generation of vacancies to replace those which have been absorbed in cavities.


6.2.4

The nucleation rate according to Raj and Ashby

The time rate at which subcritical clusters are pushed

across

the

nucleation

barrier by thermal fluctuations to form stable cavities is proportional to the


Arrhenius factor exp(-AG*/kT). The pre-exponential frequency factor must be
determined

from the kinetics rather than from the (equilibrium) thermodynamics

of vacancy cluster formation. The simplified approach of Raj and


to

determine

the

frequency

factor

will

be

presented

Ashby

first

presentation requires elementary mathematics only. A more precise

(1975)

because

its

analysis

of

the kinetics will follow in the next sub-section.


Raj and Ashby (1975) argue that there are cmax potential nucleation sites per
grain boundary area, c of which are already occupied by supercritical cavities
so that cmax - c remain for subcritical vacancy clusters. The area
critical clusters is then assumed to to be
N~

density

= (c max - c) exp(-AG*/kT).

(6.14)

This expression would be correct if the vacancy clusters were in thermal


librium,

of

equi-

i.e., if there were no drain of clusters towards supercritical cavit-

ies. Further, Raj and Ashby (1975) assume that the nucleation rate

(number

of

cavities generated per unit time and area) is given by the product
J*

B*N~,

(6.15)

16

6.

Nucleation of Cavities

where 6* is the rate (per unit time) at which single vacancies are absorbed
a

by

critical cluster. The possible emission of vacancies is neglegted, which may

cause substantial numerical errors as we shall see. The absorption rate


calculated

subject

6*

is

to the assumption that vacancies are available directly at

the edge of the cavity at their equilibrium concentration, i.e., that effective
vacancy

sources

prevent a depletion near the void. Further it is assumed that

the vacancies approach the void through the grain boundary only. Then 6*

takes

the form

(6.16)
The first bracketed term is the number of atomic sites along the
the void in the grain boundary where f

perimeter

of

was defined in eq. (6.8) and 0 is grain

boundary thickness. The second term in brackets represents the concentration of


vacancies in thermal equilibrium on a grain boundary subject to a normal stress
o. The first two terms together represent the number of

atomic

distance

vacancies

within

one

from the void perimeter. The third term is the jump frequency

of a vacancy expressed by the vacancy diffusion coefficient in a grain boundary


2

(in m Is), using the Einstein relation for two-dimensional diffusion. The
vac
factor 1/4 accounts for the fact that only one out of four in-plane jumps leads

from

the edge into the cavity. The vacancy diffusion coefficient is related to

the grain boundary diffusion coefficient through oD b = Co oD vac ' In most cases
is 00 kT so that the exponential function in the expression for ~* can be
expanded in a power series. Eventually, the final

result

for

the

nucleation

rate is obtained by combining eqs. (6.12) to (6.16), which leads to:

J*

2 f p (1jJ') Ys

04/3

(1

00 ) oD b (c
-c) exp[kT
max

4 y3 f

(1jJ')

J.

(6.11)

kT

The most important part of this result is the exponential factor. It makes
nucleation

rate

a very sharply increasing function of the stress

the

so that at

low stresses, cavity nucleation occurs at a few random nucleation sites whereas
at

high

stresses

all

nucleation sites are activated quickly. A quantitative

evaluation is deferred until a better model has been discussed in the following
subsections.

The

dominant

influence

of the exponential function will not be

altered, but a different pre-exponential factor will be obtained.

6.2

Nucleation by Vacancy Condensation

6.2.5

77

The Fokker-Planck equation

The formulation of the nucleation problem due to Zeldovich (1943)

leads

to

deeper insight and to modified and extended results compared to the formulation
of Raj and Ashby (1975).

We

denote

the

area

density

of

vacancy

clusters

containing n vacancies by N(n), which is the unknown variable to be determined.


The independent variable, n, and the
served

as

the

radius

of

curvature,

which

previously

independent variable, are related through the requirement that

they must give the same cluster volume:


(6.18)
If cavity nucleation were suppressed by some intervention mechanism,

the

sub-

critical clusters would be in thermal equilibrium and their density would be


No(n) = (c

max

(6.19)

- c) exp(-AG(n)/kT),

where cmax - c is the area density of nucleation sites available for the subcritical clusters and the free energy of a cluster, AG, is given by eq. (6.11).
InCidentally, Raj and Ashby
assumption is now dropped.

(1975)

tacitly

assume

that

N(n)

N (n).
o

This

The net flux of clusters in size space, I n , from the size class containing n
vacancies to the class containing n+1 vacancies is given by the difference in
number density

betwee~

clusters of size n which absorb a vacancy and

those

of

size n+1 which emit a vacancy (per unit time):


J

n,

= 8(n)N(n) - 8 ' (n+1)N(n+1).

(6.20)

Here 8(n) and 8 ' (n) are the rates (in number per unit time and
which

cluster

unit

area)

at

of size n absorbs or, respectively, emits vacancies. The ab-

sorption rate is given by an expression analogous to eq. (6.16),


(6.21)
if eq. (6.18) is used to replace p by
expressed

n.

The

emission

rates

8 ' (n)

can

be

by the absorption rates employing the principle of detailed balance.

This principle requires that in thermal equilibrium the net flux


classes is zero. Setting I n = 0 and N

No in eq. (6.20) leads to

between

size

78

6.

Nucleation of Cavities

(6.22)
Using this expression

in

eq.

(6.20)

and

replacing

finite

differences

by

differentials for convenience, gives the cluster flux in size space:


(6.23)
The cluster flux must satisfy the continuity condition in size space

aN(n)

(6.24)

at

This equation is
governing

sometimes

differential

called

the

Fokker-Planck

equation.

It

is

the

equation for the area density of clusters containing n

vacancies. The equilibrium density No(n) and the absorption rate S(n) are known
from eqs. (6.19) and (6.21). The partial differential equation is linear and
resembles the diffusion equation in one-dimensional space.
6.2.6

The steady-state nucleation rate and the critical nucleation stress

The steady state is characterized by a cluster flux which is independent of

n,

and that constant flux defines the cavity nucleation rate, J*. With I n = J*,
eq. (6.23) can be integrated by separation of the variables, Nand n. The integral extends from the smallest clusters(n

1) to the largest ones (n

= ~):

- Jd[N(n)/N (n)]
1

(6.25)

The lower limit of integration on the right-hand side is unity since the number
of

'clusters'

containing one

vacancy is equal to its equilibrium value. The

upper limit on the right is zero since, for large n, No goes to infinity

while

N remains finite. The value of the integral on the right is simply 1. The integral on the left-hand side is conventionally evaluated in an approximate way by
arguing that the greatest contribution to the integral comes from the region
where No has a minimum. The minimum of No exp(-~G/kT) occurs where ~G has a
maximum, i.e. at n = n*. Hence ~G is expanded in a power series up to quadratic
terms in n-n* at the maximum,

~G*;

and Sen) is assumed to be a

slowly

varying

function with a value S* at the critical point. Then eq. (6.25) gives
J*

Z S*

N~

(6.26)

6.2

79

Nucleation by Vacancy Condensation

with the so-called Zeldovich factor


(6.27)

which was missing in the result of Raj and Ashby (1975),


second

form

of

eq.

(6.15).

In

the

eq. (6.27), 6G was inserted from eq. (6.11). For 0

2
Ys = 1.5 JIm,
T = 850K and

fv(~')

100 MPa,
~
= 10 ,the Zeldovich factor is Z = 4.10- 4

Inserting eqs. (6.14), (6.21) and (6.27) for N~, B* and Z into eq. (6.25) leads
to the final form of the steady-state nucleation rate:

J*

fp(~')

0 /jD b (cmax - c)
exp(- 4 Y; fv(~') ).
1/3
2 0
[3w kT Ys fv(~,)]1/2
02 kT

(6.28)

Here, 00 was neglected compared to kT for simplicity. Since the stress

appears

in the exponent, this result is so strongly stress dependent that one can reasonably, albeit not in a .strict sense, speak of a

critical

stress

for

cavity

nucleation, 0nuc
This nucleation stress can be calculated by the following
argument. Practically all nucleation sites are activated within a given testing
time,

t,

if J*

cmax/t. The stress at which this occurs is obtained from eq.

(6.28) by replaCing J* by cmax/t, setting c = 0 and resolving for the stress in


the exponential:
4 y3 f (~') IkT
s

o
nuc

(6.29)

The logarithm in this equation is a slowly varying function of its argument. It


assumes

values

between 30 and 40 for times between a day and a few years, for

temperatures around half the melting temperature, for 0 between 10 and 1000 MPa
and for material parameters of a-iron. Taking the long-time value, 40, gives
(6.30)
In the second form, Ys = 1.5 J/m 2 and T = 850 K were inserted as representative
numerical values. At an arbitrary location in the grain boundary, the volumetric function of a lenticular cavity is f (70 0 ) = 2.1. Cavity nucleation by vacancy

condensation

would therefore require a stress of 0nuc = 7,800 MPa, which


is far greater than the stresses at which nucleation is observed to take place

80

6.

(below

Nucleation of Cavities

100 MPa in engineering alloys and below 10 MPa in pure metals). This is

a problem, which will be taken up again repeatedly starting in Section 6.3.


Two alternative interpretations of the nucleation condition,
(6.30)

are

as

follows.

First

it

eqs.

(6.29)

and

is interesting to calculate the number of

vacancies which constitute a critical cluster.

From

the

critical

radius

of

p* = 2Ys /o, and from eq. (6.18), the number of vacancies follows to
be n* = f (2Y 10)3 /U The uncertain value of f is eliminated using eq. (6.29).
v
s
v
If the logarithm in eq. (6.29) is again chosen as 40, there results

curvature,

n*
Since generally kT

80 kT/(oU).

(6.31 )

oU, a critical cluster is predicted to contain at least a

few

hundred vacancies. Alternately, the condition for cavity nucleation can be


interpreted in terms of the nucleation barrier bG*
4 y3f 102 Then eq. (6.30)
implies that nucleation occurs if bG* < 40 kT.

6.2.7

s v

Transient solutions of the Fokker-Planck equation and incubation times

At the very beginning of a creep test, vacancy clusters

still

have

the

size

distribution of the unstressed state. Therefore, a certain incubation time, t i ,


must elapse until the cluster size distribution N(n) has adjusted itself to the
new,

stressed state. Not before the new equilibrium state is reached, nucleat-

ion can take place at the steady-state rate J* calculated above.


The basis for calculating the transient behavior quantitatively is the
Planck

Fokker-

eq. (6.24). No exact closed-form solutions are available, but upper and

lower-bound estimates for the incubation time, t i , can be made. An upper


is obtained if

eq~

bound

(6.24) is written in the form

aN
at
and the second term on the right-hand side is neglected.

(6.32)

Since

this

term

is

positive in the interesting range of nand t, the first term alone under-estimates the time rate aN/at of approach to the steady state. The
is

incubation

time

therefore obtained too large. With this approximation made, eq. (6.32) is a

diffusion equation for NINo with an n-dependent 'diffusion coefficient',

Sen).

With Sen) ~ n 1/3 as in eq. (6.21), the diffusion equation has similarity solut-

6.2

Nucleation by Vacancy Condensation

81

ions of the form N(n,t)/No(n) = f(n/t 3/5 ), where the function f is obtained by
substitution into eq. (6.32) truncated after the first term on the right-hand
side. The resulting ordinary differential equation for f can be solved by separation of the variables. The cluster current, from eq. (6.23), is found to be
(6.33)
The numerical factor

cannot be determined Since, by

truncating

eq.

(6.32),

one looses the boundary condition that I n must approach J* for l0ng times.
Since the value of ~ is of no interest here, it is arbitrarily set equal to 1.
Back-substitution

of

eq.

(6.33) into eq. (6.32) shows that the ratio between

the first and the second terms on the right-hand side varies as 1/t,
first

term

is

i.e.

the

much larger than the second for short times. Hence, eq. (6.33)

r.epresents a short-time solution of the complete eq. (6.32).


With n

n* and B

B*, eq. (6.33) gives the nucleation rate, which is

plotted

in Fig. 6.4a. Time is normalized by the characteristic time


t(Ub) _ (9/25) n*2 /B *,

(6.34)

which appears in the exponential of eq. (6.33) and which is our upper bound for
the
.n*

incubation
=

time. Inserting the number of vacancies in a critical cluster,

f v (2Ys /o)3 /0 , and B* from eq. (6.21) with n

n* gives

t(Ub)

(6.35)

A lower-bound estimate was derived by Trinkaus and Yoo (1986), who solved the
full eq. (6.32) making the approximations to expand ~G(n) at the critical point
up to quadratic terms in (n-n*) and to consider the reaction rate as a constant, B(n) = B*. Since B(n) is actually smaller than B* for subcritical
clusters, this approximation overestimates the nucleation rate, particularly at
short times. The non-steady nucleation rate is then obtained as
J*(t)/J*(m) ~ (1 - e)-1/2 expf(1-1/e)-1~G*/(3kT)},
where J*(m) is the steady-state nucleation
abbreviation defined for local use only, and
t (R.b)
i

3 kT n*/(o 0 B*)

rate,

(6.36)
(R.b) )

e = exp(-2t/t i

is

an

(6.37)

6.

82

Nucleation of Cavities

is the lower-bound estimate for the incubation time. It differs from the upperbound estimate given in eqs. (6.34) and (6.35) by a factor 4.2 kT/AG*, which is
typically 0.1. Figure 6.4b shows the nucleation rate given by eq. (6.36).
Interestingly, the characteristic time given in eq. (6.37) was also obtained by
Binder

and Stauffer (1976) who analyzed the asymptotic approach of the cluster

size distribution function to the steady state. Thus, the lower-bound

estimate

of ti refers to the long-time behavior, while the upper bound was obtained from
a short-time analysis. Less accurate upper and lower-bound estimates than those
given here were reported by Raj (1978b).
Numerically, incubation times are generally short compared to the duration of a
creep test. For Y = 1.5 J/m 2 , a = 100 MPa, f2 If = 10-4 and for the grain
v p
s
boundary diffusion coefficient of ferritic steel given in Appendix A, the upper
1 sec at T = 850K and ti = 20sec at T = 750K,
1
while the lower bound gives ten times smaller values.
eq.

bound,

(6.35),

yields t.

6.3 Discussion of Cavity Nucleation Theories


There are two apparent discrepancies between the observed
and

the

nucleation

behavior

basic theories of cavity nucleation. First, the calculated nucleation

stress is substantially higher than observed ones. Second, theory predicts that

0.1,

t a3

0.75

.....8

-...

.8

f;-

~ a2
.....
....

:::f;-

f;""")

b)

0)

a5

""")

0.1

025

o
,

tlt.(ub)

Fig. 6.4. Transient cavity nucleation rate: a) Short-time solution,


eq. (6.33) with n = n* and a = 13*; b) eq. (6.36). AG*/kT = 40.

6.3

Discussion of Nucleation Theories

83

nucleation occurs more or less instantaneously once the

nucleation

stress

is

applied (apart from the incubation time which was found, however, to be small),
whereas actually nucleation occurs continuously throughout the creep life,

and

the nucleation rate is controlled by the strain rate rather than by stress.
6.3.1

A theoretical remark

The classical nucleation theory described above has been criticized for several
reasons,

in

particular for the use of macroscopic quantities such as the sur-

face energy on an almost atomistic scale. More advanced models have been worked
out

(for example, Feder et aI, 1966; Binder and Stauffer, 1976) the outcome of

which has been that classical nucleation theory cannot predict absolute nucleation

rates

correctly but the onset of profuse nucleation as a function of the

driving force (in our case: stress) is predicted well. Indeed, some

phenomena,

such as the nucleation of droplets in supersaturated vapor, can be described by


the classical theory (Volmer, 1939). In the case of

creep

cavities,

however,

even the observed nucleation stress cannot be predicted reasonably accurately.


6.3.2

On possible causes for the discrepancy between theoretical and experimental nucleation stresses

According to eq. (6.30), the theoretical nucleation stress is typically of

the

fv 1/2

order anuc = 5,360


MPa. A first possibility to reconcile this prediction
with the low observed nucleation stresses is that the wetting angles at the
cavity

nucleus

could

be such that fv(w') takes values very much smaller than

unity. If nucleation is to occur at a stress level of 10 MPa, eq. (6.30) requires f = 3'10-6 , which means practically non-wetting particles. It is difficult
v

to believe that such small values of fv are a general phenomenon

in

materials

which cavitate, even if impurity segregation is taken into account (Chapter 8).
Second, the theoretical nucleation stress calculated above is

local

stress

acting on a size scale comparable to the void nucleus dimensions. Compared to


the applied stress, the local stress may be enhanced by grain boundary sliding
or

by

slip

bands. This possibility will be considered in Chapter 7, although

grain boundary sliding is most pronounced on boundaries inclined to the applied


stress, while cavitation preferentially occurs on transverse boundaries.
Third, a gas pressure developing in the cavity n,ucleus may promote
There

are

certainly

special

situations

nucleation.

in which a gas pressure plays a key

84

6.

Nucleation of Cavities

role, but in other cases it cannot be of great importance (Chapter 9). Finally,
stresses around growing precipitates will be considered in Chapter 10.

6.3.3

The problem of continuous cavity nucleation

Considering the smallness of the incubation times calculated in Section


it

6.2.7,

can be excluded that the incubation time, together with its possible varia-

tion from one nucleation site to another, can


occurs

continuously

over

explain

why

cavity

likely that continuous cavity nucleation 1s a consequence of some


centration

process

nucleation

substantial fractions of the creep l1fe. It is more

which

stress

con-

occurs throughout the creep life and which triggers

cavity nucleation. In nickel base superalloys, there is relatively clear

evid-

ence what this triggering event might be. Shiozawa and Weertman (1983) found an
almost perfect one-to-one correlation between the cavity spacing and the
ing

spac-

between coarse slip bands, which are typical for these alloys. The density

of both increases in linear proportion to strain.


cavity

Thus,

in

these

materials,

nucleation is continuous since the slip bands which initiate the cavit-

ies are formed continuously.


Argon, Chen and Lau (1981) propose that the triggering
cavity

events

for

continuous

nucleation come from the spasmodic nature of grain boundary sliding ob-

served by Chang and Grant (1953) in coarse-grained high-purity aluminum.


from

the

lack

of

Apart

evidence that sliding occurs intermittently in engineering

alloys as well, this suggestion does not explain how cavities are nucleated

on

transverse boundaries which hardly slide but nevertheless cavitate.


Evans (1984) argues that the observed increase in cavity

number

might

be

an

artefact the real situation being as follows. All cavities are nucleated virtually instantaneously and grow by a mechanism of grain boundary
the

rate

sliding.

Since

of sliding varies from boundary to boundary, the cavities become ex-

perimentally detectable at different times, and an observer would conclude that


nucleation occurs continuously. However, this explanation can hardly apply to
materials in which small cavities grow by diffusion, since in this case the
cavities

reach

the

detectability

limit quickly, as Dyson (1983) has pOinted

out. Further, Evans' assumption of growth by grain boundary sliding


only

to

situations

in

apply

which inclined boundaries cavitate preferentially. In

commercial materials, however, cavitation


important.

can

on

transverse

boundaries

is

more

7 Cavity Nucleation by Stress Concentrations


During Creep

The analysis of cavity nucleation in the preceding chapter indicates that local
stress

concentrations

by

grain boundary sliding or by slip within the grains

are necessary to initiate cavities. This aspect is now quantified. It should be


mentioned,

however,

that the theoretical calculation of stress concentrations

will not substantially improve our understanding of the nucleation process


yond

the

be-

facts described in previous sections: in commercial materials, cavi-

ties are formed preferentially at particles on boundaries normal to the applied


stress, which suggests that sliding does not play an exclusive role; nucleation
occurs at low stresses, which points to stress concentrations; and the nucleation

rate is approximately proportional to the strain rate, which suggests that

slip within the grains is involved in the stress concentration process.


This chapter is organized according to different configurations, starting
the

stress

stresses focussed on triple grain junctions and on particles by grain


sliding

with

analysis of a slip band impinging on a boundary. Subsequently, the

are

described.

Each

boundary

of these configurations is analyzed for elastic

material response, power-law creep and diffusion. In early papers, for

example

in that by McLean (1956/57), stress concentrations were calculated elastically,


using and extending results which Stroh (1954) had developed in the

theory

of

cleavage crack initiation at low temperatures. At elevated temperatures, however, diffusion and dislocation creep tend to relax stress concentrations. Nevertheless,

the

elastic

approach

will also be pursued, not only for historical

reasons but also because elastic stress concentrations are generally the severest and, in this sense, represent a limiting case. Furthermore, elastic material behavior cannot

a~ways

be excluded. For example, when grain boundary sliding

occurs spasmodically, as observed by Chang and Grant (1953) and by Intrater and
Machlin (1959/60), the elastic response of the material determines
field

at

short

times

the

stress

after such a sudden sliding event. Rupturing of atomic

bonds, which ensues instantaneously once a

critical

stress

is

reached,

intervene before stress relaxation by diffusion or creep has taken place.

can

7. Stress Concentrations

86

7.1 An Isolated Sliding Grain Boundary Facet (Shear-Crack Model)


The simplest configuration to be discussed here is
facet

of

length

d.

Depending

sliding

factor near unity from the grain size. Nevertheless, the same
for

both

quantities.

Due

grain

boundary

on the shape of the grains, d may differ by a


letter

is

used

to the presence of particles and ledges, the grain

boundary transmits a shear stress Tb , which possibly depends on the sliding


rate U
b ' At the ends of the facet, sliding is piled up against impenetrable
barriers representing, for example the triple pOints in a polycrystal (Fig.
7.1a). A similar configuration is that of a slip band within the grain piled up
against the grain boundary (Fig. 7.1b).
7.1.1

Elastic analysis of a sliding facet

The equations of elasticity must be solved subject to the


that

the

sliding

boundary

conditions

boundary transmits the shear stress Tb' while at infinity a

uniform resolved shear stress Tm is applied. Tensile stresses do not affect the
analysis in linear material. The problem is identical with a dislocation
pile-up problem (Bilby and Eshelby, 1968) or with a Mode-II crack,
ions

the

solut-

for which are well known (see, e.g. Rice, 1968a, or Appendix B). If rand

e are polar coordinates centered at the end of the sliding boundary segment,
~ress

field for rid

aeo

0)

/reo

I
I
I
I

I
I

I
I

0 has the general form

(if uniaxial
tension)

b)

taco

8/tT,,;,

Fig. 7.1. a) Sliding grain boundary facet. b) Slip band.

7.1

A Sliding Boundary Facet

87

Kn

Y'21Tr

(7.1)

f. J.(6).
1

The dimensionless angular functions f .. (6) are independent of the outer speclJ
imen geometry, and they are given in Appendix B. The stress intensity factor
KII depends on the specimen geometry and on the load.

In

particular,

for

an

isolated crack (or sliding boundar'y segment) in plane strain or plane stress in
an infinite body is

(7.2)
For a penny-shaped crack (or boundary facet) of diameter d in an infinite body,
KII has the maximum value

The maximum of KII occurs at the pOints where the crack front is normal to
shear direction (Rooke and Cartwright, 1974).
For later use, the shear displacement across the penny-shaped crack

is

the

quoted

from Westmann (1965):


(7.4)

Here, p is the distance from the center of the circular crack measured

in

the

plane of the crack. It varies from 0 to d/2.


If the applied loadirtg is uniaxial tension with the tensile axis being inclined
at an angle a to the normal on the boundary facet, the resolved shear stress on
the boundary,

Tm'

is related to the uniaxial tensile stress,


Tm =

(1/2)

am

am'

through

sin2a.

(7.5)

The tensor transformations for general loading are not needed here.
In relation to cavity nucleation, the tensile stress occurring locally near the
end

of

the

sliding

boundary

stress is given in Appendix B as

is

of interest. The maximum principal tensile

7.

88

Stress Concentrations

1.322

The second form is for

(7.6)

121fT'

-103.2 0 where the maximum of aI occurs.

Next the question is examined whether this


cause

2.!.....

stress

concentration

suffices

to

rupturing of atomic bonds. This requires that the maximum tensile stress

exceeds the ideal strength of the material over one atomic spacing, b.

Setting

aI = aid with aid from eq. (6.3), r = b, 'b


0 and a = 45 0 , and taking KII
from eq. (7.3) for the penny-shaped facet (with v = 0.3) yields the critical
applied tensile stress for cavity nucleation:

Typical numerical values for metallic materials are E = 1.5.10 5 MPa, Ys


1.5 J/m 2 and d = 30~m. For m = 0.7 this gives the stress for cavity nucleation:
a~ = 245 MPa. Compared to the stress levels at which cavity nucleation is still
observed to occur, this is a rather high value, although the elastic stress

analysis gives an upper bound for the stress concentration and, hence, a
bound

for

the

applied

lower

stress which is necessary to nucleate cavities. Thus,

rupturing of atomic bonds at the end of a

sliding

boundary

facet

is

not

viable mechanism for cavity nucleation, unless the cohesive strength is reduced
substantially, say, by the presence of a weakly bonded particle.
Smith (1966) using an energy balance argument derived a result which is equivalent

to eq. (7.7) apart from a numerical factor. This coincidence is not fort-

uitous since the energy balance argument and the cohesive strength argument adopted

here are equivalent in relation to the rupturing of atomic bonds in ela-

stic square root singularities. The energy balance

can

be

expressed

by

the

following formula, which is well known in fracture mechanics (Rice, 1968a):

Cavity nucleation by vacancy condensation requires somewhat lower stresses


maintained

but

over a greater region having the size of the cavity nucleus. Taking

r = 10 nm in eq. (7.6) and d = 30

~m

in eq. (7.3) shows that the stress is

en-

hanced by a factor 13.5 compared to the applied tensile stress. In the light of
eq. (6.30), this is hardly sufficient to cause nucleation at

applied

stresses

around and below 100 MPa, unless the volumetric function fv is very small.

A Sliding Boundary Facet

7.1

7.1.2

89

The sliding boundary facet (shear crack) in creeping material

A power-law creeping material is described by the nonlinear

which

(3.6),

reduces

to

E =

Ban

One

nonlinear

materials

viscous

have

been

the results is that the stress field at the end of a slip

of

plane is the HRR-field of a Mode-II crack, given in eq. (3.25). In a


material

eq.

in uniaxial tension. In Section 3.4, a few

general features of stress fields in


summarized.

law,

viscous

nonlinear

it is not a simple matter, however, to calculate the value of the C*-

integral as a function of the applied stresses a~. and of Lb. Due to the nonlJ
linearity of the equations, the components of the applied stress can enter into
the problem in a complicated
resolved

shear

stress

plays

(1981) have developed an

manner,
a

whereas

in

linear

material

only

the

role. For a tensile crack, He and Hutchinson

approximate

method,

which

has

turned

out

to

be

sufficiently accurate. Riedel (1984b) has applied the idea to a shear crack.
The idea is based on a linearization of the nonlinear viscous material law. The
linearization procedure conveniently starts from an initial state of stress for
which the stress analysis is trivial. Such a starting solution is obtained, for
example,

if only a shear stress is applied, and this shear stress is just bal-

anced by the tractions on the crack faces, LOO


shear

stress

field.

Lb. This simply gives a uniform

In the next step, Lb is allowed to deviate slightly from

LOO. The linearization around the initial state for small (Loo-Lb)la: leads to
problem of linear, anisotropic elasticity, which can be solved.
For plane

strain~ th~

final result for C* and for

the

sliding

rate

averaged

over the facet, Ub' are found in this way to be


(7.9)

C*

(7.10)
Here a
(/3/2) [(a;2-o~1)2 + 4L:]1/2 is the equivalent applied tensile stress
e
in plane strain and 00 = B(ooo)n. This linearized solution is exact for n = 1
e
e
and asymptotically exact for small (Loo-Lb)/o:. He (1983) has performed
numerical calculations for the shear crack with Lb
up to n

0 [i.e. (Loo-Lb)la:

10. Comparison with his results shows that the error

of

eqs.

1//3]
(7.9)

and (7.10) due to the linearized treatment of the problem is never greater than
3%. For Lb > 0, the situation should
Lb

become

Loo' the linearization becomes exact.

even

more

favorable

since

for

7.

90

Stress Concentrations

A penny-shaped shear crack is a more realistic model for a grain-boundary facet


than

is

a plane-strain crack. In the literature, the penny-shaped shear crack

in nonlinear material has not been analyzed. The


ploying

the

to the problem of a circular shear crack in


problem

problem

can

which

is

em-

anisotropic

elastic

material,

transversely

iso-

materials, which would be directly useful for our problem. However, his

result is obviously incorrect, because it


limit

solved

a special case of an elliptic crack under arbitrary loading

considered by Hoenig (1978). He works out the solutions for


tropic

be

same linearization method as in the plane-strain case. This leads

correctly.

Since

does

not

reproduce

the

isotropic

re-calculation of Hoenig's results would be rather

time-consuming, the author proceeds by assuming that only the quantity S given
in Hoenig's eq. (17) is incorrect, whereas the general form of the solutions is
accepted. Now a supposedly correct value of S is obtained by adjusting Hoenig's
results

to

He

and

Hutchinson's (1981) solution for tensile loading. Since S

depends on material parameters only but not on the

loading

system,

the

same

value is used for the shear crack. In this way, one obtains the average sliding
rate across a circular, isolated grain boundary facet:
(7.11)
In the linear case, n
for

1, this coincides with Westmann's (1965)

exact

result

the limit of incompressible material [cf. eq. (7.4) with v=1/2J. It should

be noted that this result for the penny-shaped crack differs from that for

the

plane-strain crack only by an n-dependent numerical factor.


The viscous stress concentration is weaker than the elastic one.

For

example,

under

the

conditions where the elastic stress concentration is 13.5-fold


0
0
(t b = 0, a = 45, e = -103.2 , r = 10 nm, d
30 llm), eq. (3.25) with eq. (7.9)
gives a 3.6-fold vi~cous stress concentration if n = 5.

7.1.3

Relaxation of elastic stress concentrations at a shear crack by creep

The elastic solution and the power-law viscous solution described in


ceding

pre-

subsections can be regarded as limiting cases occurring in elastic/non-

linear viscous material. Such a material is


(3.6)

the

which,

in

characterized

uniaxial tension, reduce to

E=

olE

by

eqs.

represents

the

long-time

limit.

discussed in greater detail in connection with

The

transient

cracks

to

Bon. The instantaneous

response of such a materlal to load application is elastIc, while


response

(3.4)

in

the

viscous

behavior

will be

fracture

mechanics

7.1

A Sliding Boundary Facet

(Chapter

91

23). Without deriving the results here we use the characteristic time

for the transition from the initial, elastic behavior to the long-time limit:
2

KII (1-\1 )/E

(7.12)

(n+1) C*
Inserting the expressions for KII and C* given
leads to

in

the

preceding

subsections

for the plane-strain shear crack. In order to obtain a numerical estimate,

the

above

equation is divided by the lifetime, t f and the Monkman-Grant product,


Ee t f = CMG , is introduced (Section 2.3.1). Further, we set n = 5 and \I = 0.3
to obtain

'"
If ae/E

10 -3 and CMG = 0.03, this ratio is t1/tf = 11332. Thus the elastic
stress concentration is relaxed within a small fraction of the lifetime.
7.1.4

The time to build up elastic stress concentrations

In the preceding section, the relaxation of elastic


described.

On

the

other

hand,

it

requires

stress

concentration

a certain time to build up the

elastiC stress concentrations. This is so because the external loading rate


finite,

but

even

was
is

if the load is applied rapidly, the resistance of the grain

boundary against sliding implies a finite rise time for the elastic stress concentration. For an isolated grain boundary facet this is modeled in the following, approximate manner. The grain boundary is assumed to

slide

according

to

the linear viscous law


(7.15)
with the friction coeffiCient, n, of the grain boundary as discussed in Section
1.5.1.

The

surrounding material is elastic. Besides the geometrical idealiza-

tion of the grain boundary network in a polycrystal by an isolated


approximation

consists

in

employing

eq.

facet,

the

(7.15) averaged over the facet and

92

7.

ignoring the variation of the shear stress

~b

Stress Concentrations

on the facet.

This

reduces

the

mathematical complexity substantially. Now, only ~b from eq. (7.15) needs to be


inserted into eq. (7.10). For the special case of elastic material (n = 1),

a:/E:

in eq. (7.10) must be identified with Young's modulus. There results the

following differential equation for Ub :


(7.16)

with
t

= (3w/8)

d n/E.

(7.17)

The solution of this ordinary differential equation is


(7.18)
Thus the characteristic time for the build-up of elastic stress
by

viscous grain boundary sliding is of the order tv

concentrations

dn/E. (The subcript, v,

indicates viscous sliding).


The friction coefficient, n, is often controlled by
around

grain-boundary

particles.

stress-directed

diffusion

It is then given by eq. (1.8). Depending on

the material parameters involved, the rise time, tv' for the elastic stress
concentration may be shorter or longer than the relaxation time, t 1 If tv is
found to be larger than tl' the elastic stress concentrations are
up.

The

same

is

true

if

diffusive

stress

never

built

relaxation occurs within times

comparable to, or shorter than, tv (Section 2.2.3).

7.2 The Triple Grain Junction in Polycrystais


In this section, the geometrical conditions at triple grain junctions are

con-

sidered more accurately than in the preceding section. This has two effects.
First, the stress singularities at triple junctions of sliding grain boundarief
are different from those at the end of an isolated sliding boundary facet. Second, grain boundary diffusion, which plays no role if an isolated grain
ary

facet

bound'

is considered, contributes to the deformation process and to stresl

relaxation in a contiguous network of grain boundaries. The relative importanci


of

grain

boundary diffusion, compared to power-law creep, can be expressed 11

terms of the diffusive length,

~,

defined in eq. (4.17). In Section

7.2.2

thl

7.2

limit

93

The Triple Junction

~/d +

will

be

considered

where diffusion is negligible compared to

creep. The opposite limiting case, in which diffusion predominates,

while

the

grains behave effectively rigidly, is described in Section 7.2.3.


Apart

from

the

consideration

of

possibility
contiguous

of

grain

networks,

boundary
instead

however,

diffusion,
of

isolated

the

grain boundary

facets, does not lead to basically new conclusions, as we shall see.


7.2.1

The triple junction in elastic material

Figure 7.2 shows a hexagonal array of grains and geometrical definitions

at

triple junction. Since the author does not know of a published solution for the
stress near a triple junction, the problem is worked out for
Appendix

B.

plane

strain

in

Starting from the equation for the Airy stress function, V4~ = 0,

one obtains the stress field near the junction in the form
(7.19)
Here gij(S) are dimensionless angular functions normalized such that gss(O) =
1. The dimensionless factor K is undetermined by the asymptotic analysis and
will be estimated from other arguments later. The exponent s is calculated

for

incompressible material in Appendix B and is determined by the relation:


(2-s) cos[(2-s)(n-2a)]

+ S

cos[s(n-2a)]

2 cosns.

(7.20)

Fig. 7.2. Hexagonal array of grains deformed in plane-strain tension.

94

7.

2a
s

180

160
0.191

120
0.449

109.5
0.500

Stress Concentrations

20
0.832

60
0.689

Table 7.1. Exponent s of the stress singularity at a triple


junction of sliding grain boundaries in elastic material.
Table 7.1 gives the solution of this equation for a few values of the angle
Note that for 2a

a.

< 109.5, the stress field has a stronger singularity than the
whereas for 2a > 109.5, the singularity is

crack-tip singularity, r- 1/2 ,


weaker.

This

difference in the degree of the singularity has consequences for

the stability of a crack nucleated at the triple junction (Riedel,


obtuse

2a > 109.5,

angles,

1976).

the attainment of the ideal strength suffices to

initiate an unstable crack Since, once the crack nucleus is formed, the
singularity

rises.

For

2a

For

stress

< 109.5, the crack is formed readily, but, due to

the drop of the stress Singularity, it does not propagate, unless

the

applied

stress is increased.
The factor K can be estimated from a finite element analysis of
grid

of

grains (a

ions pertain to nonlinear viscous rather than


Extrapolating

their

the

results,

to

elastic

material

60 0

Unpublished

calculations

of

the

of

length

that

0.29

present author indicate that

K = 0.21 for a = 45. (These calculations were done for two


planes

behavior.

which will be given in eq. (7.21) below, to the

linear case and recalling the elastic-viscous analogy one finds


for

hexagonal

60) by Lau, Argon and McClintock (1983). Their calculat-

intersecting

slip

Which is geometrically similar to a network of grain

dll2,

boundaries albeit not identical).


Numerically, the stress concentration factors at triple junctions
to

those

at

the

end

of

occurs on the boundary normal to the applied


r

10 nm

cavity

similar

stress

(8

= 0).

At

distance

from the triple pOint, which corresponds approximately to the stable

nucleus

size,

088(8=0)/o~ = 9.2

was taken as d
7.2.2

are

an isolated slip plane. The maximum tensile stress

if

= 30~m

the

applied

stress

is

concentrated

a = 60, and by a factor 22.2 if a = 45 0


in this example.

by

factor

The grain size

The triple junction in power-law creeping material

The geometrical configuration is again the array of hexagonal

grains

deformed

in plane strain shown in Fig. 7.2. The grain boundaries slide freely, while the

7.2

The Triple Junction

95

grains are nonlinear viscous as described by

E = Ban

eq. (3.6),

simplifies

to

in uniaxial tension.

The asymptotic analysis of the stresses near the triple


the

which

governing

junction

starts

from

nonlinear equation for the Airy stress function , eq. (3.19a),

with the elastic terms deleted. This problem has factorized solutions of the
form ~ r 2 - s f(6). The angular distribution f(a) and the exponent s have been
determined numerically by Lau and Argon (1977). Their

results

will

be

shown

shortly. Because of the scaling properties of stress fields in power-law material (cf. Section 3.4.2), the asymptotic stress field must have the same form as
in

linear

elastic

material, eq. (7.19), but with sand gij(S) taken from the

nonlinear analysis of Lau and Argon


functions

are

shown

in

(1977).

Their

results

for

the

angular

Fig. 7.3. For the exponent of the stress singularity

they find s = 0.231 if n = 3, s = 0.150 if n = 5, and s = 0.112 if n = 7.


The determination of the factor K in eq. (7.19) requires a
of

numerical

analysis

the whole arl'ay of grains, rather than an asymptotic analysis only. Several

authors (Crossman
McClintock,

1983),

and

Ashby,

1975,

Gharemani,

1978,

and

Lau,

Argon

and

have performed finite element analyses of hexagonal grains

with sliding boundaries. Lau et al.

(1983)

lay

particular

emphasis

on

the

behavior near the triple points. They give the following interpolation formula,
which is based on finite element calculations for n

= 3 and n = 5:

Fig. 7.3. Asymptotic angular stress distribution at a triple grain


junction with an included angle of 120 0 under plane-strain tension.

7.

96

Stress Concentrations

(1.11 - 0.017n) (1-s) (2/3)-s/f.

(7.21)

Here, f is the stress enhancement factor defined in

eq.

(1.9).

According

to

6 = O.

If

Gharemani (1978) is f - 1.2.


The maximum tensile stress occurs on the grain boundary lying along
r = 10nm

(=

size of stable cavity nucleus) is inserted into eq. (7.19), stress

concentration factors are obtained as a66(6=0)la~ = 9.2, 2.8 and 1.8 for n = 1,
3 and 5, respectively, if d = 30 ~. Thus, for power-law creeping material with
n

3, the local stress concentration at

triple

points

is

moderate.

It

is

further reduced by diffusion (Section 7.2.4).


7.2.3

Stresses during Coble creep (rigid grains)

In the stress and temperature range, where Coble creep dominates


map

of

deformation

on

an

Ashby

mechanisms, the grains are displaced as effectively rigid

blocks by the diffusive plating of atoms on grain boundaries. In

other

words,

the deformation by creep of the grains can be neglected compared to the deformation by grain boundary diffusion. This limiting case
mathematically.

is

particularly

simple

The only differential equation for the stress that needs to be

solved is the equation for grain boundary diffusion, eq. (4.4). For

two-dimen-

sional problems involving plane boundaries, this equation takes the form
(7.22)
where x is the co-ordinate along the

boundary

and

un

is

constant

(in

the

absence of grain rotation; otherwise un would be a linear function of x).


As an example, eq. (7.22) is solved for the hexagonal array of grains shown
Figs.

7.2

geometrical
rigid-body

and 7.4 for plane strain. If ~ denotes the extensional strain rate,
compatibility
displacement

and
rates

the
of

conservation

of

mass

demand

that

the

the grains have the form indicated in Fig.

7.4. Note that the strain rate in horizontal direction must be


pressible,

in

-E

for

incom-

plane-strain deformation. From the relative displacement rates, the

normal displacement rates across grain boundaries are found to be


u
u

on horizontal boundaries

n
n

(7.23)
-Ed/2

on inclined boundaries.

7.2

97

The Triple Junction

Fig. 7.4. Displacement rates of grains for incompressible plane-strain


deformation, and normal component of displacement rate at grain boundaries.

The displacement rates of the grains indicated imply that


must

slide

at

rate

Ed/l3.

inclined

boundaries

This reiterates the fact that diffusion creep

cannot progress without grain boundary sliding (Raj and Ashby, 1972).
For the constant values of
eq. (7.22)

is

un

on each of the

boundaries,

the

integration

straightforward giving the parabolic stress distributions shown

be~ow.

One of the constants of integration follows

about

the

from

the

symmetry

of

triple

an

center of each boundary segment. Further, the diffusive flux, which

is proportional to dan/dx, must guarantee the conservation of mass, so that


the

of

points

dan/dx

on

at

the horizontal boundary must be equal to twice

dan/dx on the inclined boundary. The normal stress must be continuous at the
triple points since discontinuities would cause infinite fluxes. Finally the

average of the horizontal stress component must be zero for global

equilibrium

since no stress is applied in this direction. These conditions determine all of


the constants of integration, and the solution of eq. (7.22) is found to be
[(d/3)2 -

x2 J

[x 2 /2 - d 2 /72J

on horizontal boundaries

(7.24)

on inclined boundaries

where x is measured from the center of the grain facets.


Global equilibrium in the tensile direction requires that the integral over

an

7.

98

on

horizontal

boundary

is

equal

to

am d1312

Stress Concentrations

(the

average on inclined

boundaries has already been made equal to zero), a condition which connects the
strain rate and the applied stress

(7.25)

e:

Note that this is Coble's law for


(1.5).

When

the

diffusional

creep

authors it should be noted that Raj and Ashby


engineering

of

polycrystal,

eq.

numerical factor above is compared with the pesults of other


(1972),

for

example,

use

the

shear strain and the shear stress, which implies a 3 times greater

numerical factor (108 instead of 36). Inserting

eq.

(7.25)

into

eq.

(7.24)

gives the maximum tensile stress


(7.26)
which occurs at the center of the horizontal boundaries. The triple points

are

no singular pOints during diffusion creep.


7.2.4

A combination of power-law creep and grain-boundary diffusion

In the presence of diffusion,


power-law

creep

are

the

stress

singularities

which

occur

during

removed by the plating of atoms on those portions of the

grain boundary which experience high tensile

stresses.

Argon,

Chen

and

Lau

(1981) estimate the length, A~, over which diffusion relieves the stress
concentrations caused by power-law creep. They find:
A~ ~ d (~/d)1/[1-s(n-1)/3J

where
the

(7.27)

is the characteristic diffusive length defined in eq. (4.17) and

exponent

of

is

the stress singularity at the triple point in the absence of

diffusion. In the low-stress creep regime, ~ is generally not smaller than


0.2 pm (cf. Section 4.5) and tends to be larger in creep-resistant alloys.
Taking ~ = 0.2
concentrations

n = 5 and d ~ 20 pm gives A! = 60 nm. Therefore the stress


caused by power-law creep at triple junctions are relieved by

pm~

diffusion over distances of at least 60 nm. This means that the already
ate

stress

concentration

factors

expected

for

moder-

power-law creep are further

reduced by diffusion. Inserting r = 60 nm and d = 30 pm into eq. (7.19) gives a


stress concentration factor of a:~x/o~ = 1.32 for n

5.

7.2

The Triple Junction

7.2.5

99

Relaxation of elastic stress concentrations at triple junctions by creep

The relaxation of the elastic stress

concentrations

at

triple

junctions

by

creep can be described in a manner analogous to stress relaxation at

power-law

shear cracks (Section 7.1.3) or at tensile cracks (Chapter 23). The details
the

less, the transition from the initial elastic behavior to long-time


be

of

analysis are different due to the different type of singularity. Neverthedescribed

creep

can

by the growth of a creep zone around the triple points as in the

case of cracks. The characteristic time for the transition, t 1 , must

have

the

same form as in eq. (7.13) apart from possibly n-dependent numerical factors.
7.2.6 Relaxation of elastic stress concentrations at triple points by diffusion
At triple junctions, stress relaxation additionally
diffusion.

The

occurs

by

grain-boundary

process is described by the equations of elasticity within the

grains, while grain boundaries are assumed to

slide

freely

with

the

normal

stresses controlled by diffusion according to eq. (4.4).


If stress, strain, displacement, length and time are normalized by cr oo ' croo/E,
croo/(Ed) , d and kTd 3 /(EQOD b ), respectively, the relevant equations [Hooke's law
and eq. (4.4)] and the boundary conditions contain none of the
meters

(except

for

material

para-

Poisson's ratio) nor the applied load nor the grain size.

Therefore the normalized stress and strain fields cannot depend on these

para-

meters, while stress and strain themselves can depend on them only by virtue of
the normalization. Hence, the normal stress on

the

horizontal

boundary

must

have the form


(7.28)

For time t
elastic

0, the as yet unknown function Ln

field,

must

be

compatible

with

the

while for long times, Ln must exhibit the diffusion-controlled

parabolic behavior of eq. (7.24). The natural time scale for this transition is
defined by the second argument of the function Ln' More specifically, Raj
(1975) models diffusive stress relaxation in a regular polycrystal subject to
shear.

From

his

numerical

analysis, one finds that stress redistribution is

effectively completed after the time

7.

100

Stress Concentrations

(7.29)

which we call the diffusive relaxation time.

Taking

material

(or a ferritic steel) from Appendix A and d = 30

a-iron

~m,

parameters

for

gives td = 70 h at

800 K and 16.h at 850 K.


Thus diffusive relaxation can take a considerable time. In relation
nucleation,

however,

it

which matters rather than


junction,

is

the

the

stress at small distances from the junction

relaxation

of

the

overall

field.

calculation

of

the

elastic

field

still

persists.

This

Near

allows

the

remote

for

stress field based on a boundary layer formulation of the

short-time problem: the remote loading and the grain geometry can
the

cavity

substantial relaxation is expected to occur within a short period of

time, when far away the

by

to

boundary

condition

be

replaced

that the stress field must asymptotically

KO m (d/r)s, eq. (7.19). This implies


that om and d can enter into eq. (7.28) only as the product 0mds. Hence the
the general functional form of the stress, eq. (7.28), specializes to

approach the elastic Singular field, 0ij

(7.30)
which is valid for times short compared to td as defined in eq. (7.29) and

for

distances from the triple point small compared to the grain size.
In order to determine the as yet
imately,
which
2a

it

is

occurs
109.5 0

convenient

if

the

The

unknown

included

factor

dimensionless

function

angle

between

K appearing

in

was

the

sliding

along

1/2,

boundaries

is

inverse

w rather

point
than

by

along

planar
=

configuration;

in

Fig.

7.5;

at

diffusion

wa. This simplified model can be

solved in closed form using a Fourier transformation technique. The


shown

square-root

modeled by Evans, Rice and Hirth (1980). As an approximation,

they replace the actual triple


occurs

approx-

the elastic starting solution, eq.

(7.19), is taken as 0.29. Diffusive relaxation of such an


singularity

Fn

to specialize the discussion to the case s

the triple point is Fn(O)

result

is

0.255. The maximum of Fn

occurs slightly ahead of the triple junction and exceeds Fn(O) by 19%.
Now the time can be calculated which elapses until the
junction

has

stress

at

the

triple

dropped to, say, 20 m, which is the maximum stress during steady-

state Coble creep. Setting on

20 m, S

1/2

and Fn = 0.255 in eq. (7.30) gives

7.2

The Triple Junction

0.1.

101

\ .,,- initial stress


\,
F := K sin2a
n

Fig. 7.5. Normalized stress, Fn'


vs. normalized distance from triple junction, p.

the diffusive relaxation time for the elastic


triple point

stress

concentration

near

the

(7.31 )
This is 2000 times smaller than the global relaxation time given in eq. (7.29).
For

a-iron

at

850 K, this means that the elastic stress concentration at the

triple junctions is relaxed within 29 seconds by diffusion.


short

time

compared

to

Although this is a

typical rupture lifetimes, cavities may be nucleated

during the elastic transient. Recall that the incubation time for cavity nucleation

given in eqs. (6.34), (6.35) and (6.37) is numerically of the same order

as td. Of course, elastic transients can only playa role if the piling
slip

against
up

of

the triple junction occurs faster than diffusion and creep relax

the elastic stress concentration. The rate at which grain boundary


piled

up

against

the

junction

is

sliding

is

limited by the viscosity of the boundary

provided that the applied load rises fast enough or internal slip instabilities
cause

rapid

loading

events

locally.

The

sliding, tv' was given in eq. (7.17). Taking

characteristic
the

friction

time

for viscous

coefficient

of

grain boundary, n, from eq. (1.8) and td from eq. (7.31) gives
(7.32)

7.

102

Stress Concentrations

Thus, the ratio td/tv depends only on grain size, particle size and particle
spacing. For large, closely spaced grain boundary particles, this ratio is
smaller than unity, i.e., elastic stress concentrations are relaxed before they
are built up.

7.3 Stress Concentrations at Particles on Sliding Grain Boundaries


Engineering alloys usually contain hard second-phase particles such as carbides
which

are

precipitated on grain boundaries in order to prevent grain boundary

sliding. These particles transmit the shear stresses which the

grain

boundary

would otherwise not be able to support at elevated temperatures. Creep cavities


are frequently nucleated at the particle/matrix interfaces. The
of

stress

concentrations

in

the

nucleation

process

is

possible

role

described in this

section.
7.3.1

Elastic stress concentrations at two-dimensional particles

In a first attempt to model elastic stress concentrations at

particles,

Smith

and Barnby (1967) considered an infinite grain boundary subjected to an applied


shear stress '= and containing two-dimensional (cylindrical) particles with
diameter p and spacing A.
Both the particles and the matrix are modeled as
p
isotropic elastic

materials

having

the

same

elastic

moduli.

Between

the

particles, the grain boundary is assumed to slide freely. In other words, Smith
and Barnby solve the elastic problem of an infinite periodic array of

coplanar

shear cracks of length Ap-P separated by uncracked material of length p. Such a


geometry leads to the stress intensity factor at the edge of the particles:
(7.33)
(Rooke and Cartwright, 1974), where the second form is an
for small particles, p

approximation

valid

Ap.

Cavity nucleation is assumed to occur once the stress intensity

factor

satis-

fies the energy criterion, eq. (7.8). Combining eqs. (7.33) and (7.8) gives the
applied shear stress which is necessary to nucleate cavities
the rupturing of atomic bonds:

at

particles

by

Stress Concentrations at Particles

7.3

103

(7.34)

oo

The second form is approximately valid for small particles, p


It is clear from eq. (7.34)
nucleation

that

for

widely

spaced

Ap.

particles,

the

cavity

stress becomes small. If an unusually small value (in comparison to

engineering alloys) for the particle size is assumed, p = 0.1


AP = 311m, Y = 1.5 Nlm and
s
found to be T = 93 MPa. I f
unchanged, there results Too
<II

E
p

150 GPa, the


1.5 IJffi whith

~,

together with

nucleation stress in shear is


the other parameters being

406 MPa, which is far greater than the stresses


at which nucleation is observed to occur.
If a finite, rather than an infinite grain boundary facet in a
considered,

part

polycrystal

is

of the applied shear stress is supported by the triple grain

junctions, so that the stresses on the particles are reduced. To a first approximation, the importance of the finiteness of the sliding facet can be assessed
by considering a finite number of
for

coplan~~r

shear

cracks.

Numerical

solutions

this problem are shown by Rooke and Cartwright (1974). Their Fig. 93 shows

that the number of cracks has no great effect on the stress

intensity

factor.

For large particles (p ~ Ap)' the difference in KII between infinitely many and
a finite number of cracks vanishes. Even for relatively small particles,

= 0.1

and 10 particles interspersed between 11 shear cracks, the stress


intensity factor at the central particles is reduced by only 10% compared to an

.piA p

array

of

infinitely

difference becomes

many

greater

particles. Only for extremely small particles, the


and

diverges

logarithmically

as

A Ip

goes

to

infinity. Thus, in the two-dimensional case, the finiteness of grain boundaries


generally plays no

do~inant

role. For the more realistic case of

three-dimens-

ional particles, however, the effect is greater.


7.3.2

Elastic stress concentrations at three-dimensional particles

The three-dimensional problem of more or less


grain

boundaries

in

polycrystals

equi-axed

particles

on

finite

can only be treated in an apprOXimate way.

Most of the assumptions, however, have been discussed

by

Riedel

(1984b)

and

have been shown to be justified.


The first assumption is to model the grain boundary facets as isolated,

penny-

shaped shear cracks of diameter d, i.e. we apply the shear crack model of grain

7.

104

Stress Concentrations

boundary sliding in polycrystals introduced in Section 7.1.1. This


not

model

does

describe the conditions near triple junctions accurately, but the -stresses

on particles some distance away from the triple pOints are

obtained

approxim-

ately correctly. The stress transmitted by the facet, 'b' whose origin was not
specified in Section 7.1.1, is now ascribed to the presence of hard particles,
each

of

2
which carries a shear force 'bAp'
If the particles are well separated

from one another, i.e. if p Ap , each of them can be considered as a circular


junction of diameter p connecting two elastic half spaces. This is a configuration which has been treated in the literature. According to

Westmann

(1965),

the shear force transmitted by the junction, 'bA~' is related to the relative
shear displacement across the boundary by

(2-v) (1+v) 'b

A~

The next approximation made is to distribute


discrete

particles

continuously

stress is assumed to be uniform.

(7.35)

(p E).

the

forces

transmitted

by

the

over the grain boundary. The resulting shear


This

implies

an

inconsistency,

since

the

displacement calculated from eq. (7.35) with a uniform 'b is also uniform,
whereas eq. (7.4) gives a p-dependent, i.e. non-uniform, displacement. Therefore

compatibility

of the displacements calculated from eqs. (7.4) and (7.35)

can only be achieved in the average sense that

the

averages

of

ub

must

be

equal. This requirement determines both u b and 'b:

'b

From 'b the average


calculated as

= '",

shear

,p

2
311 (2-v)
[ 1 + -6
1
1-v

stress

"10.

(7.36)

p ]-1

-pd

transmitted

by

particle,

can

be

(7.38)

Along the circumference of the particle the stress intensity factor varies and,
according

to Westmann (1965), the maximum Mode-II stress intensity factor at a

7.3

105

Stress Concentrations at Particles

circular junction is
(7.39)
Assuming that

cavity

nucleation

occurs

once

the

stress

intensity

factor

satisfies the energy criterion, eq. (7.8), one obtains a critical applied shear
stress for cavity nucleation

C7.40)

where tb was replaced by

t~

through eq. (7.36). The

second

term

in

brackets

accounts for the finiteness of the grain boundary and, correspondingly, vanishes for d
well

~ ~.

as

As eq. (7.40) above shows large and closely spaced particles,

as

small grains are beneficial in preventing cavity nucleation caused by

grain boundary sliding.


v = 0.3,

150 GPa,

As

p = 0.5

numerical

~m,

Ap = 3

example,
and

~m

assume

d = 30

~m.

Ys = 1.5 N/m, E =
Then the stress for

cavity nucleation is predicted from eq. -(-7.40) as t~


85 MPa. This is a relatively high value compared to observed nucleation stresses; (note that the
tensile nucleation stress is 2t~ if uniaxial tension is applied
sliding boundary is inclined by 45 0 to the tensile axis).
7.3.3

and

if

the

particles

are

Stresses at two-dimensional particles during power-law creep

As in the linear
calculated

ela~tic

depend

on

case, the methods by which stresses on

whether the particles are two- or three-dimensional.. In

this subsection the two-dimensional case


briefly

described

following

work

of

an

infinite

grain

boundary

is

of Lau and Argon (1977) and Lau, Argon and

McClintock (1983).
These authors consider an infinite
particles

on

an

otherwise

periodic

freely

array

sliding

of

grain

rigid,
boundary

diamond-shaped
subjected

to

plane-strain shear. Toe geometry is the same as that shown earlier in Fig. 6.2.
The

particle/matrix

interface

is assumed to transmit no shear tractions. The

matrix is modeled by Norton's creep law [eq. (3.6), which reduces to

E=

Ban in

uniaxial tension]. Elastic deformation and diffusion are neglected.


Before details of the calculations are described, it should be mentioned that a
continuum-mechanical

description

of

the

material

on

the size scale of the

7.

106

particles (typically 1
ions

~m)

Stress Concentrations

is questionable. On this scale, individual dislocat-

start to playa role. But in an average sense, the analysis appears to be

reasonable, and it will be found that it is primarily the average stress

on

particle which matters rather than local details.


For power-law creeping material, the stress field at the apex of the
in

the

sliding

grain

particles

boundary is singular. Lau and Argon (1977) analyze the

singularity starting from the

equation

for

the

Airy

stress

function

Ceq.

(3.19a)

with the elastic terms deleted]. They obtain factorized solutions having the form 0ij r- s gij(B). For the exponent, they give s = 0.192 and
s = 0.101 for n = 3 and 7, respectively, for a particle angle of ~ = 60 0 . Lau,
Argon and McClintock (1983) report values of s = 0.225 and 0.152 for n = 3

and

5, respectively, for ~ = 45 0 The particle angle ~ was defined in Fig. 6.2.


In the analysis of the singular field at the particle, a factor
termined.

remains

unde-

This factor was calculated by Lau, Argon and McClintock (1983) using

the finite element method. Based on calculations for ~ = 45 0 , p/Ap

1/3, n = 3

and 5 they infer the following formul~ for the normal stress, 0p' acting on the
particle/matrix interface near the apex of the particle:

where p is the diameter of the particle in


nonlinear

material

the

grain

boundary.

Although

assume that in plane-strain tension, Tm must be replaced by om/13 for


boundaries.

Using

in

generalization to other stress states is not simple, we

the

special

values

5 and ~

inClined

45 0 , eq. (7.41) gives a

stress concentration factor at particles of


C7.42)

In connection with cavity nucleation by


distance

from

the

apex

must

be

the

condensation

of

vacancies,

the

chosen to be of the order of the cavity

nucleus size. For r = 20nm and p = 1~, eq. (7.42) becomes op /0 m = 0.65 Ap /p.
It is interesting to compare this stress concentration occurring in the
Singular field with the average shear stress transmitted by

two-dimensional

particle, Tp = Tm(Ap/p). Obviously, the difference between the average shear


stress on a particle and the stress at a small distance from the singularity is
small

(12%

in

the

present

example).

This

means

that

the

local

stress

7.3

Stress Concentrations at Particles

concentration in the

singular

field

107

is

numerically

to the fact that particles occupy an area fraction of


ensional case), but carry the whole shear load.
7.3.4

moderate

in

power-law

with n > 3. Stress concentrations at particles are then primarily due

material

piA

[in

the

two-dim-

Stresses at three-dimensional particles during power-law creep

The three-dimensional case is modeled in close analogy to the elastic


of

analysis

Section 7.3.2; a sliding grain boundary is described as an isolated, penny-

shaped shear crack. Grain boundary sliding is restrained by circular


which

represent

the

hard

particles

between

the

grains. The junctions are

treated as if they were well separated from one another. In the


shear

forces

stress,

~b'

slides

freely

which

they

transmit

are

analysis,

the

smeared out to give an average shear

transmitted by the boundary. Outside


but

junctions

the

junctions

the

boundary

transmits normal tractions. The grains deform by power-law

creep according to eq. (3.6).


To transfer the linear elastic results to power-law viscous
place

material,

we

re-

displacement by displacement rate recalling the elastic/viscous analogy,

set v
1/2, and replace Young's modulus by ae/~e. This provides the linear
viscous solution. Finally, the generalization to arbitrary stress exponents,
n > 1, is aChieved by using the
Hutchinson

linearization

method

introduced

by

He

and

(1981). For the shear crack as a whole this has been carried out in

Section 7.1.2. The analogous calculation for a circular junction connecting two
power-law creeping half-spaces has not yet been worked out. Instead of performing such a calculation, it is assumed that the displacement rate of the
ion

has

the same n-dependence as the circular shear crack. For n = 1, the ex-

pression is fitted to Westmann's (1965) result for a junction, eq.


this

way

junct-

(7.35).

In

one obtains the sliding displacement rate across a circular junction

loaded by a shear force

~bAp:

Here, ae is the equivalent tensile stress acting near the grain


example ae = l3~b for pure shear.

boundary;

for

As in the elastic case, global compatibility demands that the average displacement

rate

calculated for the whole grain-boundary facet must be equal to that

of the single junctions. Equating Ub from eq. (7.11) with Ub

from

eq.

(7.43)

7.

108

Stress Concentrations

gives the average stress transmitted by the facet:


Lb

Lm 1 [1

+ (27n/32)

A~

1 (pd)].

(7.44)

Here it was assumed that the equivalent stress near the sliding facet is
to

the

remotely

applied

a:.

equal

equivalent stress, which is a good approximation if

either Lb + Lm or Lm
Equation (7.44) is independent of n [since in eq.
(7.43) the same n-dependence was assumed as in eq. (7.11) l, and is therefore
identical with the linear elastic result given in eq. (7.37) for-v
The two-dimensional analysis showed that in
n

power-law

viscous

= 1/2.

material

with

> 3 the local stress concentration beyond the average shear stress acting on

the particle is usually negligible. Therefore a good estimate for


stress

the

maximum

acting on a particle is provided by the average shear stress carried by

the particle:

(7.45)

Just as in the elastic


particles

lead

to

case,
small

small
stress

grains

and

concentrations

coarse

and

and

should

beneficial in preventing cavity nucleation as far as stress

closely

spaced

therefore

be

concentrations

by

grain boundary sliding are crucial for nucleation.


It should be noted that the stress concentration

factor Lp/Lm given by eq.


not very large: if O/p)2 is small, Lp/Lm is small in any
case; if O/p)2 is large, the denomi,nator becomes also large except when the
grain size is also very large. Thus in commercial alloys, the stress
(7.45)

is

usually

concentration factor at particles can hardly exceed a value of 10 or 20.


7.3.5 Diffusion and creep around particles during power-law creep of the grains
If both,

power-law

creep

and

grain-boundary

diffusion,

take

place

in

polycrystal, diffusion is the dominant transport mechanism over short distances


while power-law creep dominates on a larger scale. This
characteristic

is

expressed

the

diffusion length 1, defined in eq. (4.17). Here we consider the

case where power-law creep determines the deformation behavior of


i.e.

by

the

grains,

this section refers to the power-law creep regime on an Ashby deformation

map. On the scale of the particles, however, grain-boundary diffusion is

taken

7.3

109

Stress Concentrations at Particles

into account in addition to creep.


Sliding of a particle-strengthened grain boundary in the presence of
and

creep

diffusive

accommodation is described by adding the sliding rates due to creep

flow Ceq. (7.43)] and due to diffusion around the

particles

[eqs.

(1.6)

and

(1.8)]. Such an additive superposition can be only an approximation, but it reproduces the limiting cases of creep-controlled or diffusion-controlled sliding
if

is

very large or very small, respectively, compared to t. Otherwise the

additive superposition provides an interpolation formula between

the

limiting

cases.
Equating the sliding rate obtained in this way to the sliding rate of the grain
facet

as

whole,

eq.

(7.11),

gives the average shear stress carried by a

particle:
(7.46)

Here, the definition of the diffusive length, t, is based on the diffusion

co-

along the particle/matrix interface, 6D i , which usually differs from


oD b (Burton and Beere, 1981), and on the bulk diffusion coefficient, Dv ' which
may contribute to the diffusive flux around the particle:
efficient

Equation (7.46) is an extension of eq. (7.45) by the diffusion term


term

in

last

the denominator), which reduces the stress concentration on particles

substantially if the diffusive length is large, i.e. if the applied


relatively

(the

low.

In

this

case,

the

stress

is

applied shear stress is focussed on the

triple grain junctions, while the particles are circumvented by diffusion.


In the presence of diffusion,
particles,

the

stress

singularity

at

the

apex

of

the

which was a weak singularity anyway, is removed so that the average

shear stress carried by the particles is definitely a

good

approximation

for

approximations

and

the maximum local stress.


The problem analyzed above was so
idealizations

had

to

be

complex

that

numerous

invoked. However, most of the approximations become

exact in at least one well defined

limiting

case

(the

linearization

method

7.

110

applied
Tb

Stress Concentrations

in the shear crack model, for example, becomes exact for n

~ T~)'

particles

as

flat,

were

of

circular junctions, which are well separated from one an-

other) are intuitively expected to be appropriate. Some of


involved

1 and for

while the geometrical idealizations (for example, the description

checked

the

approximations

globally by Riedel (1984b) who applied the shear crack

model to grain boundary sliding

in

creeping

polycrystal

having

linearly

viscous grain boundaries. The closed-form results of the model agreed well with
finite element results of Gharemani (1980).
7.3.6

Stresses on particles during (free and inhibited) Coble creep

In this section, situations are considered in which the applied stress


or

enough,

the

is

low

grains are small enough, that the deformation rate of a poly-

crystal is carried by diffusive fluxes of atoms, while power-law creep is


ligible.

neg-

If diffusion occurs along grain boundaries, which it does at moderate

temperatures, this is called Coble creep (cf. Section 1.3).


If diffusion around the particles is not strongly retarded or inhibited, it
clear

that

the

stresses

at

particles

is

are relaxed since diffusion over the

particle size can easily keep pace with diffusion over the whole

grain,

which

controls the rate of Coble creep. If, however, diffusion in the particle/matrix
interface is much slower than in the grain boundary, oDi oD b , interface
diffusion may become rate controlling, the whole shear force being focussed on
2

the particles. The stress on the particles is then of the order crp/a~ = (Ap/P)
simply because the area fraction of the particles is of the order (A /p)2. This
limit is approached when oD./oD b
1

(p2/A d)2. For a more


p

detailed

discussion

of this case, see Burton and Beere (1981).

The inhibition of diffusion creep at low stresses is often ascribed to the fact
that

the plating of atoms onto the particle/matrix interfaces may be inhibited

as described in Section 1.4. In this case, the plating of atoms on clean


of

parts

the grain boundary (i.e. diffusion creep) can only continue if the thicken-

ing of the boundary is accommodated at the particles by some

other

mechanism,

for example by power-law creep.

A simple model of inhibited diffusion creep with power-law creep


near

the

inhibiting

particles

is presented next. It resembles the parallel-

fibre model proposed by Hart (1967) for


boundary

normal

to

an

accommodation

similar

applied tensile stress,

problems.
a~.

Consider

grain

The average stress trans-

7.3

Stress Concentrations at Particles

11 1

mitted by a particle is denoted by ap ' whereas outside the particles, the average stress is abo There are two requirements that determine these two stresses.
First, the displacement rate by the diffusive plating of atoms during Coble
creep,
(7.48)
must be compatible with the displacement rate by creep near the particles:
(7.49 )
Equation (7.49) must have the form shown due to the scaling properties of
er-law

viscous

pow-

fields. The factor a can be estimated by conSidering the part-

icle as a circular junction connecting the grains, just as in the case of rigid
particles

in

otherwise

freely

sliding grain boundaries (Section 7.3.4). The

linear elastic solution for a junction under

tension

(Rooke

and

Cartwright,

1974) is extrapolated to power-law viscous material by assuming the same dependence on the stress exponent n which holds for the penny-shaped crack (He and
Hutchinson, 1983). This results in a = (3~/4)'(1+3/n)-1/2. (Recently, Cocks,
1985, has proposed an alternative model in which he models the

particle

as

plane-strain punch). The second condition for the determination of ap and ab is


that they must together balance the applied stress taking into account the respective area fractions on which ap and ab act. Combining the equilibrium condition with the compatibility of eqs. (7.48) and (7.49) one obtains:
(7.50)
For arbitrary n, this equation cannot be
limiting
becomes

cases.
small

For
and

small
all

applied

stress

is

resolved

stresses,

for

We

consider

two

the second term in eq. (7.50)

concentrated

on

the

particles,

i.e. ,

In this limit, the rate of deformation is controlled by


localized creep of the grains near the particles. The strain rate, i = UId,
ala""

then

(4/~)(J..lp),

follows

as

power

function of stress,

pB[(J.. IP)2 a ]n /d At high


P
""
stresses, on the other hand, the second term in eq. (7.50) dominates compared
to the first. This generally leads to an only moderate stress concentration. In
this limit is ab
Coble

creep.

The

a"" so that the strain rate is the usual strain rate of

inhibiting

effect

power-law creep near the particles.

of

the

free

particles is overcome by rapid

7.

112

Stress Concentrations

Figure 7.6 shows the behavior of the strain rate as calculated from eq. (7.50).
(The strain rate follows from ap by = a p B ~ / d). Apparently, there is no
real threshold to diffusion creep in this model, but the steep decrease of at
low

stresses,

if

observed

experimentally,

might

well

be interpreted as a

threshold. This quasi-threshold occurs approximately where the two terms on the
left-hand side of eq. (7.50) are equal. This gives a 'threshold stress' of
n-1
aeD
As an example, eq. (7.51) is
if p2 is neglected compared to A2.
p

evaluated

for

the melting temperature with ~ = 50, n = 6.9, p = 1 ~m,


A = 3 ~m and d = 20 ~. Then eq. (7.51) gives a threshold stress of 1.3 MPa,
P
which is compatible with the order of magnitude observed in experiments (cf.
a- iron

at

half

Section 1.4).
7.3.7

Relaxation of elastic stress concentrations at particles by creep

Relaxation of the initial elastic stress concentration at particles by creep is


similar

to stress relaxation at triple grain junctions and at cracks (Sections

7.1.3 and Chapter 23). In particular, the relaxation time must be of the order
n-1 -1
t1 '" (E B 't p )

eq.(7.51)

Fig. 7.6. Strain rate in diffusion creep wnen inhibited by particles


(schematic).

(7.52)

7.3

113

Stress Concentrations at Particles

in analogy to eq. (7.13) when all numerical factors


(7.52)

it

are

neglected.

From

eq.

is apparent that local stress concentrations are relaxed the faster

the higher the average stress on the particle, ~p' is. Numerically, t1 is
usually only a minute fraction of the lifetime, and it is questionable whether
elastic stress concentrations can be built up

fast

enough

compared

to

this

relaxation time.
7.3.8

Relaxation of elastic stress concentrations at particles by diffusion

In addition to
concentrations.

creep,
As

also

one

diffusion

reduces

the

initial

elastic

stress

would expect, the global relaxation time of the whole

stress field around a particle is given

by

an

expression

analogous

to

eq.

(7.29), which was derived for stress relaxation on the scale of the grains. Now
the grain size, d, must be replaced by p and, as Koeller and Raj (1978) show in
detail, the numerical factor in the denominator is 37 instead of 130.
As in the case of the triple junction, however, it is not the global relaxation
time, but rather the relaxation time at a small distance from the stress Singularity, which is of primary interest for the cavity nucleation process. Diffusive

stress

relaxation

near

the

particle/grain boundary intersection can be

treated in close analogy to the analysis of the

triple

junction.

Again,

the

stress field has similarity solutions, eq. (7.30) with d replaced by p and with
Fn having a different functional form. Other than at the triple

junction,

the

normal stress on the interface, and therefore Fn' must now be zero at any time
t > 0 directly at th~ intersection because of the symmetry of the problem.
Figure

7.7

schematically shows the behavior of the normal stress on the part-

icle/matrix interface. At very small distances, r, from the


the

grain

boundary,

the

form

intersection

with

of the similarity solutions requires that the

stress decreases according to


(7.53)
where a is numerical factor of order unity
Similarity

which

cannot

be

determined

from

arguments alone. Now we define the diffusive relaxation time as the

time after which an as calculated from eq. (7.53)


average shear stress on the particle,

~p.

has

dropped

to,

say,

the

This leads to:


(7.54)

7.

114

Stress Concentrations

~---------r---------'

____ initial elastic singularity

diffusive flux

OL---------~--------~

p/1I2

r-

Fig. 7.7. Normal stress, an' on particle/matrix interface vs. distance from
apex, r. Diffusive relaxation of an elastic stress singularity (schematic).

If r is taken as the cavity nucleus size, typically r


time

is

second if

extremely

20 nm, this

relaxation
to

be

Since elastic stress concentrations at particles are relaxed so quickly, it

is

cD i

the

small. For a-iron at 800 K, for example, td is less than a

unknown

oD b , and p

unlikely

that

~m

interface
and a

diffusion

coefficient

is

they playa general role in cavity nucleation. Usually the rate

of loading is much slower than the rate at which elastic stress


are

relaxed

assumed

1.

by

creep

concentrations

and by diffusion. Only if a grain boundary is stressed

very rapidly, say by some sliding instability in an adjacent grain, the time to
pile

up

sliding

against

the

particles

relaxation time depending on the intrinsic


between

mayor may not be shorter than the


viscosity

of

the

grain

boundary

the particles which is, however, rarely known. For details, see Argon,

Chen and Lau (1980).

7.4 Stresses at Grain-Boundary Ledges


The author does not know of a stress analysis of ledges in sliding grain boundaries in elastic or creeping materials. However, the configuration is geometrically similar to two-dimensional particles except that now the sliding boundary
segments are offset by the ledge height h rather than separated by the particle
diameter, p. Hence, until a detailed analysis is available, it is

proposed

to

use the results derived in the preceding section for two-dimensional particles.

7.5

Summary

115

7.5 Summary of Stress Concentrations


The local stresses at triple grain junctions and

at

particles

due

to

grain

boundary sliding were found to depend on the predominant deformation mechanism,


and this in turn depends on time and on stress. For short times after load application,

the

stress distribution is elastic, while for long times, diffusion

and creep predominate. Figure 7.8 summarizes the


occur

stress

concentrations

which

in the steady state after elastic stress concentrations have been relax-

ed; 0p is the average stress focussed on a particle by grain boundary sliding


or by inhibited Coble creep. At high stresses, the polycrystal is deformed by
power-law creep, and also the particles are circumvented by creep.

The

stress

concentration factor at the particles is then given by eq. (7.45). For somewhat
lower stresses, the particles start to be circumvented
that

by

diffusion

provided

diffusion is not inhibited. As long as power-law creep still dominates on

the size scale of the grains, the stress concentration factor is then given
eq.

(7.46).

At

by

even lower stresses, one enters the regime of Coble creep. If

diffusion is inhibited by particles, the stress concentration factor

rises

to

values of the order (A /p)2 during Coble creep. On the other hand, the stresses
p

themselves are small in that range. In summary, the stress concentration

fact-

ors in the power-law creep regime of engineering alloys, are moderate. Only for
widely spaced, small particles and large grains can
factor

the

stress

concentration

exceed a value of 10. During elastic transients, the stress concentrat-

ion may be larger on a small scale due to the relatively severe elastic

stress

singularities.

(~p~z f---___... particle - irhibited


J

diffusion

eq. (7.50)

_ad
n

eq. (7.45\
-----~------''-f

-- eq. (7.46)
power-law
creep

Coble creep

Ga>

(log scale)

Fig. 7.8. Stress concentration factor at particles vs. applied stress.

8 The Role of Impurity Segregation


in Cavity Nucleation

8.1 Qualitative Observations


The presence of trace levels of impurities like sulfur, phosphorus, arsenic and
others has a strong, and often deleterious, effect on the mechanical properties
of metals and alloys. If such
material

tends

elements

segregate

to

grain

boundaries,

the

to fail by brittle intergranular fracture at low temperatures,

probably due to a reduction of the cohesive strength. Observations relating

to

this phenomenon, which is called temper embrittlement, are described in Section


8.1.1, and the theory of cohesive strength in the

presence

of

impurities

is

developed in Section 8.2.4. Impurities also influence the ductility under creep
conditions.

If

embrittlement.

they

do

so

in

an

adverse

manner

this

is

called

creep

This may be caused by a reduction of the cohesive strength, but

the precipitation

of

(Section

Not

8.1.2).

nonadherent
all

particles

elements

which

also
are

plays
present

quantities are harmful. Some elements such as carbon

or

an

important

in

alloys in small

zirconium

are

role
added

deliberately to improve an alloy's properties.


8.1.1

Grain-boundary brittleness at room temperature (temper embrittlement)

If materials which exhibit segregation of impurities to grain boundaries


low-alloy,

creep-resistant

steels)

are

(e.g.

tempered or crept in the temperature

range where segregation occurs (350 0 -600 0 C in steels), a drastic loss of ductility is observed in impact testing at room temperature. The phenomenon has long
been known as temper embrittlement. The subject has been
for

reviewed

repeatedly,

example by McMahon (1968), by Guttmann (1980) and by McMahon et al (1985).

Fracture in temper embrittled materials occurs by a rapid separation

of

grain

boundaries. It has nothing in common with intergranular cavitation, except that


both phenomena are influenced by impurity segregation.
A remarkable observation related to segregation has recently been made

by

Liu

8.1

Qualitative Observations

117

et al (1983) on the intermetallic compound Ni 3Al. Such ordered alloys have good
high-temperature mechanical properties but at room temperature they usually
fail

intergranularly

in

an

extremely

impurity level, adding 200 to 500


slightly,

ppm

fashion.

and

By controlling the

increasing

the

Ni-content

Liu et al were able to produce completely ductile Ni 3Al-allOYs. The

gain in room temperature ductility


cohesion

brittle
boron

due

to

the

segregation

is

ascribed

to

improved

grain

boundary

of boron to grain boundaries. Later it was

found that the boron-doped material exhibits brittle inter granular fracture
temperatures between 600 and

800 0

at

in oxygen-containing atmospheres (Liu, White

and Lee, 1985). This effect is probably not related to segregation and

can

be

should

be

avoided by the addition of chromium to the alloy.


For completeness, the embrittlement of grain boundaries by hydrogen
mentioned.

It

plays

a role, for example, in 'cold cracking' of welds, but at

elevated temperatures cracking


author's

knowledge.

by

atomic

hydrogen

does

not

occur

to

the

However, at elevated temperatures hydrogen can react with

carbon to form methane bubbles on grain boundaries in carbon-bearing alloys


('hydrogen attack'). This is related to cavitational failure and is described
in connection with the effects of other pressurized gas bubbles in Chapter 9.
8.1.2

Embrittlement by impurities under creep conditions

Apart from affecting the room temperature ductility, impurities can also impair
the

creep

rupture properties (creep embrittlement). Progress in understanding

the role of different impurities has been made in the past few years by examining

freshly exposed cavity surfaces and grain boundaries employing Auger spec-

troscopy and, more recently, field ion microscopy. These


ques

experimental

techni-

allow one to determine the degree of segregation of various elements. The

correlation with the creep rupture properties is rather complicated


completely clear to date. A few examples will be described next.

and not

Examples on nickel and its alloys containing traces of P, As, Sb

Sn

been

reported

by

Bieber

and

and

have

Decker (1961), White and Padgett (1982, 1983),

Schneibel et al (1983) and Thomas and Gibbons (1984). A comprehensive review of


beneficial

and

detrimental trace elements in nickel-base superalloys has been

given by Holt and Wallace (1976). Schneibel et al (1983) show that small additions of Zr restore the creep ductility of a Ni-20%Cr alloy, which was previously embrittled by trace levels of sulfur. They suspect a twofold effect, namely,
part of the zirconium preCipitates as sulfides in the matrix thereby preventing

118

8.

Impurity Effects on Nucleation

the segregation of sulfur to grain boundaries, and part of


resulting

in

reduction

of

order of magnitude. Such a reduction would have little


nucleation

stress

the

Zr

segregates

the grain boundary diffusion coefficient by an


effect

on

the

cavity

given in eq. (6.29), but diffusional cavity growth would be

retarded in proportion to 6Db . In any case, the beneficial effect of small additions of Zr and B on the creep rupture properties of nickel-base superalloys
has long been known and utilized. Oxygen has an
boundaries

in

nickel

and

embrittling

effect

on

grain

of its alloys. Bricknell and Woodford (1984)

some

discuss several causes for this; the possibility of gas bubble

formation

will

be studied in Section 9.1.


Impurities have a great effect on the creep rupture
stainless

steels

properties

of

austenitic

(White et aI, 1981; Swindeman et aI, 1981, 1983). The under-

lying mechanisms do not seem to be altogether clear. Indirect effects


an

influence

on

the

carbide

morphology

such

as

are important. Hence, the observed

interactions between impurity content and processing history are conceivable.


The greatest amount of work on the effects of impurities on creep ductility has
been performed on ferritic and bainitic steels. Generally, it is beneficial for
the creep ductility to keep the impurity level very low. In high-purity steels,
the

cavity

nucleation

rate

is

considerably reduced compared to commercial-

purity steels (see Table 5.1 in Section 5.7.1 and Hopkins, Tipler

and

Branch,

1971; and Tipler and Hopkins, 1976). However, the mechanism by which impurities
be easily identified in later

lead to enhanced 'cavity nucleation could not


studies

(e.g.

Tipler, 1980; Needham and Orr, 1980; Yu and Grabke, 1983; Jager

et aI, 1984). Often it was found that deliberately


severe

temper

addea

impurities

lead

creep ductility. Under special circumstances, the addition of phosphorus to


alloy,

which

(Takasugi

had

and

to

embrittlement at room temperature, but had little effect on the


been

Pope,

embrittled

1982,

1983),

by

sulfur,

although

an

restored the creep ductility


is

known

to

cause

temper

embrittlement. These seemingly conflicting observations on the effects of


impurities on the creep ductility of steels were reviewed and interpreted by
Pope and Wilkinson (1981), Chen, Takasugi and Pope (1983) and Pope (1983).
At least three different effects are conceivable by which impurities
the

influence

creep ductility. First, they can reduce the cohesive strength of interfac-

es. Although this is probably not a generally dominant factor in creep embrittlement,

it

plays

a role in special cases. George (1985), for example, showed

that oxide particles decohere from the iron matrix only if sulfur

is

present.

8.1

119

Qualitative Observations

This

observation

suggests

that

sulfur

segregation weakens the oxide-matrix

interface. Second, impurity segregation affects the


coefficient.

As

grain

boundary

will be shown in Section 8.2.5, this has an only small effect

on cavity nucleation, while the effect on cavity growth will


Section

be

discussed

serve

as

heat

boundaries,

effective cavity nucleation sites and thereby reduce the creep

ductility. The size and the distribution of the sulfides


the

in

11.1.5. The third effect is often the most important one; sulfur tends

to form sulfides, and, if these sulfides are preCipitated on grain


they

diffusion

treatment

depends

strongly

on

and on the presence of other elements, which leads to the

complicated behavior observed. Since this effect has been investigated primarily

in

connection

with

stress

relief

cracking, it will be described in the

following section.
8.1.3

Stress relief cracking or reheat cracking

A particularly harmful distribution of sulfides is generated by a


ment

which

is

often

zone of a weld in steels (King, 1980; Middleton, 1981; Hippsley


1982;

heat

treat-

applied to simUlate the conditions in the heat affected


et

al,

1980,

Gooch, 1981; McMahon, 1984). In the part of the heat affected zone which
say, 1100 oC, sulfides, which are normally

experiences temperatures above,

precipitated as large MnS particles within the grains are dissolved. During the
rapid cooling after the welding pass, the sulfides are

preferentially

re-pre-

cipitated as finely dispersed (Mn,Cr)S particles on austenite grain boundaries,


where they nucleate cavities in a subsequent creep
stress-relief

heat

treatment

after

welding.

test

or

even

during

the

Besides an unfavorable sulfide

distribution, welding also produces undesirable residual stresses in

the

heat

affected zone of the parent material. Therefore after welding, the structure is
reheated to some 700 0 C in order to remove the internal stresses by creep relaxation. Materials Which are susceptible to stress relief cracking cannot sustain
these strains. This susceptibility often arises from cavity

formation

at

the

finely dispersed sulfides. The coarse grain structure, which is also obtained
in that part of the heat affected zone and Which is known to lead to a reduced
ductility

at low temperatures, is probably less important for creep embrittle-

ment and for stress relief cracking. As for the effect of other impurities,
is

observed

that

it

small additions of boron, in the range of a few ppm, affect

the sulfide distribution adversely thus aggravating the problem. Cerium, on the
other

hand,

seems

to

retain

sulfur in the grains and therefore removes the

problem. Boron itself, in the presence of Ce, i.e. in


boundary

sulfides,

the

absence

of

grain-

is innocuous. Phosphorus seems to have a beneficial effect

8.

120

Impurity Effects on Nucleation

on the sulfide distribution (Takasugi and Pope, 1982, 1983).


There is yet another, even more drastic effect of impurities
ture

cracking

which

has

been

on

cracking. Under certain conditions, the cavitational mode of


replaced

by

high-tempera-

investigated in connection with stress relief


cracking

may

be

a very brittle, but slowly progressing, separation of grains. The

fracture surface is practically featureless


particular,

the

characteristic

dimples

down
of

to

scale

of

10 nm.

In

cavitational failure are absent.

Hippsley et al (1980, 1982) show that the occurrence of this brittle mode in
21/~Cr-1Mo

steel

multiaxial

states

depends
of

on the presence of S, P, Sn or Sb and is favored by

stress

and

by

medium-temperature

conditions.

To

rationalize their observations they assume that the cohesive strength is reduced sufficiently by grain boundary segregation to allow for

brittle

separation

of the grains. Indeed, Hippsley, Rauh and Bullough (1984) show by high-resolution Auger spectroscopy that sulfur is enriched near
interfacial

the

tip

of

the

brittle

crack. McMahon (1984) and Shin and McMahon (1984) propose that the

sulfur enrichment stems from grain boundary sulfides which might release sulfur
atoms quickly to the newly created surfaces of the interfacial crack.
An alternative explanation would be that segregation reduces the dihedral
tip

angle

which

facilitates

cavity

nucleation

void

by vacancy condensation and

enhances the cavity growth rate. In fact, there is evidence in Hippsley et aI's
(1982)

paper

that this could be the case. Their Fig. 9b shows a cavity at the

carbide/matrix

interface

located

at

triple

grain

junction.

As

shown

schematically in Fig. 8.1, the cavity has a concave radius of curvature against
the matrix.

For such a cavity shape,

there would be no nucleation barrier and

cavity with
negative

positive curvature
carbide - decorated
grain boundaries

Fig. 8.1. Cavity formation at a triple junction of carbide-decorated

grain boundaries. After Hippsley et al (1982).

8.1

Qualitative Observations

121

spontaneous nucleation and growth would be predicted even


stress.

The

concave

without

an

applied

cavity shape is a consequence of the small wetting angle

where the cavity surface intersects the

carbide

and

this

could

be

due

to

impurity enrichment in the interface.


It should be mentioned that creep cracKing near welds is not necessarily related

to creep embrittlement caused by sulfide precipitation. In practice, cracKs


780 0 e and

also develop in the zone which has experienced temperatures between


900 0 e.

In

this two-phase (ferrite-austenite) regime, carbides coarsen rapidly

so that the material looses its creep resistance. Therefore, large deflections,
which

can

occur,

for

example, as a consequence of thermal strains in piping

systems, are accommodated primarily in these

weak

zones.

The

material

then

fails because of the large strains imposed on it.

8.2 Theories Related to Segregation and Cohesion


8.2.1

Segregation eguilibria

The reason why substitutional impurity atoms are segregated to grain boundaries
is

that

sites

on

grain

boundaries

are energetically favorable compared to

lattice sites. This is in part a consequence of the

size

misfit

of

impurity

atoms which can be accommodated more easily in the grain boundary (Seah, 1975),
but the state of chemical bonding also plays a role (Wynblatt and Ku, 1980). At
high temperatures, entropy tends to favor a homogeneous distribution of impurity atoms with no enrichment on grain boundaries.
The first type of theory relevant for segregation effects deals with this thermodynamic

between energy and entropy. In the simplest case, it is


assumed that there is a gain in free enthalpy, - flG i , if a mole of atoms of
species i is segregated. A negat i ve flG i implies a tendency towards segregation.
The number of impurity species present is denoted by m (so that i = 1,2 m)
each

equilibrium

having a dilute molar fraction xi in the grains. The segregation free en-

thalpy is assumed to be independent of the number of atoms which


segregated,

i.e.,

there

is

no

have

already

energetic interaction between the segregated

atoms. However, there is site competition since only a limited number of sites
is available. The total coverage, Lei, cannot exceed unity, where e i is the
partial coverage of the available sites with atoms of species i, and
summation

sign.

Of

course,

it

is

the

would be conceivable that different Kinds of

8.

122

Impurity Effects on Nucleation

atoms segregate to different kinds of sites. This would complicate the formulation

slightly.

Similarly,

if impurities can be segregated in several layers,

the theory needs to be modified. Assuming, for simplicity, that only


and

one

one

type

layer of segregation sites is available, the coverage is given by the

so-called Langmuir-McLean isotherm (McLean, 1957, Guttmann and McLean, 1979):


(8.1)

-I~
J

Segregation enthalpies have been determined from Auger


phosphorus

segregating

to

electron

studies.

For

grain boundaries in iron, Erhart and Grabke (1981)

find AGb = - (34.3 + 0.0215T/K) kJ/mol, which is consistent with Hondros's


(1965) indirect measurement at 14500C, AGb = -67 kJ/mol. Moller et al (1984)
obtain AGb = -46.5 kJ/mol for a commercial steel between 400 0C and 500 oC, which
also

lies

in

the

same range. (Subscripts indicate the location where segre-

gation occurs, i.e. b


note

the

spe~ies

grain boundary and s

surface, while superscripts

de-

of atoms which segregate). For the segregation of phosphorus

to a free iron surface, Hondros's (1965) data give an estimate of AG =


s
-101 kJ/mol at 1450 0C (Asaro, 1980) which is slightly higher than the value
obtained by Hartweck (1979) at lower temperatures, AG = -(75 + 7.10-3 T/K)
s
kJ/mol. The segregation of tin in iron has been studied by Seah and Hondros
(1973) and by Seah and Lea (1975). At 550 0C they obtain AGb = -45.3 kJ/mol for
the grain boundary and AG = -76.6 kJ/mol for the free surface. Sulfur seems to
s
playa special role in that its segregation enthalpies are so large that even
at

sulfur

bulk concentrations of as little as 0.01 ppm and at temperatures up

to 850 0 C, iron surfaces are completely


1977).

An

indirect

estimate

Grabke (1978), gives AG


estimate is available.

for

like

and

with

sulfur

(Grabke

et

aI,

sulfur on iron surfaces, due to Tauber and

= -165 kJ/mol, while for

An obvious extension of eq. (8.1) would


between

covered

be

to

grain

admit

boundaries

energetic

no

such

interactions

unlike segregated atoms. This has been proposed by Guttmann

(1975) in order to explain why temper embrittlement occurs in low-alloy

steels

but does not occur in plain carbon steels even in the presence of high phosphorus levels. According to Guttmann,
phosphorus

due

elements

like

chromium

cosegregate

to an attractive interaction. Recent work of Erhard and Grabke

(1981), however, has shown that the effect of alloying elements is an


one:

In

carbon

with

steels,

indirect

carbon displaces phosphorus from grain boundaries by

simple site competition which is already included in

eq.

(8.1).

Addition

of

8.2

Theories Related to Segregation and Cohesion

chromium,

molybdenum

or

vanadium

123

drastically reduces the carbon activity by

carbide formation so that phosphorus is free to segregate to grain


where

it

causes

embrittlement.

In

other

boundaries,

cases, cosegregation by energetic

interactions apparently occurs, as Gas et al (1986) show for Ni and Sb in Fe.


8.2.2

Segregation kinetics

A second class of theories is concerned with the kinetics of segregation.

This

is important in cases like the following example. The heat treatments of steels
are usually done at relatively high temperatures, at which impurity
dissolved

in

the

grains

due

quenching, impurity atoms start to diffuse to


establish

the

new

equilibrium

associated

within

practically

Therefore, only upon

are

grain

boundaries

interesting

tempering

above

in

order

to

with the lower temperature. Below

350 0 C however, diffusion of substitutional impurities


that,

atoms

to thermodynamic equilibrium, eq. (8.1). After

times,

in

steels

segregation

350 0 C,

is

the

is

does

phenomenon

so

slow

not

occur.

of

temper

by

McLean

embrittlement observed.
Theoretically, diffusion-controlled segregation

has

been

modeled

(1957). He solved the diffusion equation for the impurity atoms, subject to the
boundary condition that the concentration at the
local

equilibrium

then found

with

the

grain

boundary

must

be

in

current grain-boundary coverage. The coverage is

increase according to

~o

(8.2)
where

is the time normalized as

thickness,

D is

the

complementary
t

= O.

9~,

error

from

function.

4 Dt (x B I 9 0)2,

is the grain boundary

coefficient of ;he impurity, xB is its molar

diffusion

fraction in the grains far


value,

the

grain

boundary

and

erfc

denotes

the

For long times, 9 saturates to the equilibrium

which can be calculated from eq. (8.1); 90 is the initial

value

at

For short times, eq. (8.2) can be approximated by the leading term of a

series expansion, which gives the square-root of time dependence:


(9-9 ) I (9 -9 )

4 (Dt/~)1/2 x B I (9_0)'

Figure 8.2 shows the grain boundary coverage at


plotted

as

(8.3)

fixed

time

of

tempering,

function of tempering temperature according to eq. (8.2), where

the temperature-dependent quantities are D and

9~.

At

high

temperatures,

the

8.

124

Impurity Effects on Nucleation

(I.,

e
s~
O'i

5
kinetic / ;
control

@
o

..Q

8<x.Vl5t'

O'i

o~--~~~~~~~~--~--~~

1,00

600

800

1000

tempering temperatureTinK ___

a predicted by eqs. (8.2) and (8.1)


after a fixed time of tempering. Impurity contents in weight %.
After Hippsley et al (1982).

Fig. 8.2. Grain boundary coverage

area coverage is low since impurity atoms are dissolved in the lattice
at

whereas

low temperature equilibrium cannot be aChieved because of low diffusivities

so that the area coverage is again low.


M61ler et al (1984) confirmed the validity of eq. (8.2) by measuring the segre
gation

kinetics

of P in a commercial steam turbine steel using Auger electron

spectroscopy. Hippsley et al (1982) also obtained good

agreement

between

the

observed and calculated segregation kinetics of phosphorus.

8.2.3

Calculation of interface energies from adsorption data

A third class of theories is strictly thermodynamic. It


that,

as

rule,

is

intuitively

clean

impurities with a strong tendency towards segregation will

reduce the grain boundary energy

strongly.

Quantitatively,

the

decrease

in

grain boundary energy due to segregation, dYb , can be calculated from the Gibbs
adsorption equation

(8.4:
which, for Simplicity, is given here
pressure;

ri

for

constant

temperature

and

constant

is the number of atoms per unit area of grain boundary and ~i is

the chemical potential of species i. The summation in eq.

(8.4)

includes

all

8.2

125

Theories Related to Segregation and Cohesion

constituents

of

the

alloy, including the matrix atoms. A brief derivation of

eq. (8.4) is given, for example, by Guttmann


(1980).

Its

meaning

is

that

if

the

(1975)

chemical

and

by Hirth

and

Rice

potential of an impurity is

increased by increasing its concentration, the grain boundary energy

decreases

in proportion to the area coverage ri.


In order to evaluate the Gibbs adsorption equation it is necessary to know
area

coverage

the

as a function of the chemical potential. As a simple example we

consider an ideal solution of an impurity with an atomic fraction xB in a solid


matrix having an atomic fraction x A = 1 - x B. In an ideal solution, the
chemical potentials of a host atom, ~A, and of an impurity atom, ~B are

given,

respectively, by

(8.5)

(8.6)
(apart from additive constants which play no role here).- The area coverages, r A
and r B, are related to the fraction of grain boundary sites occupied by host
and impurity atoms through
(8.7)

(8.8)
where Nb is the number of available sites per unit grain boundary area
is

given

byeq. (8.1) setting

a =

~aJ

and

aB

= a for a binary alloy. Inserting the

expressions for ri and ~i into the Gibbs


integrating on xB from 0 to xB leads to

adsorption

equation

(8.4)

and

(8.9)
where Ybo is the grain boundary energy of the clean material and the abbreviation e = exp(6G b /RT) is defined for local use. For strong segregants, is usually
-6Gb RT so that e can be neglected compared to unity. In this case, eq.
(8.9) simplifies to
(8.10)

8. Impurity Effects on Nucleation

126

The last form involves the additional approximation of low grain-boundary coverage, i.e., the logarithm is expanded for arguments near 1. Seah (1980) obtains
a similar result except that he writes the gas constant, R, where k belongs.
Equations (8.9) and (8.10) were derived
surface

for

grain-boundary

is the same with Yb , Ybo ' Nb and 6Gb replaced by their


surface, viz., Ys ' Yso ' Ns and 6Gs ' respectively.
The utility of theoretical formulas like those above
grain

segregation;

for

segregation a completely analogous procedure is applicable. The result


values

lies

in

for

the

the

free

fact

that

boundary energies and surface energies are difficult to measure directly

at moderate temperatures. On the other hand, segregation enthalpies have become


available

during

the

past few years [see the paragraph following eq. (8.1)].

Also the structure of the adsorbed layers has been clarified so that the number
of adsorbtion sites, Nb , is known. The distance between adsorption sites is
often equal to the size of the unit cell of the host lattice (- 0.3 nm) so that
19 2
Nb = (0.3 nm) -2 ~ 101m.
Inserting this value together with 6Gb = -50 kJ/mol
(which is characteristic for grain-boundary segregation of phosphorus in iron
and steel at 500 0 C) and Ybo = 0.9 J/m 2 into eqs. (8.9) and (8.10), leads to the
grain boundary energy plotted in Fig. 8.3 as a function of the bulk phosphorus
content xp At xP = 0.4% the grain boundary energy is found to be reduced by
some 30% compared to pure iron.

1.0

'"E
"'.......
.~

.1G=-50 k)/mo/e

09

T=773K
0.8

07

~.Q

06

/inear approximation

0.5
0

01

02

0.3

04

impurity content x P in %

--0.5

0.6

Fig. 8.3. Decrease of grain boundary energy due to equilibrium segregation


of an impurity. Numerical values are representative for P in Fe at 773 K.

B.2 Theories Related to Segregation and Cohesion

127

Similar evaluations have been performed by Yoo et al (19B5) for


alloys

and

tures

where

copper-bismuth

by Asaro (19BO) for iron-phosphorus alloys at much higher temperaexperimental

data

on

interface

energies

are

available.

The

agreement of the computed curves with the experiments is found to be fair.

B.2.4

The relevance of segregation for decohesion

In the preceding subsection the reduction of grain-boundary and

surface

ener-

gies by segregation was calculated. Now the results are applied to the cohesive
strength of interfaces. The reduction of the cohesive strength is important for
temper embrittlement and sometimes also for creep embrittlement.
In order to separate a solid along an interface
bonds

mechanical

loading

device must do the

by

the

wor~

rupturing

of

atomic

Y per unit area of newly

created surface. The magnitude of this specific work depends on whether or

not

the impurity atoms have enough time to segregate to the free surface during the
process of separation. Two limiting cases will be considered. The
is

that

of

first

tained (constant chemical potential). The other limit is that of fast


ion

limit

slow separation in which an equilibrium surface coverage is mainseparat-

which implies that each free surface just inherits half of the impurities,

which had been present on the interface before separation (constant coverage).
Hirth and Rice (19BO) analyze these limiting cases carefully. For the slow-separation limit the work of separation is

(B.ll )
where Ys is the energy of the free surface with equilibrium segregation and Yb
is the energy of the contaminated grain boundary. This is a result which one
would naively expect: the work of separation is twice the surface energy (since
two surfaces are created) minus the energy of the grain boundary which had been
present before separation occurred. Both, Ys and Yb can be calculated from eq.
(B.l0) with the mOdifJcations for Ys indicated. Insertion into eq. (B.ll) gives

For small coverages, (6 b and 6s )

1, expansion of the logarithms leads to


(8.13)

128

8.

Impurity Effects on Nucleation

where rs = Ns 6s and rb = Nb6b are the numbers of impurity atoms per unit area
of surface and grain boundary, respectively. For strongly segregating impurities, the linear approximation, eq. (8.13), is usually not accurate enough.
The loss in cohesive strength by segregation,

~(2Y) = 2Y - (2Yso -Y bo )' is shown


in Fig. 8.4 for sulfur and phosphorus in steel according to eq. (8.12) with the

material parameters given in the figure caption.

For the clean material is


2Y = 2Yso - Ybo ' which has values of typically 3 J/m 2 for metals. As the figure
shows, the decrease in cohesion, ~(2Y), may reach 30% of that value if
0.1 at-ppm

sulfur

or 0.1 at% phosphorus are added (at 800 K). Clearly, sulfur

is much more effective than phosphorus

in

embrittling

grain

boundaries.

In

steels, however, the atomic fraction of sulfur in solid solution is limited due
to the presence of manganese, which precipitates together with sulfur
The

solubility

product

as

MnS.

of Mn and S in 3.25% silicon iron is given by Ainslie

and Seybolt (1960) as


[Mn].[S]

5.25 exp(-12,818 KIT),

(8.14)

where the brackets mean concentration in weight per cent. This formula is based
on

measurements

above 1,300 K. The dashed line in Fig. 8.4a represents the so

calculated sulfur concentration in solid solution in

the

presence

of

1%

by

weight manganese. As the figure shows, the presence of 1% Mn limits the loss in
cohesive strength

by sulfur

segregation to at most

~(2Y) ~ 0.7 J/m 2 ,

nearly

0
0) Sulphur

'"E

b) Phosphorus

-0.5

"'-,
.;

---

;0.,.

-1.0

850K
0

0.05

0.1

x S, at-ppm 5 -

0.15

0.1

0.5

xP,at-% P

Fig. 8.4. Loss of cohesive strength, ~(2Y), by equilibrium segregation


of Sand P in iron predicted from eq. (8.12) with Ns = Nb= 10'9/m'.
Free enthalpies (in kJ/mol) for S: ~Gs = ~~ = - 165 (~Gb assumed),
for P: ~Gs = -(75 + 710-3 T/K), ~Gh = -(34.3 + 0.0215 T/K).

8.2

129

Theories Related to Segregation and Cohesion

independent of temperature.
The result in the fast separation (or constant coverage) limit is less

obvious

than eq. (8.11). It was given in eqs. (28) and (43) of Hirth and Rice (1980) in
general thermodynamic terms, and was evaluated for
(1980).

For

brevity,

ideal

solutions

by

Asaro

it is not reproduced here. It should only be noted that

the loss in cohesion is smaller in the fast separation limit than it is in

the

slow, equilibrium limit described above (Rice, 1976).


8.2.5

The effect of segregation on cavity nucleation by vacancy condensation

As an alternative to decohesion of interfaces, cavity nucleation can also occur


by

the

thermally

activated condensation of vacancies as described in Section

6.2. Now the interaction of impurities (or of deliberately added alloying elements)

with

the

process of vacancy condensation is considered. The nucleation

stress, which was given in eq. (6.29), varies as 0nuc y~/2. Hence, an
impurity which segregates to the surface of a vacancy cluster and reduces the
surface energy is expected to reduce the nucleation stress correspondingly.
However, the problem is complicated by the fact that the surfaces of the vacancy clusters mayor may not be covered with impurity atoms, depending on whether
the impurity atoms diffuse fast enough to
This

problem

the

fluctuating

vacancy

has been examined by Trinkaus (1983) and by Chen and Yoo (1984),

based on older work on nucleation in multi-component systems (e.g.


Stauffer,

clusters.

1976).

The

Binder

and

analysis starts from the Fokker-Planck equation for the

cluster density, which depends now on the

additional

variable

as'

the

area

coverage of the cluster surface with impurity atoms. Instead of one self-diffusion coefficient, oD b , there are now the diffusion coefficients of the host
A,
B
atoms, oDb , and of the impurity atoms, oDb The results are surprisingly
B
A
simple. Over a wide range of the ratio oDb/oDb , the free activation enthalpy
stays

within

1%

of

the free activation enthalpy of the equilibrium-coverage

case, even if the diffusivity of the


smaller

than

that

of

impurity

is

treated by Chen and Yoo (1984), the most important


rate,

exp(-~G*/kT),

is

many

orders

of

magnitude

the host atoms. This means that, at least in the model


nearly

independent

of

factor

in

the

nucleation

the mobility of the impurity

The kinetic factor in the nucleation rate must be modified such that oDb
replaced by the geometric mean, [a b OD~ (l-a b ) oD~ll/2. The dependence of

~toms.

is

the Zeldovich factor on oDB/oDA given by Chen and Yoo (1984) is neglected here.
b

8.

130

Impurity Effects on Nucleatior

Hence, in the presence of impurities, the cavity nucleation stress, eq. (6.29),
becomes

60 - (Q~

~)/(2RT)

A
B
Here, Qb and Qb are the activation energies for grain boundary diffusion of the
host and impurity atoms, respectively, and typical numerical values were
inserted in the argument of the logarithm appearing in eq. (6.29),

the

result

being insensitive to the choice of these values.


Based on eq. (8.15), the possible effects of impurities on cavity nucleation by
vacancy

condensation

can

clearly

be

demonstrated. If the surface energy is

lowered to, say, 10% of the pure material, the nucleation stress drops

to

59%

of the original value since anuc y;/2. For a-iron at 850 K this means
a
= 3,160 f v1/2 MPa, using the same numerical values as in eq. (6.30). This
nuc
is still a. rather high numerical value for the nucleation stress. In the
section on stress concentrations it was found that the local stress at
particles may be at most 20 times higher than the applied shear stress. In many
cases

the

stress

concentration

concentration

factor

of

20

will

be

much

lower.

Taking

stress

as an upper limit the applied tensile stress for

cavity nucleation must be at least a... - 316 f1/2


MPa, .where a... = 21:... was
v
assumed. In order to reconcile this with the observation that at 50 MPa
nucleation still occurs, it must be concluded that

= 0.025.

fv
~

+ ,

For

must

be

as

small

as

For the dihedral angles defined in Section 6.2.2, this means that
~ < 35 0 according to eq. (6.10). This does not appear to be impossible.

example,

the

combination

of

values

stress

concentration

factor

,= 30

10 0 ,

fv = 0.014. This allows for cavity nucleation at a...


shear

0 ,

10 0

gives

40 MPa provided that

is 20. However, no values for , and

generally available, particularly not in the presence of segregation,

so

the
~

are
that

this estimate remains somewhat unfounded.


Finally, segregation affects
thereby

the

the

cavity nucleation

grain
stress.

boundary
However,

diffusion
a

coefficient

reduction

of the grain

boundary diffusion coefficient by the relatively large factor of 60, which


occur

and
may

in iron due to phosphorus segregation (Stratmann et aI, 1983), increases

the nucleation stress by as little as 6% according to eq. (6.29).

9 Cavity Nucleation Assisted by


Internal Gas Pressure

Gases or solid phases which are precipitated in alloys may exert high pressures
on

the

surrounding

matrix.

The driving force for precipitation against that

pressure comes from the high chemical potential which the constituents
precipitate

equilibrium gas pressure, p, is


clusters,

of

the

may have in solid solution. If one assumes for the moment that the
established

within

the

subcritical

vacancy

the effect of the pressure on nucleation is exactly the same as that

of a tensile stress (Russell, 1978, Raj, 1982). This means that in the formulae
for
G +

ure

cavity

nucleation (Section 6.2) the normal stress,

G,

must be replaced by

p to account for internal pressure. The magnitude of the equilibrium pressis

calculated in Sections 9.1 to 9.3 below for a number of representative

examples. It is, however, a very difficult question whether the


pressure

~quilibrium

gas

is actually built up fast enough compared to the lifetime of the sub-

critical vacancy clusters. This kinetic aspect is addressed in Section 9.4. The
effect

of

a gas pressure on the growth of cavities, as distinct from nucleat-

ion, will be described in Section 11.1.6.

9.1 Oxygen Attack and Related Phenomena


Nickel and some nickel-base superalloys undergo considerable embrittlement when
heat treated in oxidizing environments at temperatures above 1,000oC (Bricknell
and Woodford, 1982, 1984; Pandey, Dyson and Taplin, 1984). Among
causes

for

this

the

possible

embrittlement we consider the formation of carbon-dioxide or

carbon-monoxide bubbles by the reaction of oxygen with carbon contained in


metal.

the

In subsequent creep tests, these bubbles, which are present in a typic-

ally 0.5 mm thick surface layer, lead to rapid grain boundary cavitation and to
brittle failure.
A related effect is the formation of water vapor bubbles in silver or copper by
tempering in hydrogen and subsequently in oxygen. This phenomenon was exploited

132

9.

Gas Pressure in Cavities

by Goods and Nix (1978), by Nieh and Nix (1979, 1980a) and by Stanzl, Argon and
Tschegg

(1983)

in

order

to

bypass the cavity nucleation stage and to study

cavity growth alone.


9.1.1

The equilibrium carbon-dioxide pressure in nickel

A chemical reaction is in equilibrium when the sum of the


of

the

chemical

potentials

reactants equals the chemical potential of the reaction product. It is

convenient to express the chemical potential by the fugacity, f (for gases), or


by the activity, a (for solid phases):
11

I1 g + kT In f
o

11

liS +

( 9.1)

kT In a.

(9.2)

Here, the fugacity is conventionally measured in atmospheres (1 atm

0.1 MPa) ,

and 11~ is the uhemical potential of a gas atom in the pure gas at f
1 atm. In
general, eqs. (9.1) and (9.2) can be considered as mere definitions of a and f,
and

insofar they do not contain any information. For ideal gases, however, the

fugacity of a constituent becomes equal


allows

to

pressure,

and

this

11

liS

by

definition.

For

ideal

solutions, the activity of a constituent is equal to its concentration.

The chemical potentials or


phases

partial

a calculation of the chemical potential. The activity, a, is dimension-

less. For the pure solid phase is a = 1 ~nd


dilute

its

have

equivalent

thermodynamic

functions

of

the

pure

been measured and tabulated (Landolt-Bornstein, 1961, Kubashewski

and Alcock, 1979).


A calculation of fugacities and activities is carried out now for the
carbon-dioxide

(ormation

in

nickel

and

nickel

case

of

alloys (oxygen attack). The

reaction to be studied is
(9.3)
For simpliCity, the possibility of CO-formation is disregarded. In
formation

is

fact,

CO 2 -

preferred in oxygen attack of nickel (Dyson, 1982; Bricknell and

Woodford, 1982). The equilibrium condition for the chemical potential demands
(9.4)

9.1

133

Oxygen Attack and Related Phenomena

Here, the oxygen pressure (but not the


enough

to

CO 2-pressure)

is

assumed

to

be

low

permit the use of the ideal gas approximation, f = p. Resolving eq.

(9.4) above for p(C0 2 ) and replacing f by P through f = rp, which


fugacity coefficient r, gives
(1/r)

defines

P(02) a(C) exp(-6G/RT),

the

(9.5)

where the molar free enthalpy of the reaction is given by


(9.6)
These free enthalpies are also tabulated. For CO 2-formation

according

to

eq.

(9.3), Kubashewski and Alcock (1979) give


6G

396

0.84.10- 3 T/K) kJ/mol.

(9.7)

In order to obtain the gas pressure of CO 2 it remains to determine the pre-exponential factors in eq. (9.5), a(C), P(02) and r. The activity of carbon in
pure nickel is calculated by Dyson (1982) in the following way. According to
Hansen (1958), the solubility limit of carbon in pure nickel is
(9.8)

x~ = 0.5 exp[-39.6(kJ/mol)/RT].

Since no carbides are formed in pure nickel, the activity of carbon


at

the

solubility

limit

must

be

it

nickel

equal to that of pure solid carbon, i.e.,

a(C) = 1. If the atomic fraction of carbon, xC ,is


limit,

in

less

than

the

solubility

is plausible (because of the dilute carbon concentrations) that the

activity varies in proportion to xC:


a(C)

2 xC exp(4,762 KIT),

(9.9)

where the gas constant was inserted numerically.


The partial pressure of oxygen in the metal usually cannot

be

assumed

to

be

equal to the outer atmospheric oxygen pressure. This would only be so if oxygen
were an inert gas. In reality, however, Ni and O2 form an oxide
low oxygen pressures. Therefore the reaction
2NiO

even

at

very

(9.10)

134

9.

Gas Pressure in Cavities

determines the oxygen pressure in the metal. Equilibrium of

the

nickel-oxygen

reaction in the presence of the pure phases Ni and NiO gives


P(02) -

(9.11)

exp(2~G/RT);

with a formation free enthalpy of one mole NiO (Dyson, 1982):


~G

= - ( 235 - 0.086 T/K ) kJ/mol.

Numerically, P(02) varies between 9.3.10- 22 Pa

at

(9.12)
2.7.10- 9 Pa

700 K and

at

1400 K according to eq. (9.11).


It remains to determine the fugacity coefficient
neither

theoretically

r, which is not a simple task,

nor experimentally. Therefore an approximation proposed

by Raj (1982) is adopted here. Based on a van der Waals description of


gas

and

on

some

additional

real

approximations, he obtains the universal result

valid approximately for all van der Waals gases:


In r

0.086 plpc
TlTc

where Pc and Tc are pressure and temperature,

(9.13)

respectively,

at

the

critical

point of the gas. For CO 2 is Pc = 7.3 MPa and Tc = 304 K according to textbooks
in physical chemistry. Raj
3.5 MPa, Tc

134 K) and for

(1982)
CH~

(pc

additionally
=

4.6 MPa, Tc

Well away from the critical point, a van der Waals

gives
=

values

for

CO

(pc

191 K).
gas

can

approximately

be

described by the following equation of state


(9.14 )
with a constant Vo. For this special case, Raj's (1982) analysis leads
alternative approximation for the fugacity coefficient:

to

the

(9.15)
With eqs. (9.7) to (9.13), the equilibrium pressure can be calculated from
(9.5).

eq.

The result cannot be resolved for p(C0 2 ) in closed form since r depends

on pressure, too. However, the equations can be solved for T, and in

this

way

9.1

Oxygen Attack and Related Phenomena

the

result

shown

in

Fig.

9.1

was

135

obtained.

Obviously,

the

ideal

gas

approximation becomes very inaccurate at CO 2 -pressures exceeding 50 to 100 MPa.


Compared

to the ideal gas approximation the van der Waals description leads to

lower pressures. For example, the

pressure

at

Woodford (1982) did their experiments, is p

1273 K,

where

Bricknell

and

700 MPa for the van der Waals gas

instead of 4,740 MPa for the ideal gas. But still, the pressure calculated
the

van

der

Waals

gas

is

far

greater than the stresses which are usually

applied in creep rupture tests. The gas pressure should therefore play
in

cavity,

or

bubble,

for

nucleation

role

provided that the equilibrium pressure is

built up fast enough to stabilize subcritical vacancy clusters.


9.1.2

Carbon-oxides in nickel-chromium alloys

Dyson (1982) also examined the equilibrium

CO 2 -pressure

in

Ni-20Cr

alloy

containing enough carbon to form Cr 2 .C.-carbides. The effect of Cr is to reduce


the activity of carbon and oxygen, since Cr has
these

elements

much

greater

by

pressure

10

to

ttran has nickel. In the example considered by Dyson, the over-

whelming effect comes from the reduction of the oxygen partial


drops

affinity

to

in the

pressure

which

20 orders of magnitude. Correspondingly, the calculated CO 2 bubbles is vanishingly

small.

At these low

---T inK
1000

1250

Q,
CII

....

500

\
\-ideal-gas

500

~
S

750

1(1)()

200
100

approx imation

eq. (9.13)

::::J

I,/)
I,/)

CII
....

50

Q,

Nickel +

130 wt. -ppm carbon

0'"
lJ

20

\
~

(x c =5.35-1O-')

10~--~------~----~--~

1.5

1000K/T - - -

Fig. 9.1. Equilibrium pressure in CO 2 -bubbles in


carbon-bearing nickel exposed to oxygen.

oxygen

partial

9.

136

Gas Pressure in Cavities

pressures, the formation of carbon-monoxide is preferred.


pressure

The

calculated

CO-

is much greater than that of CO 2 , but it is still below 1 MPa. Hence,

cavity nucleation is not substantially assisted by CO 2 - nor by CO-formation.


Contrary to this prediction, exposure of nickel-base superalloys to high-tempeair (1150 oC) causes some form of cavitational damage near the surface

rature

(Pandey, Dyson and Taplin, 1984). Pandey, Taplin and Mukherjee (1984) report
reduction

ductility of a super alloy if tested in air as compared to vacuum


in the temperature range 550 to 880 oc. These authors ascribe the embrittlement

to

in

CO/C0 2 -bubble

formation

on

grain boundaries near the surface. If this is

correct the oxygen partial pressure in the metal must be far greater
equilibrium

pressure,

maybe

than

the

as a result of Cr-depletion in a zone near grain

boundaries, or because carbon is much more mobile than chromium and reacts with
oxygen

before Cr 2 0 a can be formed (Dyson and Hondros, 1984). Another mechanism

which can account for sub-surface cavity formation during oxidation is


bed

by

Weber

and Gilman (1984): elements with a high affinity to oxygen, for

example chromIum, aluminum or rare earth metals, diffuse to the


they

surface

where

form an oxide scale. Nickel, which has a low affinity to oxygen, diffuses

into the material but, being a slower diffuser, cannot quite balance
ward

descri-

flux

of

the

out-

matter. This leads to a supersaturation of vacancies, which can

condensate into cavities.

9.2 Hydrogen Attack


A classical example for pressure-induced bubble formation

is

what

is

called

hydrogen

attack. If a carbon steel is exposed to high-pressure hydrogen (a few


MPa) in the temperature range 300 0 C to 500 oC, hydrogen diffuses into the metal

and reacts with carbon to form methane. This reaction provides the pressure for
the nucleation of methane bubbles, preferentially on grain boundaries. After
certain

time,

the

growing

bubbles interlink to form grain-boundary fissures

which deteriorate the strength and ductility of the material drastically.


Review-type papers on

hydrogen

example,

(1976,

by

Shewmon

attack
1985)

and

in

steels

have

been

prepared,

for

by Shih and Johnson (1982). Hydrogen

attack in nickel was described by Mancuso and Li (1979).

It

should

be

noted

that these papers place their emphasis on bubble growth rather than nucleation,
since growth seems to be rate controlling. Therefore the subject will be
up again in connection with cavity growth in Section 11.1.6.

taken

9.2

137

Hydrogen Attack

Hydrogen attack has been of concern primarily to the


also

to

other

industrial

processes

petroleum

industry,

but

which involve high-pressure hydrogen at

elevated temperatures. In the early days of the Haber-Bosch process for ammonia
synthesis, failures occurred which were caused by hydrogen attack. This problem
was avoided later by fabricating the pressure vessel from a low-strength
shell,

tent, jacketed in an outer shell, which had higher strength due to


content.

its

carbon

The outer shell was intentionally perforated by drilled holes so that

the hydrogen which diffused through the inner shell could escape and no
ure

inner

which was not susceptible to hydrogen attack due to its low carbon con-

could

build

press-

up in contact with the carbon steel. Nowadays the problem is

handled by employing low-alloy steels, which are less susceptible

to

hydrogen

attack than carbon steels for reasons which will become clear shortly.
The relevant reaction for hydrogen attack in a carbon steel is
Fe.C

2 H2

3 Fe

The pertinent equilibrium pressure of the


Shewmon

(9.16)

+ CH~.

CH~-bubbles

has

been

calculated

by

(1976) using the ideal-gas approximation, and by McKimpson and Shewmon

(1981) using a van der Waals description. Figure 9.2 shows the results. For
ideal gas,

one

would

predict p(CH~)

[p(H 2 )]2 (the

dashed lines),

an

but the

2000

1000

&
~
5

---

~~

100
50
30
20
10
03

as

2 3

p~)~M~

~~

Fig. 9.2. Hydrogen attack. Equilibrium CH~-pressure in iron containing carbon


in equilibrium with Fe.C (from Kimpson and Shewmon, 1981) and in 21/~Cr-1Mo
steel (from Parthasarathy et aI, 1985). Dashed lines: ideal-gas approximation.

9.

138

deviations

Gas Pressure in Cavities

from ideality are substantial. Numerically, the equilibrium methane

pressures are large enough to be relevant for cavity nucleation.


In low-alloy steels containing Cr, Mo or V, the carbon activity
by

the

equilibrium

with

is

determined

carbides of these elements. This reduces the carbon

activity, and therefore the methane pressure, compared to the sole presence
FesC.

Parthasarathy,

Lopez

and

Shewmon

analysis for the special case of a


in Fig. 9.2.

(1985)

21/~Cr-1Mo

of

have carried out a detailed

steel. Their result

is

included

9.3 Helium Embrittlement


Another problem which arises from gas bubble formation is helium embrittlement.
Core

materials

of

fast

breeder reactors and, even more so, of future fusion

reactors, are exposed to high fluxes of fast neutrons. Several

of

the

atomic

nuclei contained in common structural materials (for example, saNi or SB) react
with fast neutrons and subsequently
These

(n,a)-reactions

emit

a-particles,

i.e.,

helium

exceed the solubility of He. At elevated temperatures (around 600 0 C in


less

nuclei.

produce helium concentrations in the metal which by far


stain-

steels), helium is therefore precipitated into small bubbles, preferenti-

ally at grain boundaries. This leads to a severe loss of ductility.

It

should

be mentioned that this is not the only form of material damage caused by irradiation. If helium embrittlement is to be studied in isolation, special

techni-

ques of helium injection are used (Braski et aI, 1979).


The driving force for helium bubble formation is not the potential of a chemical

reaction as in the previous examples, but rather a supersaturation which is

forced into the material by irradiation. The fugacity of helium is

related

to

the supersaturation according to


( 9.17)

in atm

by an argument analogous to that leading to eq. (9.9). Here xHe is


fraction

of

He,

the

atomic

and

x*
He is the solubility limit of He at 1 atm. Since the
solubility limit of He in iron is far less than 1 ppm and He concentrations of

several

hundred

ppm

are

easily

achievable

by

implantation, fugacities of

several thousand atmospheres are possible. From the critical data Pc = 2.26 atm
and

Tc

5.3 K,

and

from eq. (9.13) the fugacity coefficient is estimated as

9.3

Helium Embrittlement

139

r = 2.7 if P = 5,000 atm ( = 500 MPa) and T


pressures

are

of

the

1000 K.

Thus

the

equilibrium

order of a few hundred MPa, which may be sufficient to

nucleate bubbles.

9.4 Kinetic Aspects


Whether or not the equilibrium gas pressure is actually built up
critical

in

the

sub-

cavity nuclei depends on how fast the constituents of the gas diffuse

to the nucleation site and how fast the chemical reaction takes place. The
precipitation

of

vacancies

and

gas

atoms

(1978), Parker and Russell (1981) and Trinkaus


Planck

equation.

co-

has been investigated by Russell


(1983)

based

on

the

Fokker-

The formulation by Trinkaus includes both, the precipitation

of an inert gas, like helium, and the precipitation of a reaction product, like
methane.

However,

the

application

of

Trinkaus'

rather complicated. His own application to the


stops

theory to special cases is

nucleation

of

helium

bubbles

at the calculation of nucleation enthalpies, whereas the pre-exponential

factors

of

evaluation

the

nucleation

rate

are

not

explicitly

derived.

A complete

of the nucleation rate like that by Chen and Yoo (1984) for segreg-

ation-assisted nucleation is lacking.


A theoretical prediction by Russel (1978) is worth mentioning
that,

here.

He

finds

at least for the kinetic laws he uses, there exist metastable bubbles in

the subcritical size range. These bubbles are stable against small fluctuations
in

size

and in gas

~ontent.

Large fluctuations, of course, may take them into

the supercritical regime where they grow deterministically.


There is yet another complicating kinetic aspect of bubble formation
like

oxygen

in

cases

attack. It is possible that oxygen from the atmosphere and carbon

contained in the metal react at the specimen surface rather than at CO 2 -bubbles
formed

within the material. This happens if the outward diffusion of carbon is

much faster than the inward diffusion of oxygen. In iron and nickel for
example, Bohnenkamp and Engell (1962) find that the oxydation of carbon occurs
primarily on the surface of the

metal

with

the

effect

that

the

metal

decarburized near the surface without internal gas bubbles being formed.

is

10 Internal Stresses Due to the Precipitation of Solid


Phases and Thermal Expansion

The precipitation of solid second phases is often accompanied by an increase in


specific

volume.

If the matrix is mechanically soft, the excess volume can be

accommodated readily, while in a rigid matrix high stresses are built up around
the

particle.

Lopez

and

Shewmon

(1983) speculate that these stresses might

initiate cavities. They were led to their assumption by the occurrence of small
cavities

at

carbides in a bainitic steel directly after tempering, before the

material was exposed to creep or any other treatment which is known to initiate
cavities. They ascribe the cavities to the stresses caused by the precipitation
of carbides during tempering.
Lopez and Shewmon's idea will be examined in
principles

of

the

light

of

an

analysis

the

which are due to Hillert (1957, 1976). The analysis consists of

two steps. First the diffusion of carbon to a carbide is considered the carbide
being

under as yet unspecified, hydrostatic pressure, p. Second, accommodation

mechanisms are examined which, together with the influx

of

carbon,

determine

the pressure.
Mechanically similar is the effect

of

differential

thermal

strains

of

the

matrix and of particles, a possibility which is considered in Section 10.5. For


example, certain processing histories lead to a decohesion of manganese sulfide
particles from the iron matrix in steels.

10.1 The Flux of Carbon to the Carbide


The chemical potential of a carbon atom dissolved in the iron matrix is
(10.1)

where xC is the atomic

fraction

of

carbon

in

solid

solution,

x~

is

the

10.1

The Flux of Carbon to the Carbide

equilibrium

141

value in contact with (unstressed) Fe,C and Po is its potential in

unstressed Fe,C. If the carbide is under hydrostatic pressure, p, the

transfer

of a carbon atom from the matrix to the stressed carbide requires the work p6Q,
where 6Q is the increase in volume per carbon atom reacting with iron

to

form

Fe,C. According to Lopez and Shewmon (1983), the increase in specific volume by
Fe,C-formation is 10%, i.e. the increase per carbon atom is 6Q = 0.3 Q =
3.5'10-30 m3 , since each carbon atom increases the atomic volume of three iron
atoms. The chemical potential of carbon in the stressed carbide is
(10.2)
where we assume that the pressure dependence of Po is
p6Q, which is usually justified in solids.

negligible

compared

The difference in chemical potential between the supersaturated matrix and


stressed

carbide

to

the

drives a diffusive flux of carbon towards the carbide. For a

spherical carbide of radius HC as shown in Fig. 10.1, the solutions to the then
spherically symmetric diffusion problem can be taken from textbooks on
diffusion, or can be developed directly
result,

the

from

the

diffusion

equation.

As

rate at which excess volume is generated by the chemical reaction

in the carbide is obtained as:


(10.3)

outward
creep

~ \U

inward diffusion
of carbon
/

~B8

/'

FeJ C

'"
\

'"

outward
diffusion of
iron along
grain boundary

Fig. 10.1. Spherical grain boundary carbide, whose excess volume is


accommodated by creep of the matrix and by grain boundary diffusion.

10.

142

Stresses by the Precipitation of Solid Phases

where the term in brackets is proportional to the difference in chemical potential

far

from the carbide and in the carbide. This increase in volume will be

equated to that provided by several accommodation mechanisms (elastic deformation, creep and diffusion) to determine the pressure p.
Before the accommodation processes are studied, it should

be

noted

that

the

pressure is limited thermodynamically to the maximum value:


(10.4 ;

If the pressure grew greater than that, precipitation would stop and
reversed

according

to

eq.

(10.3).

This

upper

accommodation mechanism. Its numerical value

would

be

limit is independent of the

depends

on

the

supersaturation

xC/x~ which may be high directly after quenching from the austenitic regime, in
which a great amount of carbon is

soluble.

During

tempering

of

bainitic

steel, when carbides are already present and primarily undergo Ostwald ripening
(i.e., growth of larger carbides at the expense of smaller ones), the supersaturation

is

likely to be small. At 800 K, supersaturations of XC/X~

and 17.5% would be necessary to


respectively,

according

to

generate

eq.

(10.4).

pressures

of

50 MPa

and

1 = 1.61
500 MPa,

These are upper bounds which can be

reduced by accommodation processes.

10.2 Elastic Accommodation


Consider a spherical Fe 3 C-particle

An/(3n)

compared

to

whose

specific

volume

has

increased

the surrounding matrix, which is regarded as an i finite

body. The misfit of the particle is accommodated by elastic deformation of


particle

the

and the matrix, which are assumed to have the same elastic moduli. In

such spherically symmetric problems, one favorably uses spherical


r,~,e,

by

centered

at

the

component is the radial one, ur ' while


components arr and
related to ur by

co-ordinates

center of the sphere. The only non-zero displacement

aee =

and

a~~,

the
the

stress

tensor

has

the

non-zero

components of the strain tensor are

(10.5
The only independent
problems demands:

variable

is

r.

Equilibrium

in

spherically

symmetric

10.2

143

Elastic Accommodation

o.

(10.6)

Combining eqs. (10.5) and (10.6) with Hooke's law, and eliminating

stress

and

strain, leads to an equation for ur valid for spherical symmetry:

a ( a~
ar ar

+ 2

~ )

(10.7)

This ordinary differential equation is readily solved giving the two

solutions

-2

ur r and ur r
Within the particle, the singular solution can make no
contribution, because ur must be finite at the origin, while outside the
particle, the first of these solutions must have a vanishing coefficient, since
otherwise the stress at infinity would not vanish. Thus, the

solution

of

eq.

(10.7) for the spherical particle must have the form


inside the particle
(10.8)
outside the particle,
where a 1 and a 2 are constants of integration. They are obtainable from the
following two conditions. First, the misfit of the particle must be balanced by
the displacements at r

HC:

(10.9)
Second, the normal stress orr on the interface must be continuous, a
which

condition

is evaluated using Hooke's law and eq. (10.5). Thus, one finally obtains

the pressure exerted by the elastic matrix on the particle

- rr

2 E

lln

(the stress state in the particle being hydrostatic


calculated

pressure

is

greater

(10.10)

3 (1-\1) 9n

compression).

If

the

so

than the thermodynamically possible pressure

given in eq. (10.4), precipitation cannot take place unless other accommodation
mechanisms

become active. In relation to cavity nucleation, the hoop stress in

the matrix is relevant, which turns out to be


(10.11)

144

Stresses by the Precipitation of Solid Phases

10.

10.3 Accommodation by Power-Law Creep


Now a spherical carbide is assumed to
described

by

eq.

which

(3.6) ,

analysis of stress and strain


Equations

(10.7)

and

grow

parallels
are

(10.8)

in
to

reduces

the

replaced

creeping
. a Bonpower-law
in uniaxial tension.

matri~

e;

elastic
by

the

case

described

requirement

The

above.

that

the

displacement rate in incompressible flow must decay as l/r2:

(10.12)

ur = a.3 I r

Here a.3 is a constant, which follows from the condition that the normal stress
on the interface be continuous, i.e. orr = -po This condition is evaluated
using the creep law, eq. (3.6), and leads to a volume growth rate of the
particle by creep of the matrix
(10.13)

The pressure in the particle is now defined by the requirement that the
growth

volume

rates given in eqs. (10.3) and (10.13) must be equal. This leads to the

implicit relation for p:


(10.14)

The numerator represents the thermodynamically admissible


denominator

describes

the

reduction

of

pressure

while

the

the pressure due to yielding of the

matrix by creep flow. Creep becomes effective in relaxing the pressure when the
second

term

in

the

denominator

becomes comparable to unity. As an example,

consider the following numerical values which are typical


tempering

temperature

of

Ashby, 1977, for a.-iron), x


second

term

pressure

is

xC/x* - 1

in

the

631 MPa.

39%

according

for

accommodate

steel

at

700 C: B = 710
MPa Is, n = 6.9 (from Frost and
C
-10 2
= 10 -3 ,DC
410
mis, RC = 0.1 ~m. Then the

denominator in eq. (10.14) becomes equal to unity if the


supersaturation
This
corresponds
to
a
of
to eq. (10.14). Therefore, only for relatively high

supersaturations and correspondingly high pressures, creep is


accommodation

-15-n

fast

enough

to

the excess volume effectively. For smaller supersaturations, creep


controls

the

rate

of

precipitation,

accommodation, which is described next, is faster.

unless

diffusi ve

10.3

Accommodation by Power-Law Creep

145

In relation to cavity nucleation it is important to note that the

hoop

stress

around the particle has the form:


aee

~ p

(RC / r ) 3/n [ 3/ ( 2n ) -1 ]

(10.15)

Interestingly, the hoop stress is negative (compressive) if n > 1.5,


usually

which

is

the case. Therefore, if creep accommodation prevails, carbide precipi-

tation does not lead to tensile stresses on

the

grain

boundary,

and

cavity

nucleation is not expected to occur as a result of carbide growth.

10.4 Accommodation by Grain Boundary Diffusion


The excess volume of the carbide can also be accommodated by diffusion of

iron

atoms away from the carbide. Here we are interested in grain boundary carbides.
Then, at intermediate temperatures, the
primarily

along

diffusive

transport

of

iron

occurs

grain boundaries. The driving force is the pressure p exerted

by the carbide on the grains. This is exactly the same problem as that of a
cavity growing by grain boundary diffusion under an internal gas pressure. The
analysis will be presented later (in Section

11.1)

and

the

result

for

the

growth rate is given in eq. (11.11):


(10.16)
with the abbreviation,
(10.17)
2

where OOc is defined as OOc = (2RC/AC) , and AC is the carbide spacing


grain boundary. A possible sintering stress due to the energy

in
of

the
the

carbide/matrix interface was omitted for simplicity.


Equating the growth

r~tes

from eqs. (10.16) and (10.3) gives

the

pressure

in

the carbide:

(kT/AO) In (xC/x~)
1

(10.18 )

2 (O/AO)2 6D b / (q RC xC DC) .

The numerator is the thermodynamically possible

maximum

pressure,

while

the

10.

146

denominator

describes

the

Stresses by the Precipitation of Solid Phases

correction

due

to diffusive accommodation. As an

example, we again consider a steel tempered at 700 0 e with the numerical values
used in the preceding subsection and, additionally, oD b = 5.3'10 -22 m3 /s,
we = 0.1, i.e., q = 2. Then the second term in the denominator is 0.15, i.e.,
grain

boundary

diffusion reduces the maximum pressure by 15% in this example.

This also means that grain boundary diffusion of iron away


limits

the

carbide

growth

rate

to

from

the

particle

15% of the value which it had if carbon

diffusion were rate controlling and accommodation occurred readily.


In relation to cavity nucleation, it is important to realize how the stress
distributed

in

the

vicinity

of

the

growing

potential of iron atoms must be continuous across


stress

on

the

grain

boundary

directly

on = -p, which would prevent, rather

carbide.
the

Since

interface,

is

the chemical
the

normal

at the particle must be compressive

than

promote,

cavity

nucleation

at

growing carbide.
In conclusion, cavity nucleation at growing carbides is unlikely to occur under
all conditions examined in this section, except when elastic deformation is the
only possible accommodation mechanism. Only in the latter case
exerted

by

is

the

stress

the growing particle on the adjacent grain boundary tensile. There

is, however, no obvious reason why grain boundary diffusion or creep should not
occur

10.5 Decohesion of Particles by Thermal Expansion


During their processing
contractions.

Since

history,

materials

undergo

thermal

expansions

and

the thermal-expansion coefficient of particles is usually

different from that of the matrix, stresses will develop around the

particles,

and these might cause decohesion.


If Aa = a p - am is the difference between the linear expansion coefficients
the particle, ~p'
misfit develops is

and

the

matrix,

~,

of

the rate at which the thermal volume

(10.19)
where

is the rate of temperature change.

10.5

Decohesion of Particles by Thermal Expansion

147

The analogy to the growth of a misfitting preCipitate is obvious:


above

replaces

eq.

(10.19)

eq. (10.3), while the accommodation processes remain the same.

Thus for elastic accommodation, the pressure in the particle is found to be

2 E tJ.a tJ.T
3 (1-\)

(10.20)

where tJ.T is the difference between the current temperature and the

temperature

at which the misfit is zero.


For creep accommodation, the pressure in the particle is found by
from eq. (10.13) and

(2n/3) 12 tJ.a

t /

the

matrix

(10.21)

Bll/n sign(tJ.at).

If the pressure is positive, both the radial and the


in

~c

from eq. (10.19) which gives

tJ.~

= - arr

equating

circumferential

stresses

are compressive provided that n > 3/2 as shown in eq. (10.15),

and cavity nucleation is not expected to occur. For negative pressure, however,
the stresses are tensile, which might cause decohesion of the particle/matrix
interface or of the adjacent grain boundary.

10.6 Grain-Boundary Decohesion by Thermal-Expansion Anisotropy


If the crystal lattice has lower than

cubic

anisotropic

between

with

thermal-expansion
accommodated

difference

coefficients.

2tJ.a
If

the

symmetry,
the

resulting

and

expansion
the

is

minimum

incompatibilities

are

elastically in a polycrystalline aggregate, stresses of the order

EtJ.atJ.T arise. If the grain boundaries are able to slide,


are

thermal

maximum

concentrated

additionally

at

triple

grain

Evans, Rice and Hirth (1980) have investigated this


cavitation in ceramics.

the

thermal

stresses

junctions. Evans (1978) and


problem

with

respect

to

11 Diffusive Cavity Growth

Cavities in solids can grow by several mechanisms. The basic


under

creep

rupture

conditions

growth

mechanism

is probably the stress-directed diffusion of

atoms away from the cavity into the grain boundaries, where they can be deposited

(Hull and Rimmer, 1959). This process is described for rapid and slow sur-

face diffusion separately in Sections 11.1 and 11.2, respectively.

Rapid

sur-

face diffusion implies that a growing cavity preserves its equilibrium lenticular shape, while for slow surface diffusion, the drain

of

atoms

through

the

void tip by grain boundary diffusion leads to flat, crack-like cavities.


It will be found, however, that neither of the diffusive growth
for

an

models

allows

understanding of the creep rupture behavior of real materials as it is

commonly observed. Cavity growth rates are usually grossly


correspondingly,

overestimated

and,

rupture lifetimes of engineering materials are underestimated

by the diffusive growth models.


In the past few years, it has been recognized that the material
cavitating

grain

surrounding

boundary facet may exert a constraint on cavity growth rates

(Dyson, 1976, 1979). Indeed this constraint does reduce the cavity growth rates
to

level

which

is

compatible

with observed ones, but it is difficult to

understand how constrained growth can control the rupture lifetime, as will

be

pointed out in Section 12.3.


In Chapter 13, the effects of a possible inhibition of grain boundary diffusion
are examined. Whereas the conventional treatments of diffusional cavity growth
anticipate that atomic vacancies are always and everywhere available
thermal

equilibrium

concentration,

models

of

inhibited

constrained) diffusional growth admit that vacancies may


stress-directed

diffusion.

be

(as

at

their

distinct from

exhausted

during

Then the rate of vacancy generation at appropriate

sources may control the rate of cavity growth.

11.1

149

Diffusive Growth of Equilibrium Cavities

The growth of cavities by creep flow of the surrounding material


in

Section

14.1

(Hancock,

1976),

although

is

described

this mechanism is probably less

important under typical low-stress creep rupture conditions. Another mechanism


by deformation is cavity growth by grain boundary sliding (Section 14.2; Evans,
1971). Diffusive growth can interact with growth by creep in

an

approximately

additive way (Section 15.1; Beere and Speight, 1978). Also elastic deformation
can enhance diffusive cavity growth (Section 15.2).
Rupture lifetimes are calculated in the chapters on various cavity growth mechanisms assuming that cavities nucleate readily at the beginning of the test. If
nucleation occurs continuously, the calculation of lifetimes is a more complicated task (Chapter 16). The theoretical results will be compared with measured
rupture lifetimes of various materials including commercial alloys. Chapter

17

presents a summary of cavitational failure mechanisms, while Chapter 18 applies


the models to creep-fatigue conditions.

11.1 Diffusional Growth of Lens-Shaped (Equilibrium) Cavities


A cavity on a grain boundary can grow by the diffusive removal
the

edge

of

the

cavity

of

on

from

and deposition of these atoms on the adjacent grain

boundary. The driving force comes from the mechanical work done by
stress

atoms
the

normal

a Joundary if atoms are deposited. In other words, cavities grow by

accepting vacancies generated in the grain boundary.


In this section, several assumptions are made. First, the grains are considered
as being rigid so that there is no contribution to cavity growth by creep of
the grains. Second,
spherical-caps

shape

the

cavities

are

assumed

with a void tip angle

2~

to

have

their

equilibrium,

as shown in Fig. 11.1. Thirdly,

it is assumed that there is no depletion of vacancies in the grain boundary


the

cavities

grow.

as

Vacancies that have been absorbed by cavities are readily

replaced by the operation of vacancy sources. This is the same assumption on


which the usual treatment of diffusional creep is based (Nabarro, 1948,
Herring, 1950, Coble, 1963). Then the diffusion of atoms in the grain
is

governed

by

eq.

boundary

(4.4), which relates the normal stress distribution on a

grain boundary to the normal displacement rate:


(11.1)

11.

150

oj
b)

Diffusive Cavity Growth

0 0

o
J

,/0

Fig. 11.1. Equilibrium cavities on a grain boundary.


a) Section through grain boundary, b) plan view of grain boundary

In the rigid-grain approximation, un must

be

uniform

over

the

whole

grain

boundary. This.is val id as long as the diffusive length l defined in eq. (4.17)
is large compared to the cavity spacing. Cases with arbitrary l are
in

Chapter

15.

Finally,

the

average normal stress acting on the cavitating

grain boundary facet is denoted by 0b' and i t is noted that


this

stress

may

be

different

considered

from

the

applied

difference may be very large, and this is the idea of

stress
the

in
000.

polycrystal
In fact, the

constrained

growth

model to be described in Chapter 12.


11.1.1

The stress distribution between the cavities and the cavity growth rate

Equation (11.1) can be solved for two-dimensional (cylindrical) cavities, where

i-

a2 /ax 2 ,

and

for axisymmetric problems where'; = a2 /ar2 + (l/r)<l/<lr. The

cavity diameter in the grain boundary is 2R and the cavity spacing is


11.1).

The

axisymmetric

A (Fig.

case is meant to approximate the realistic case of a

distribution of more or less equi-axed cavities

in

the

grain

boundary,

al-

though, strictly speaking, the axial symmetry of the flow field is disturbed by
the arrangement of the cavities (square, hexagonal, statistical etc.). As

long

as the size-to-spacing ratio of the cavities is not too large, the aXisymmetric
approximation can be expected to work well.
The flux of atoms half-way between two cylindrical cavities must
symmetry.

Since

the

be

zero

for

flux is proportional to the stress gradient according to

eq. (4.3), this implies the boundary condition dOn/dX = 0 at

x = A/2.

In

the

11.1

151

Diffusive Growth of Equilibrium Cavities

axisymmetric

case the requirement of no net flux from one cavity to another is

modeled by requiring don/dr

0 on a circle with radius A/2. With

these

boun-

dary conditions, the solutions of eq. (11.1) are:

n
o

kT
t

21'l15D b

(x-R) (A-R-x)

(cylindrical cavity)

(11.2)

(axisymmetric cavity).

(11. 3)

The stress at the void tip, 00' is a constant of integration which follows from
the requirement that the chemical potential, ~, must be continuous, since the
transition of an atom from the cavity surface into the grain
expected

to

be

inhibited

boundary

is

not

so that local equilibrium is granted. Equating the

expressions for the chemical potential on the grain

boundary,

eq.

(4.2),

to

that on the cavity surface, eq. (4.11), gives


(11.4)
For equilibrium shapes the curvatures are geometrically related to R through:
K2

simp/R

for the axisymmetric cavity

(11.5)

for the cylindrical cavity.

(11.6)

The conservation of mass requires that the volumetric growth rate of an axisymmetric cavity is related to the displacement rate of the grains by

V=

(A/2)

un'

(11.7)

(For the cylindrical cavity only the final result for the linear cavity
rate,

R,

will

be given). The displacement rate ~

growth

of the grains is generally

unknown. It can be related to the normal stress acting on the

grain

boundary,

0b' by requiring that the average of on equals 0b for global equlibrium


2~

A/2

f rOn

where

dr

kT Ii A2
(l-w) 00 + _ _n::..:....._ q (w),
32 I'l I5D b

(11.8)

11.

152

Diffusive Cavity Growth

00 = (2R/A)2

(11 .9)

is the area fraction of cavitated grain boundary and the abbreviation


defined by
q(oo)

- 2 1noo - (3-00) (1-00).

q(oo)

is

(11.10)

Resolving eq. (11.8) for un and inserting the result into eq. (11.7) gives

the

volume growth rate ~, as a function of 0b:


~ = 811" D c5Db [ob - (1-00) 00 ]

(11.11)

kT q(oo)

The sintering stress, 00 , from eqs. (11.4) and (11.5), is 00 - 2 Ys sinw 1 R. A


stress which equals 00 just equilibrates an isolated cavity against sintering.
For greater stresses the cavity grows, for smaller ones it shrinks.
The linear growth rate, ~, is geometrically related to

the

volumetric

growth

rate of a spherical-caps shaped cavity by


(11.12)
where

~(w)

is the cavity volume divided by the volume of a sphere

with

radius

R. It is related to the function f v (w) defined in eq. (6.9) and is given by


hew) =

~ fv(~) =
411" sin W

(1+cosw)-1 - (1/2) cosw


sinw

Typical numerical.values are h(70 0 )

0.61 and h(75 0 )

(11.13)

0.69. For very small

W,

is hew) = 3w/8 upon linearization.


Now the final result for the linear growth rate of an axisymmetric, equilibrium
cavity can be obtained using eqs. (11.11) and (11.12). There results
2 D c5Db [ob - 00 (1-00)]

( 11.14)

kT hew) q(oo) R2
For small, widely spaced voids (00

1) the result can be approximated by

11.1

Diffusive Growth of Equilibrium Cavities

153

(11.15)

2 kT h(W) ~n(A/4.24R) R2

For cylindrical cavities, the growth rate derived in an analogous way is:
3

cD b [ob - 00 (1-2R/A)]

(11.16)

2 kT f(W) (1 - 2R/A)3 A R

with the sintering stress for the cylindrical cavity 00

Ys sinW I R, and
(11.17)

The above results coincide with those given by Speight and Beere (1975) and
Needleman

and

Rice (1980) [except for a printing error in their definition of

h(W)J. Raj and Ashby (1975), Raj et al (1977) and


omitted

the

by

so-called

Chuang

et

al

(1979)

have

jacking effect in eq. (11.7) with the consequence that

their results must be divided by the factor 1-w or 1-2R/A for the

axisymmetric

and cylindrical void, respectively. Apart from the jacking factor, the paper of
Raj and Ashby (1975) contains a misprint which has been corrected by Raj et
(1977).

The

al

latter authors further claim that the surface stress (as distinct

from the surface energy) should be taken into account with a numerical value of
2Ys sinW I R. In this case, the factor 1-w adjoining 00 in eqs. (11.11) and
(11.14) can be deleted. Needleman and Rice (1980), however, point out that this
choice

of the surface stress is as arbitrary as the assumption of zero surface

stress implicit in our derivation. Fortunately, the


case.

In

earlier

work

(Hull

and

Rimmer,

effect

1959,

inappropriate boundary conditions have been used as

is

small

in

any

Speight and Harris, 1967)


pOinted

out

by

Raj

and

Ashby (1975). These older results are only approximately compatible with ours.
In all papers quoted in the preceding paragraph it was tacitly assumed that the
stress

0b acting on the grain boundary is equal to the applied stress, but the

reader is reminded that under many practically interesting conditions, the


material surrounding the grain boundary exerts a constraint on cavity growth
which reduces 0b compared to the applied stress.
Note that in the preceding analysis no use was made of

the

assumption

of

an

equili bri um voi d shape except in eqs. (11. 5) and C11. 12). Therefore many of the
equations can be used in the analysis of non-equilibrium void growth as well.

Diffusive Cavity Growth

11.

154

11.1.2

Rupture times by diffusive cavity growth neglecting nucleation

If the time for cavity nucleation can be neglected,

the

time

to

rupture

is

given by the time for void growth from an initial radius Ro to A/2 when
coalescence occurs. The void growth law, ~ = ~(R,A), can be integrated by
separation of the variables Rand t if the void spacing is time-independent (no
continuing nucleation). Integrating eq. (11.14) between R = Ro
between t

and

i.e.

A/2,

0 and t f (time to fracture) gives

A/2

dR

(11.18 )

t f = R H(R)

o
if the sintering stress, 0 , is neglected. Further, in the term in brackets,
0
terms of order (RO/A) 5 and (Ro/A) 7 were omitted, since they are generally very
small.
If the stress on the cavitating boundary, 0b' is

identified

tentatively

with

the applied stress om' eq. (11.18) predicts t f ~ 110 m Such a linearly inverse
stress dependence of the rupture lifetime has hardly ever been observed. The
vast

majority

of creep rupture tests yields much stronger stress dependencies

often compatible with the Monkman-Grant rule, i.e., t f


is

1/0.

An

example

the study of Pavinich and Raj (1977) on copper polycrystals. In addition to

its unexpectedly strong stress dependence, the observed rupture time is


ally

much

greater

than

gener-

predicted by eq. (11.18). The only case in which the

linearly inverse stress dependence was approximately obtained is the investigation

of

Raj

(1978a) on copper bi-crystal speCimens. Four or five of his test

specimens failed in the range in which diffusive cavity growth can be

expected

to

dominate, and they exhibited a behavior close to t f ~ 1/0. For a more


detailed interpretation of Raj'S experimental results, see Chuang et al (1979).

The failure of the simple diffusive cavity growth model in

relation

to

creep

rupture of polycrystals has led to a long-lasting debate regarding the possible


reasons. In the remaining chapters of Part II, several

possibilities

will

be

examined, including the effects of the sintering stress, of the polycrystalline


constraint and of continuous nucleation.

11.1

Diffusive Growth of Equilibrium Cavities

11.1.3

155

The effect of the sintering stress on the rupture time

In this subsection it is to be demonstrated that the sintering


cavity

stress

in

the

growth law, eq. (11.14), may increase the rupture lifetime markedly. To

see this, consider the extreme case when the nucleation process provides nuclei
of

just

critical

size. For these cavities, the sintering stress balances the

applied stress, so that the growth rate is exactly zero and


would

be

the

rupture

time

infinite. Therefore it is necessary to consider the growth of nuclei

by thermal fluctuations slightly beyond the critical size before

deterministic

growth can take over. Hence the deterministic law will be applied from the time
on when it exceeds the growth rate by fluctuations derived next.
The average growth rate of a vacancy cluster by thermal fluctuations
by

n(B-B'),

~ =

where

is

given

Band B' are the vacancy absorption and emission rate,

respectively, defined in eqs. (6.21) and (6.22). Evaluating

these

expressions

near the critical cluster size gives the growth rate by fluctuations as
(11.19)
In the second form, the absorption rate at the

critical

pOint,

B*,

and

the

number of vacancies in the critical cluster, n*, were taken from Section 6.2.
Equating the growth rate due to thermal
growth

rate

given

in

eq.

(11.11),

fluctuations
yields

the

with

the

deterministic

cavity radius at which the

transition from statistical to deterministic growth takes place:


(11 .20)
critical

where R* = 2Yssinw/crb is the radius of the

cluster

and

is

the

abbreviation
(11.21)
which should be small compared

to

unity

if

the

above

calculation

of

the

statistical growth rate near the critical cluster size is to be justified.


Now

the

rupture

nucleation)

by

time

can

integrating

be
the

calculated
cavity

(ignoring

the

time

for

cavity

growth law, eq (11.14) including the


sintering stress, from R = Rt to A/2. Including the sintering stress renders

11.

156

Diffusive Cavity Growth

the integration more complicated. The problem is simplified if the factor


adjoining Go in eq. (11.14) is deleted, a change
framework

of

the

present

which

is

justified

(1-00)

in

the

theory which does not distinguish between surface

energy and surface tension. Then the rupture time takes the form
(11.22)
Here, all terms of order 2R*/A and of order Et were neglected compared to
unity. The effect of the sintering stress is described by the term containing
tnE t Obviously, the effect becomes large if Et becomes small, i.e., if thermal
fluctuations push the cluster size only slightly beyond the critical size. The
logarithms in eq. (11.22) depend on quantities which do not vary a
for

different

constants

in

3/4 - tn(2R*/A)

replaced by

materials
order

to

deal

and temperatures so that they can be approximated b)


simplify

the

= 5 and tnEt = - 7,

2!ssin~/Gb'

great

further
and

if

discussion.
the

If

we

take

critical cluster radius is

eq. (11.22) reduces to


(11.23)

In this approximation the rupture time contains the well-known 11Gb-term, but
additionally a l/G~ -term is obtained which originates from the sintering
stress. The fourth-power term dominates for stresses smaller than
(11.24 )
With Ys - 1.5 J/m 2 , ~ = 75 0 and A = 1 ~m, this means that for Gb < 65 MPa the
cavities spend an appreciable fraction of the lifetime near the sintering
limit, and the lifetime is extended markedly by the sintering forces.
11.1.4

Removal of cavities by compressive loads or by surface tension forces

Despite the limited success of


model

in

the

unconstrained

diffusional

cavity

growth

predicting failure times, we apply eq. (11.14) to the calculation of

sintering times which are necessary to shrink existing cavities to zero size.
The closure of cavities is of interest in powder metallurgy during hot isostatic pressing and other sintering processes. Reviews on sintering have been
prepared by Swinkels and Ashby (1979), by Swinkels et al (1983)

and

by

Exnen

11.1

Diffusive Growth of Equilibrium Cavities

and

Arzt

(1983).

Sintering

of

treatments of used parts which

157

cavities is also one of the goals of re-heat

may

have

developed

cavities

during

service

some

initial

operation. A few examples are quoted below.


To calculate the sintering time, eq. (11.14) is integrated
cavity

size, R

from

Ri , to R = O. The integral can be evaluated in closed form if


either ao = 0, or ab = O. The first case corresponds to sintering under high
compressive stress, and the second to sintering by surface tension forces
=

alone. The resulting sintering time, t s ' has the form

o
s

(11.25)

0,

where R~ was neglected against A2. For n-iron at


and
1
0
2
A = 311m, Ys
1.5 J/m , $ = 70 and the other material parameters taken from
Appendix A, the sintering time is found to be t s = 4 mins for a compr essi ve
stress

of

ab = ~- 50 MPa i f ao is neglected. The time to close the cavities by


surface tension forces alone, Le. ab = 0, is predicted from eq. (11.25) as
ts = 13 mins. For Y-iron at 950 0 C, the corresponding sintering times are 19 sec
under 50 MPa compression and 63 sec under the action of surface tension forces
alone.
Considering the shortness of these
rejuvenile

sintering

is

appears

tempting

to

parts that have developed cavities in service by giving them a heat

treatment at a temperature well above their


stresses

times

accelerate

sintering

service

temperature.

Compressive

but even the surface tension forces suffice to

remove the cavities within relatively short times as long as the

cavities

are

still small.
To the authors knowledge, eq. (11.25) has not been

compared

with

experiments

quantitatively but it is well established that re-heat treatments substantially


prolong the creep life of gold (Davies and Evans, 1965), brass and some of
nickel-base

alloys

(Davies

et aI, 1966, 1967). In most superalloys, however,

re-heat treatments have a complex


structure

and

simultaneous

effect

on

both,

the

the

micro-

on the cavities (Davies et aI, 1973, Davies and Dennison, 1975,

Dennison and Wilshire, 1977), and this effect may be beneficial or


for

creep

detrimental

life. Further, surface-connected cavities are usually oxidized

and are therefore hard to remove (Dennison et aI, 1978). A gas pressure in
voids

or

the

surface

the

coverage by segregated impurity atoms may have a similar

11.

158

effect. Moderate improvements of the creep lives of two

Diffusive Cavity Growth

austenitic

steels

by

interspersed heat treatments were reported by EI-Magd and Jager (1984).


11.1.5

The effect of impurity segregation on diffusive cavity growth

The effect of impurities on cavity nucleation was described in Chapter


their

effect

on

cavity

growth

is

8.

Now

considered. The grain-boundary diffusion

coefficient enters into the expression for the rupture

lifetime

in

the

form

t f 1/6D b Most segregants seem to reduce the self-dlffusivity along grain


boundaries which is beneficial to the lifetime. Zirconium addition to a Ni-20Cr
alloy,

for

example, reduces 6Db by a factor of ten at 800 0 C (Schneibel, White

and Padgett, 1983). The effect of several impurities and alloying

elements

on

6Db in iron is shown in Fig. 11.2 using data of Stratmann et al (1983). Grain
boundary diffusion in iron is also retarded by tin and antimony (Bernardini et
aI,

1982;

Gas

et

aI,

1986).

Data on other materials have been reviewed by

Gleiter and Chalmers (1972) and by Nix, Yu and Wang (1983).


The reduction of the grain boundary diffusivity by segregants can be

rational-

ized by an argument due to Borisov et al (1964). Since the atomic forces across
a grain boundary are reduced compared to the lattice, which is reflected in the
grain boundary energy,

the atoms should be more mobile.

Hence, the activation

- - T inC

700

800

t
$.

10-20

,,~

IO~ '~:,~
'

~v~

0-11

.s.,

600

:~:+400ppmN ~~

-13

10

"Fe + 120 ppm S


"- v,
+Fe + 90ppm C
"DFe +0.17% P
D "
- -Fe from Swinkels and Ashby'

10-2'

(1979)
0.9

1.0

1.1

1.2

10OOK/T---

Fig. 11.2. Grain boundary self-diffusion coefficient of iron containing


impurities (from tracer diffusion experiments by Stratmann et aI, 1983).

11.1

Diffusive Growth of Equilibrium Cavities

159

energy for diffusion along a grain boundary should

be

lowered

by

the

grain

boundary energy, Yb , compared to diffusion through the lattice. This leads to


(11.26)
where b is Burgers vector, Dv is volume diffusion coefficient and 6 is the
relative coverage of the grain boundary with impurity atoms. The factor (1-6)
was introduced here since impurity atoms occupy space in the grain boundary and
do

not

contribute

to

self-diffusion

energy which appears in eq.


accounts

for

(11.26)

is

of
now

the base metal. The grain boundary


taken

from

eq.

(8.10),

which

the effect of segregation. Normalizing the diffusion coefficient

by the value of the base material and eliminating the segregation

enthalpy

by

using eq. (8.1) gives the simple result


60 16D base metal
b

(11.27)

where Nb is the area density of available segregation sites, i.e., 2b2Nb ~ 1.


Of course, the Borisov argument is not a strict theory, but it predicts qualitatively correctly that 6Db decreases with increasing grain-boundary coverage.
In conclusion, the effect of impurity segregation is usually beneficial as
as

diffusive

cavity

growth

those which do not enhance cavity nucleation, or even impede


same

time

reduce

it,

and

at

the

the grain boundary diffusion coefficient strongly to retard

cavity growth. Of course, impurity elements additionally influence the


logy

far

is concerned. The optimum trace elements will be

morpho-

of carbides and of other precipitates, and this may often be the dominant

effect of impurities on the creep and rupture


beneficial

trace

el~ment

properties.

An

example

for

seems to be zirconium in nickel, which retards grain

boundary diffusion and forms stable sulfides within the grains thus

preventing

sulfur from segregating to grain boundaries.


11.1.6

The effect of gas pressure on the diffusive cavity growth rate

If the voids contain an internal gas pressure, p, the analysis


growth

problem

of

the

cavity

must be modi-fied. The gas pressure enters into the equilibrium

eondition, eq. (11.8), and into the chemical potential on the free surface, eq.
(11.4) (the gas does the work pO in removing an atom from the void surface). In
the final results for the cavity growth rate and for the rupture lifetime,
gas pressure is taken into account by replacing [Jb by [Jb

the

P (Raj et al, 1977).

11.

160

Diffusive Cavity Growtt

An illustrative application of these results to the growth of


during

hydrogen

attack

has

that at relatively low temperatures the chemical reaction


hydrogen

methane

bubbles

been given by Shih and Johnson (1982). They find


between

carbon

and

at the bubble surface is the rate-controlling step for methane forma-

tion using reaction rates measured by Grabke and Martin (1973). At higher temperatures

the

reaction is fast enough to build up the equilibrium pressure. In

response to the methane pressure, the bubbles


mechanism

grow

either

by

the

diffusive

or by creep of the matrix (see Chapter 14). Diffusive growth is pre-

ferred at high temperatures and low hydrogen pressures, while creep

predomina-

tes at lower temperatures and higher pressures.


In a review on hydrogen attack, Shewmon (1984) analyzed data by
Shewmon

(1981)

on

bubble

McKimpson

and

coalescence in carbon steel and found satisfactory

agreement with the diffusive growth model. The lifetime was found to depend

or

pressure as t f 1/p1.2, which is reasonably close to the predicted 1/p-dependence, and the observed temperature dependence agreed with that of grain boundary diffusion.
Another application of diffusive cavity growth under internal gas
that

of

pressure

is

Trinkaus and Ullmaier (1979) who analyze the nucleation and growth of

helium bubbles in stainless steel.

11.2 Diffusional Growth of Non-Equilibrium Cavities


Diffusional cavity growth under typical creep rupture conditions occurs predominantly by grain boundary diffusion rather than by volume diffusion. This means
that atoms leave the cavity through its perimeter in
surface

diffusion

along

atoms leads to flat, crack-like void shapes, rather


shape

assumed

in

the

the

grain

boundary.

If

the cavity surface is very slow, this local drain of


than

to

the

equilibrium

previous section. The conditions for the occurrence of

non-equilibrium shapes will be discussed

next.

In

fact,

flat

cavities

are

sometimes observed (Goods and Nix, 1978).


The theoretical discussion below starts from the set of equations for diffusion
in the grain boundary and along the free cavity surface given in Chapter 4. The
presentation follows a paper of Chuang et al (1979), which is based on
work

by

Chuang

earlier

and Rice (1973). A later numerical investigation by Pharr and

Nix (1979) substantiated the analytic results of Chuang et al.

11.2

11.2.1

Diffusive Growth of Nonequilibrium Cavities

161

The procedure to solve the coupled problem of surface diffusion and


grain boundary diffusion

The most convenient way to solve the coupled problem of surface


grain

diffusion

and

boundary diffusion during cavity growth is to start from an assumed cav-

ity growth rate, ACt). Although A is usually considered as the unknown quantity
to

be

determined,

it

is

easier to revert the procedure and to consider the

stress, 0b' acting on the boundary as unknown and the


Next,

the

surface

growth

rate

as

known.

diffusion problem is solved, which can be done for certain

simple growth rate histories. This gives the void shape, the chemical potential
and the flux into the void tip. The continuity conditions for the potential and
for the flux at the void tip link the surface diffusion problem

to

the

grain

boundary diffusion problem. The latter has already been solved in the preceding
section and the stress distribution only needs to be adapted
conditions

set

to

the

boundary

by the surface diffusion problem. The analysis is completed by

requiring global equilibrium, which relates the stress 0b to the initially prescribed growth rate ACt). The only difficult step in this procedure is to solve
the surface diffusion problem for a given growth rate. This will be done next.
11.2.2

He-formulation of the surface diffusion problem

In Chapter 4, the surface diffusion problem was formulated in terms of the curvatures and of the normal displacement rate,

vn ,

of the cavity surface. A form-

ulation which is more convenient here describes the cavity shape by the height,
w, of the cavity surface above the plane of the grain boundary as shown in Fig.
11.3. For the axisymmetric cavity, w is related to the old variables by:

r=O
equilibrium

v-O

transition

crack-like limit
11 _

CD

Fig. 11.3. Transition from equilibrium cavity shape to crack-like shape.


Definitions of geometrical quantities.

11.

162

Diffusive Cavity Growth

1<:1

- w" ( 1

+ w,2)-3/2

(11.28J

1<:2

- w' ( 1

+ w,2)-1/2 1 r

(11. 29J

(w - Un /2)

(1 + w,2)-1/2,

(11. 30)

where the primes and the dots denote the derivatives with respect to r

and

t,

respectively. The rigid-body displacement rate of the grains, ~, is subtractec

from
since n is meant to originate from surface diffusion alone while ~
comprises the whole displacement rate of the surface. This jacking effect of
the atoms plated on the grain boundary can be neglected for small
to-spacing

ratios

cavity-size-

and has been omitted by previous workers. When expressed ir

(4.14),

terms of w, the fundamental equation for surface diffusion, eq.

takes

the form
Q 60 s

v2

w"

Ys V2 (

kT

(1+w,2)3/2

(11.31)

0,

= a2 /ar2

+ (l/r)a/ar for the axisymmetric cavity. For the cylindrical


2
= a /ax 2 , delete the second term in parentheses and let the prime
denote a/ax. The fourth-order partial differential eq. (11.31) must be solved
subject to a prescribed growth rate and to the symmetry condition w, = w'" = 0

where

cavity set V2

in the certer of the cavity and to w = 0 and w, = latter

condition

tan~

at the void

tip angle has its equilibrium value,

~,

impeded.

equilibrium

void

shape

Hence,

the

as defined in eq. (6.5).

A preliminary measure for whether or not surface diffusion is

an

The

is introduced since it can be anticipated that local thermo-

dynamic equilibrium of the surface tension forces is not

maintain

tip.

fast

enough

to

is given by the following dimensionless

velocity formed by the parameters of the surface diffusion problem:


v = kT R3

R 1 (n

60 s Y )
s

(11.32)

If this dimensionless growth rate is small compared to unity, the


grow at an equilibrium shape, while for v

Next two specific solutions of eq. (11.31) are presented. The first
steady-state
(v

solution

which

applies

cavi ty

will

1, crack-like shapes will prevail.


one

is

to rapid growth in the crack-like limit

1). The second one is a similarity solution for v

const, which

is

not

11.2

to either small or large growth rates, but v = const corresponds to

restricted

163

Diffusive Growth of Nonequilibrium Cavities

t- 1/4 , which is not a very realistic time dependence of

The surface diffusion problem for v

the

growth rate.

0 is trivial. Here the void has its equi-

librium shape.
11.2.3

A steady-state solution of the surface diffusion problem in the crack-

like limit
Chuang and Rice (1973) surmised that for fast cavity growth, steady-state
utions

to

sol-

eq. (11.31) exist. If the cavity grows very fast due to rapid grain

boundary diffusion, while surface diffusion is relatively slow, the diffusional


activity

on

the

cavity

surface

is confined to a small region near the tip.

Further away, the curvature and therefore diffusional flow become


the

region
to

Thus

of high surface-diffusion activity can be treated as if the cavity

were infinitely large. Only if


expected

small.

this

is

so,

steady-state

solutions

can

be

exist. In this case , the time derivative of w can be replaced by

= - R w, + U
n /2, where the prime denotes the derivative with respect to x (as
defined in Fig. 11.3), and derivatives of w must go to zero as x + -~.

The steady-state problem is readily solved

if

the

governing

eq. (11.31)

is

linearized by neglecting w,2 against unity. The linearized treatment is limited

W, but the solution will be generalized for arbitrary


W shortly. With the steady-state condition, w = - Rw, + Un 12, inserted the

to small void tip angles

linearized equation for w takes the form


kT R dw
d 4w
dx 4 - n 60 s Ys dx

(11.33)

This equation applies equally to cylindrical and aXisymmetric


the

steady-state

approximation

cavities,

since

requires that the cavity is so large that the

term w'/r in eq. (11.31) is everywhere small and can be neglected.


The ordinary differential

eq. (11.33)

has

-w

boundary

exponential

solutions.

With

conditions w = un/2 and dw/dx =


at x = 0 and dw/dx = 0 for x
one obtains the steady-state solution valid for small
dw
dx

- W exp [(

kT

n 60 s Ys

)1/3 x]

w:

the
+

-~

(11.34)

Diffusive Cavity Growth

11.

164

w - un /2

iC 1

iC2

kT it
) 1 13 x]}.
g 60s 'Vs

(11.35)

kT R ) 1 13 exp[ (
kT it
) 1 13 x].
g 60 s 'Vs
goDs 'V s

( 11. 36)

g 60 s 'Vs

1/1

kT
d2w

-~=

dx

1/1 (

1/3

{1 - exp [(

The curvature determines the chemical potential, and its derivative

gives

the

flux according to eqs. (4.11) and (4.12). The continuity conditions at the void
tip for the potential and for the flux set

the

boundary

conditions

for

the

grain boundary diffusion problem to be considered in Section 11.2.5.


Chuang and Rice (1973) have solved the steady-state problem

without

resorting

to the linearization employed here. Comparison with their result shows that the
linear solution for the flux and the curvature at the void tip is
within

2%

for

accurate

to

arbitrary 1/1, if 1/1 is replaced by 2 sin(1/I/2) in eqs. (11.34) to

(11.36).
11.2.4

Similarity solutions for the surface diffusion problem

Dimensional considerations [or insertion of the expression for w below into the
governing equation (11.31)] show that
(11.37)
represents a family of self-similar solutions of eq. (11.31). The dimensionless
coordinate

is defined as
(11. 38)

for the axisymmetric cavity. From the form of the Similarity coordinate
clear

t 1/4 ,

that the cavity expands in proportion to

as R ~ t
ordinary

-3/4

or the growth rate varies

The

solutions

will

solved

linearized

it

reproduces

the

the

of

that

growth

rate

and

equilibrium and crack-like (steady-state) shapes for

small and large v, respectively. At intermediate values of


solution

version

not be shown here. It suffices to note that the

Similarity solution is valid for arbitrary magnitude of


that

is

,or v = const. The dimensionless shape function W(~) obeys the


differential equation, which follows by insertion of eq. (11.37) into

eq. (11.31). Chuang et al (1979) have


problem.

it

v,

the

similarity

acts as a smooth transition between .the limiting cases. The price for

11.2

165

Diffusive Growth of Nonequilibrium Cavities

this seemingly comprehensive solution is a rather artificial stressing

history

which follows for the similarity solutions (see Chuang et aI, 1979).
The stress associated with steady-state, crack-like cavity growth

is

expected

to be time-independent. Its magnitude will be derived next by coupling the surface diffusion problem with the grain boundary diffusion problem.
11.2.5

The relation between growth rate and stress in the crack-like limit

The partial problems of surface diffusion


linked

together

by

the

requirements

and
that,

grain
at

the

boundary

diffusion

are

void tip, the chemical

potential and the flux must be continuous. Continuity of the chemical potential
is

expressed

byeq. (11.4). Equating twice the flux on a surface, eq. (4.12),

to that along the grain boundary, eq. (4.3), specifies don/dr at the void
which is denoted by

tip,

0' .

(11.39)
The curvatures, K1 +K2 , at the void tip are known from eq. (11.36).
The distribution of the normal stress on on the grain boundary was
connection

with

equilibrium

cavity

valid; un is &pecified by the requirement that dOn/dr = o~


which gives

u
n

derived

in

growth. The result, eq. (11.3), is still


at

the

void

0'

tip,

(11.40)

The solution of the void-growth problem is completed by equating the average of


on to the stress, 0b' acting on the grain boundary. For the axisymmetric cavity
in the crack-like limit,
following

the

equilibrium

condition,

eq. (11.8),

form if eqs. (11.4), (11.39) and (11.40) are inserted for

takes

0 ,

the

0' and
o

un' respectively, and if Kl + K2 is taken from eq. (11.36):


(11.41)
with the abbreviations
(11.42)

11.

166

Diffusive Cavity Growth

(11.43)
The normalized growth rate v, q(w) and w were defined in eqs. (11.32), (11.9)
and (11.10), respectively. Solving eq. (11.41) for v and using the definition
of v leads to the final result for the growth rate of an axisymmetric cavity in
the crack-like limit:
(11.44)

with the notation


(11. 45)
As in the case of equilibrium cavity growth, an internal gas
cavity can be taken into account by writing ab

pressure

in

the

p instead of abo

Depending on the value of the product Qr~, the growth rate given in eq. (11.44)
has two limiting ranges. For very slow surface diffusion, i.e. if Qr~ 1, a
power-series expansion of the square root in eq. (11.44)

up

to

linear

terms

leads to
(11.46)

This is a cube-law relation between Rand abo It is interesting to note that


the growth rate in this limit is controlled by surface diffusion alone while
the grain boundary diffusion coefficient does not appear in the result.
For very high stress, i.e. if

Qr~

1, unity can be neglected in the square

root in eq. (11.44), so that


f{

= _ _12_---,1"""7"<"

I(T

(y

c5D )172
s
s

3 c5Db ab
)3/2
2 A Q (1-w)2 sin(~/2)

In this range the stress exponent is 3/2. Surface

diffusion

(11.47)

has

retarding

effect since c5Ds appears in the denominator. The reason for this is that for an
increased rate of surface diffusion the void shape tends to a rounded, equilibrium shape so that more atoms must be removed through the grain boundary.

11.2

Diffusive Growth of Nonequilibrium Cavities

161

Which one of the limiting cases applies, depends on the normalized growth

rate

v defined in eq. (11.32) and on the product QtA. At low stresses, or alternately, low growth rates or high surface
vails.

At

diffusivities,

equilibrium

like mode. This regime is subdivided into a low-stress,


range,

in

growth

pre-

high stresses or high growth rates, the cavities grow in the crackslow-surface-diffusion

which the growth rate is proportional to the third power of stress,

and into a high-stress range, in which a (3/2) power law obtains. Chuang et

al

(1919) investigate the ranges of validity of the limiting cases in detail.


Finally, the growth rate of a cylindrical, as distinct

from

an

axisymmetriC,

cavity in the crack-like limit is quoted from Chuang et al (1919) to be:


21 D 60s 'Vs

( 1+ tAl 1 12 - 1

8 kT >.3 A3

- 2R/>.

(11.48)

where t has been defined in eq. (11.45) and A in eq. (11.42).


11.2.6

Rupture times for non-equilibrium grOwth

If cavity nucleation is assumed to occur readily at the beginning of a test, or


if pre-existing cavities are present, rupture times are obtained by integration
of the cavity growth law, eq. (11.44), from an initial cavity size,
coalescence, R = >'12. Thus, in analogy to eq. (11.18), one obtains:

Ro'

2 kT, >.4 A3

to

{11.49)

21 D 60 s 'Vs

Here Wo
(11.42).

(2R o l>.)2 and Q, t and A are


Since

defined

in

eqs. (11.43),

over

and

Q depends on w in a complicated manner, the integral cannot be

carried out in closed form. However, the value of Q does not vary
extent

(11.45)

the

entire

to

well if Q is set equal to 0.6 (Chuang et aI, 1919). If, furthermore,


neglected compared to 1, the result for the rupture time becomes

0.015

great

range of w, so that the integral can be approximated

D 60s 'Vs

(1 - 35 R 18>.) {
0

(1

A 172
}3
0.6tA)
- 1

Again, the result has two limiting ranges. First, if 0.6tA

Wo

is

(11.50)

1, i.e. for small

stresses and low surface diffusivities, a series expansion of eq. (11.50) leads
to the inverse cube law for the rupture time

11.

168

Diffusive Cavity Growth

(11.51)

This result is approximately valid if


growth

prevails,

i.e. if v

Chuang et aI, 1979). Thus the

0.6L~ ~

1.5

and

if

crack-like

cavity

1. (How large exactly v must be was examined by


stress

range

within

which

the

inverse

cube

relation applies is determined by


(11 .52)

These inequalities can only be met simultaneously if oDb/oDs is rather large.


The range of validity of the inverse cube law as expressed by eq. (11.52) is
shown graphically in Fig. 11.4.
The opposite limiting range of the
Neglecting 1 against

rupture

time

is

defined

by

0.6L~

oDs Ys

(1 - 4.4 RIA) (
0

Y oDs sin(w/2) 3/2


s
).
A oDb 0b

100, the 3/2 power law prevails for stresses satisfying

200 r--,--,----.-----..----,
t

150

OC

l/a:rz
2.5x

'"

~ 100

""-

.-<
b

50
tf

OC 7/a

OL-.-..L..----'--------'-------'------J

1.

in eq. (11.50) leads to

kT A4

0.058
If oDb/oDs

0.6L~

20

1,0

60

80

100

Fig. 11.4. Range of validity of the inverse cube law, eq. (11.51).

(11.53)

11.2

169

Diffusive Growth of Nonequilibrium Cavities

(11.54)
For small 6Db/6Ds the inverse cube law does not obtain and the
occupies the whole crack-like limit, i.e. it is valid for

power

3/2

law

(11.55)

11.2.7

Experiments on copper and silver containing water vapor bubbles

Nieh and Nix (1980a) measured rupture lifetimes of copper


water

vapor

bubbles

had

specimens

in

which

been implanted on grain boundaries before the creep

tests in order to bypass the cavity nucleation stage. The stress and

temperat-

ure dependence of the measured lifetimes are in good agreement with the inverse
cube law t f ~ 1/03 , eq. (11.51), which exhibits the activation energy of surface diffusion. By doing tension and torsion tests on the same kind of material
(copper with water vapor bubbles), Stanzl, Argon and Tschegg
the

(1983)

confirmed

observations of Nieh and Nix, and they provided the additional information

that the rupture lifetime is determined solely by the maximum principal stress,
but not by the von Mises equivalent stress.
There is, however, a serious problem associated with
(11.51)

the

application

of

eq.

to the experiments of Nieh and Nix and of Stanzl et al. From Fig. 11.4

it. is clear that an inverse

cube

law

is

only

possible

if

o)./Ys

> 60 and

6Db/6DS > 30. Both requirements are violated in the experiments of Nieh and
Nix, at least at first sight. The bubble spacing is ). = 3.2 ~m, the surface
2

Ys = 1.7 Jim
for copper and the applied stresses range from 4 to
60 MPa. Hence o)'/Ys ranges from 7 to 110 with only two data points, out of 17,
energy

is

having

values

greater

than 60, which is required for the inverse cube law to

hold. If the surface energy were reduced by a factor 2 due to the


water
o)'/Y s

used

vapor,

which

is

an

> 60. Further the ratio

extreme

presence

of

assumption, 10 data pOints would satisfy

6D b /6D s has values near

unity if the commonly


grain-boundary diffusion coefficient of copper is employed. This value is

admittedly uncertain. Only if the actual value of 6Db were much greater, or

if

6D s were much smaller than the commonly accepted value, could the occurrence of
the inverse cube law for the rupture time be expected.
Goods and Nix (1978) have performed the same kind of experiments on silver containing

water

vapor

bubbles.

As in the case of copper, the measured rupture

Diffusive Cavity Growth

11.

170

lifetimes approximately exhibit the inverse cube-law dependence on str.ess


dicted

by

pre-

the crack-like diffusive growth model in the slow-surface-diffusion

limit, eq. (11.51). However, the same problem arises as in the case of embrittled

copper.

As Chuang et al (1979) point out the available data indicate that

the surface diffusion coefficient is larger than the


coefficient

under

grain-boundary

diffusion

the conditions of Goods' and Nix' tests. But eq. (11.51) is

only applicable if the reverse is true. Goods and Nix (1978) suggest to resolve
the

contradiction

by

postulating that the diffusion coefficients in the pre-

sence of water vapor are very different from those usually measured.
Although such an assumption is not satisfactory, the idea that
by

diffusive

cavity

failure

occurs

growth is supported by experiments performed by Nieh and

Nix (1979). They also tested silver with implanted

water

vapor

bubbles,

but

they strengthened the silver matrix by a fine dispersion of magnesia particles.


Although the creep rate of the MgO-dispersion
orders

of

magnitude

strengthened

material

was

two

lower than that of pure silver, the rupture lifetime was

not affected. This precludes a cavity growth mechanism which is

controlled

by

creep

strain. Among the diffusion-controlled mechanisms only crack-like growth


leads to a strong stress dependence of the type t f 11a3
In conclusion, although the inverse cube law, eq. (11.51),
copper

and

fits

the

data

on

silver with implanted water vapor bubbles well, there remains some

doubt as to whether the inverse cube law should be applicable to these cases.
11.2.8

Void~shaBe

instability/finger-like cavity growth

Taplin and Wingrove (1967) and Wingrove and Taplin (1969b) have

observed

that

cavities in iron, when growing at a high rate, assume an irregular shape. Their
periphery in the grain boundary, which is usually circular, breaks
wavy

shape

and

up

into

eventually develops finger-like protrusions as shown schemat-

ically in Fig. 11.5.

Fields and Ashby (1976) have interpreted this observation

as an intrinsic instability of crack-like diffusive cavity growth. The phenomenon resembles the shape instability of a void in a thin
two

glass

plates

liquid

layer

between

which are pulled apart. The shape instability of crack-like

cavities arises from the interplay of surface tension forces and grain boundary
diffusion, which draws matter out of the crack. A protrusion of the crack front
implies that the surface is more strongly curved there, which is unfavorable in
terms

of

the

chemical potential. Hence surface diffusion tends to smooth out

the protrusion. On the other

hand,

the

protrusion

is

favored

for

kinetic

11.2

171

Diffusive Growth of Nonequilibrium Cavities

reasons:

the atoms leaving the cavity through a protrusion can spread into the

grain boundary along divergent flow lines, whereas the atoms leaving the cavity
through

an

intrusion

in the crack front must share the limited space between

the adjacent protrusions which retards the diffusive flux. Therefore the
of

atoms

is

stronger

drain

through the protrusion, so that the perturbance of the

cavity shape is enhanced. For rapid cavity growth and for slow surface

diffus-

ion, the kinetic aspect dominates leading to the shape instability, whereas for
slow growth and rapid surface diffusion, the cavities retain a regular shape.
The evolution of such shape instabilities from initially straight crack
was

fronts

investigated quantitatively by Fields and Ashby (1976). They superimpose a

sinusoidal perturbance on the cavity front and examine whether the

disturbance

increases or decreases during growth. There results that a disturbance grows if


its wavelength is greater than the critical value:
( 11.56)

"crit
Here,
as

w~

is half the height of the cavity at x = -

function

of

the

stress,

w~

and

R can

~.

If the result is

be taken from eqs. (11.35) and

(11.48), respectively. Of course, if the cavity is smaller than


wavelength,

the

desired

this

critical

instability cannot develop, but if the cavity size becomes of

the order of the critical wavelength, a shape instability is expected to occur.


Beere

(1978)

has considered the more general case of initially circular crack

fronts, which is more appropriate

to

describe

the

instability

of

cavities

having a finite size, but, on the other hand, the result is more complicated.

aj

bj

diffusion in grain
boundary

Fig. 11.5. Plan view on a grain boundary containing crack-like cavities.


a) Circular shape, b) inCipient instability, c) finger-like protrusions.

12 Constrained DitJusive Cavitation


of Grain Boundaries

Rupture lifetimes of engineering alloys are generally


would

expect

from

the

rate

of

diffusive

much

greater

than

one

cavity growth. Motivated by this

general observation, Dyson (1916, 1919) proposed his constrained cavity

growth

model, which is explained next.


Over a substantial fraction of the creep lifetime, cavitation
relatively

isolated

material. This is
cavitation

grain

so

-depends

boundary

because
on

its

the

confined

to

facets which are surrounded by undamaged


susceptibility

orientation

to

of

the

grain

stress

boundary

axis

orientations of the adjacent grains [Lim and Raj (1984a) for


Majumdar

is

and

nickel;

to

on

the

Don

and

(1986) for stainless steel]. If the material surrounding a cavitating

facet were rigid, the excess volume of the cavities could not be accommodated
and cavity growth would come to a standstill. Thus the rate of cavity growth
can be controlled by the deformation rate of the surrounding material. This
called

constrained

cavity

growth.

On

the

other

hand,

if the surrounding

material is relatively soft, the accommodation process occurs readily


rate

of cavity

~owth

cavitation

not enclosed
the

~n

constraint

the

In

bi-crystal

specimens,

is obviously not constrained since the grain boundary is

a cage of non-cavitating material. A useful


is

and

is controlled by diffusive transport as described in the

preceding chapter. This is the unconstrained limit.


diffusive

is

that

cavitating grain boundary

interpretation

of

the surrounding material exerts a back stress on the


facet.

The

resulting

stress,

Db'

on

the

grain

boundary adjusts itself such that the rate of the volume increase by diffusive
cavitation is compatible with the deformation rate of the surrounding material.
In

the unconstrained limit, the cavitating facet experiences the whole applied

stress, i.e. Db = D~, whereas in the constrained limit,


boundary is reduced to the order of the sintering stress.

the

stress

on

the

12.1

173

Constrained Cavitation of an Isolated Facet

12.1 Cavity Growth Rates for Constrained Cavitation


of an Isolated Facet
12.1.1

A tensile-crack model for the calculation of constrained growth rates

The idea of constrained cavity growth is nicely illustrated and quantified by a


model

proposed originally by Rice (1981), which was extended by Riedel (1983b)

and by Tvergaard (1984a). The model is shown in Fig. 12.1. A cavitating

bound-

ary facet is described as a penny-shaped crack of diameter d embedded in an infinite block of power-law creeping material characterized

by

material is stressed at large distances from the crack by

a;j.

below is carried out for aXisymmetric loading,


ponents

of

the

a;,

applied

i.e.

the

stress are the normal stress,

eq.

only

ai,

(3.6).

(The calculation
non-zero

com-

and the transverse

stress,
as shown in the figure). The crack transmits the normal stress
which is determined from the following compatibility argument.
The volume growth rate of all cavities in the boundary facet can be
in

two

independent

ways.

Under

the

action

the

number

ab ,

calculated

of a stress ab , the volumetric

diffusive growth rate of a single cavity is given by


with

The

eq. (11.11).

Multiplying

of cavities, (n/4)(d/A)2, gives the total growth rate of the

whole facet:
(12.1)

--

- - - (1r""

100----

----.j

Fig. 12.1. Cavitating grain boundary facet. The penny-shaped crack model
idealizes the cavitating facet as circular and ignores other boundaries.

12.

174

Constrained Diffusive Cavitation

where w is the cavitated area fraction of the boundary, q(w) was defined in eq.
(11.10) and ao is the sintering stress of an isolated cavity.
On the other hand, the opening rate of a
viscous

body

penny-shaped

crack

in

nonlinear

can be inferred from work of He and Hutchinson (1981), which was

already referred to in connection with the shear crack model of grain


sliding

in

Section

7.1.2.

Using

boundary

linearization method, He and Hutchinson

develop approximate solutions which turn out to be highly accurate under a wide
range

of

conditions

when

compared to numerical solutions. He and Hutchinson

consider the case ab = a only,


straightforward. Since He's and

is
but a generalization to finite ab
Hutchinson's method essentially consists of

reducing the nonlinear viscous problem to a linear


make

use

of

a;

replacing
linear

the
by

fact

a;-ab

approximation

that

normal

elastic

in linear elastic crack problems.


is

problem,

one

can

stress on a crack can be included by


The

accuracy

of

the

expected to increase for increasing abo In this way,

one obtains the volume growth rate of the facet by creep flow of the matrix
(12.2)
<II

where the applied equivalent stress for axisymmetric loading is ae


.CD
CD n
and the equivalent strain rate is Ee = B(ae )

For compatibility, the volume growth rates by diffusive cavitation and by creep
flow must be equal. Equating eqs. (12.1) and (12.2) defines ab as
(12.3)

where q'

w2 (1+3/n)1/2 is an abbreviation. Substitution of

eq.

(12.3)

into

eq. (11.14) gives the constrained cavity growth rate


(12.4 )

Equations (12.3) and (12.4) clearly show


surrounding

material

the

constraining

effect

which

the

exerts on cavitation on an isolated facet. If the strain

rate is large, the cavity volume produced by diffusive transport can easily
accommodated. Correspondingly, the
(12.3) and (12.4) vanish for large

second

be

terms in the denominators of eqs.

which. implies

that

the

full

applied

12.1

Constrained Cavitation of an Isolated Facet

stress

acts

on

175

the boundary, and the growth rate, eq. (12.4), reduces to the

unconstrained limit, eq. (11.14). On the other hand, if the strain rate becomes
very small, 0b becomes small and approaches the sintering stress, while the
cavity growth rate becomes accommodation controlled and approaches
the
constrained limit:
(12.5)

where q' was approximated

by

12.5

and

h(~)

by

0.61.

The

grain

boundary

diffusion coefficient plays no role in the constrained limit. The cavity growth
rate now varies in proportion to the strain rate.
A condition for whether or not the constraint is effective for a
rate

is

obtained

by

given

strain

equating the two terms in the denominator of eq. (12.4)

which gives
(12.6)

For strain rates higher than this characteristic strain rate, cavity growth
unconstrained,

is

whereas for smaller strain rates, cavity growth is constrained.

Of course, eq. (12.6) can also be expressed in terms of a characteristic stress


by

virtue

of

B(o~)n. The transition is illustrated in the figures shown

E~

below, in particular in Fig. 12.3.


12.1.2

Comparison with measured cavity growth rates

An extensive investigation on cavity nucleation and growth was carried


Needham

and

co-workers.

The

results

out

by

on ferritic steels are summarized in a

report by Needham (1983) and by Needham and Gladman (1984).

In

this

section,

cavity growth rates from these studies are compared with theoretical predictions, while the pertinent rupture lifetimes will be given in Chapter 16. Figure
12.2

shows the growth rate of the i-th largest cavity (with i * 1000) observed
by Needham (1983) in two ferritic steels at 550 oC. The dashed lines represent
the

growth rate

expected

for

the

unconstrained diffusive growth mechanism

according to eq. (11.14). The sintering stress is neglected and the


q(w)h(~)/2

is

values for nand

approximated

oD b

by

expression

1 as a representative average value. With the

from Appendix A, eq. (11.14) then leads to

176

12.

10-10

55O"C

.s

------------

----R":02IJ.m
10-11

R=05/-L1JI_-------

------

III

"'E

Constrained Diffusive Cavitation

10-12

R= 1JJ.T----

------

Q::

100
(J

150 200 250 300


in MPa __

Fig. 12.2._Cavity growth rates in two ferritic steels measured by Needham

(1983). Solid lines: Constrained growth, eq. (12.4). Dashed lines:


Unconstrained growth, eq. (11.14); (after Riedel, 1985c).

(12.7)
where strgss is in MPa, length in meters and time in
shows

the

unconstrained

diffusive

mechanism

seconds.

12.2
for

figure

solid

lines

in

represent the constrained growth rates according to eq. (12.4) with

the grain size d


formula

the

overestimates the growth rates

considerably and,leads to an incorrect stress dependence. The


Fig.

As

H,

18

~m

strain

Needham (1983), and the

as given by
rates

are

cavity

Needham

and

with

h(~)

= 0.61.

In

the

given the experimental values reported by

spacing

is

estimated

cavity
densities. The cavity density varies from (0.5 to 2.5). 1011 /m 2 in the 1Cr- 1/ 2 Mo

steel investigated and is by a factor 2.5 smaller in the


11 2
representative
values we take 1/), 2
101m,
l.e.

from

reported

steel.

21/~Cr-1Mo

).

3.2

for

~m,

As
the

1Cr-1/2Mo steel ~nd 1/).2 = 0.5'10 11 /m 2 , i.e. ). = 4.5 ~m, for the 21/~Cr-1Mo
steel. With this choice of the parameters, the agreement with the experimental
data is good, but it should be kept in mind that Rand
tests

substantially'

whereas

the

calculation

average values. Nevertheless, the agreement of


with

Needham's

data

strongly

suggests

constrained under these test conditions.

that

was
the

).

varied

during

the

done using representative


constrained

diffusive

growth

growth

is

model
indeed

12.1

177

Constrained Cavitation of an Isolated Facet

A similar conclusion is suggested by experiments on 21/~Cr-1Mo steel


by

Cane (1979). The material was austenized at

grain size of d
act

as

1300 0 C.

This gives an austenite

= 150 pm and produces fine sulfides on grain boundaries, which

cavity nucleation sites. The cavity spacing was typically A = 4.5 pm,

and the cavities observed had radii of typically R = 1.8 pm.


the

565 0C

at

cavity

growth

rate,

are

~,

plotted

The

results

for

Ale

as a

in Fig. 12.3 in the form

function of stress. For comparison, the prediction of

the

constrained

growth

model, eq. (12.4),

ru:

(12.8)

A were
inserted numerically. (Units are m, sand MPa). Strain rate and stress were
taken from plots given by Cane (1979). At low stresses, cavity growth is
constrained
and
proceeds much slower than would be predicted by the
unconstrained diffusive growth model, which is represented by the dashed line.
Between 100 and 200 MPa a transition to unconstrained growth occurs, which
means a decrease in HI: since now diffusion becomes rate controlling. The
constrained growth model agrees with the measurements of Cane to within a
factor 2.3, which is satisfactory, while the unconstrained model exhibits a
totally wrong behavior at low stresses.
is also plotted, where the material parameters, as well as

In a later paper, Cane (1981) investigates cavity growth in


of the same steel.

For his

notch geometry is

150

a;la;

d,

R and

notched

specimens

= 2 so that, according to

\ unconstrai

E 100

(.onstrained \

:::t

.5;

.~

'0:::

50f----

Cane(exp

\
~

o~~~--~--~~

60

100

150

200

UinMPa ---

Fig. 12.3. Cavity growth rate in over-heated 2 1/.Cr-1Mo steel at 565 0 C


measured by Cane (1979). Constrained and unconstrained growth rates
from eqs. (12.4) and (11.14), respectively.

12.

178

Constrained Diffusive Cavitatior

the constrained growth model, eq. (12.4), the growth rate

is

expected

to

a;.

twice that of the uniaxial tension tests for the same value of
presents only four measurements on notched specimens. On the average,
data

lie

by

factor

1.5

above

the

data

bE

CanE
thesE

for uniaxial tension, which iE

reasonably close to the expected factor 2.


Finally, the constrained growth model is compared with data on Type 347
less

steel

measured

sults. The theoretical curves were computed using the grain boundary
coefficients

diffusion

given by Needham and Gladman, which agree with those reported for

other stainless steels and for Y-iron by Frost and Ashby (1982) and
given

stain-

by Needham and Gladman (1980). Figure 12.4 shows the re-

in Appendix A. The cavity radius was chosen as R

spacing as A = 6.3

~m,

0.6

been obtained in this range. The grain size is reported as d

are

and the cavity

~m

since the bulk of the experimental data

which

seems
12

~m.

to

have

The creep

rate can be described by = 2.10-35 all at 550 0 C and by E = 4.3.10-26 0 8 1 at


650 oC, where time is in seconds and stress in MPa. Again the expression
q(w)h(~)

was approximated by 2.

It is apparent from Fig. 12.4 that the constrained growth

model

overestimates

the growth rates in this case. However, the theoretical result is sensitive

10-

to

347 stainless steel

-I--

10- 17

I.

--f-~

650C

.s:

00:: 10-12

10-13

100

7
150

/.;
.I

200
(J

/
b 55OC

300

400

in MPa - -

Fig. 12.4. Cavity growth rates in rype 347 stainless steel, from Needham and
Gladman (1980). Solid lines: eq. (12.4). Dashed lines: Completely constrained
and unconstrained limits. Dashed-dotted line: Plastic hole growth mechanism.

12.1

Constrained Cavitation of an Isolated Facet

179

the choice of R and A, and it is not exactly clear for which values of R and
Needham

and

Gladman

(1980)

give

their

experimental

A
growth rates. If, for

example, the theoretical curves had been plotted for A = 3 urn


they

and

0.8 urn,

would have approximated the experimental data closely. The calculation of

lifetimes does not involve these uncertainties and the agreement of theory
experiment

is

and

good, if continuous cavity nucleation is taken into account, as

will be shown in Chapter 16.


Figure 12.4 also shows the growth rate by the
mechanism

for

continuum

plastic

hole

growth

550 0 C (dashed-dotted line). The continuum hole growth mechanism

will be described in Chapter 14. This mechanism usually underestimates the


served

growth

rate

ob-

significantly and there are no uncertain parameters which

could account for the difference.


Mancuso and Li (1979) investigated the growth
under

of

methane

bubbles

nickel

the internal gas pressure and under an applied tensile stress. They show

that at high streSSes


diffusive

growth

the

model,

observed
eq.

growth

(11.14).

rates

are

compatible

qualitative

with

the

At lower stresses, the growth rate is

substantially less than predicted by the diffusive model. This


in

in

observation

is

agreement with the constrained growth model, eq. (12.4), but a

quantitative comparison cannot be made since the relevant creep

data

for

the

therefore

the

material used by Mancuso and Li are not available.


12.1.3

Additional remarks on constrained cavity growth rates

At very low stresses the


accommodation

of

the

deformation
cavity

power-law, dislocation creep.


viscous

flow

of

polycrystal,

and

volume,occur by diffusion creep rather than by


Diffusion

creep

can

be

described

as

linear

as in eqs. (1.4) and (1.5), at least on a scale greater than the

grain size. If this macroscopic description of diffusion creep is adopted, eqs.


(12.1) to (12.4) remain valid with n = 1 and with ~:/a: = ~ QoD b/(kTd 3 ), in
the range where Coble creep dominates.
A certain amount of cavity volume can also be accommodated elastically.
growth

progresses

Cavity

until

the stress on the boundary, ab , equals the sintering


stress, which we neglect here. For ab = 0, the elastic opening volume of a
penny-shaped crack is
(12.9)

12.

180

Constrained Diffusive Cavitation

If (v/4)(d/A)2 cavities share this volume, their size-to-spacing ratio is


(12.10)

Greater

cavities

accommodation
v

0.3, h(tjI)

cannot

mechanism
=

be

accommodated

intervenes,

0.61, d/A

'"
10, aIlE

cavity

elastically,
growth

cavitation.

Nonlinear

strain-rate

viscous

which

relation

&~

is

sliding

is

grains
nonlinear

B(fa)n

sliding on

greater

other
0.17.

constrained

separated by freely sliding boundaries


viscous

also,

but

with

describes the grains.

If

stressl

grain

boundary

taken into account in this macroscopic way, the constrained cavity

growth rate retains its form given in eq. (12.4). Only the remote
is

no

where f is the stress enhancement factor de-

&= B~

scribed in Section 1.5.2 and

if

at this size. For

210 -4 ,eq. (12.10) gives 2RIA

A third remark refers to the effect of grain boundary


constitute a solid

and

stops

by

factor

fn

than

it

is

strain

rate

for non-sliding grain boundaries.

According to Rice (1981), who uses a different argument, grain boundary sliding
should enhance the volume growth rate of a penny-shaped crack by a factor of
about two. This is consistent with the estimate based on the stress enhancement
factor. It should be noted that for closely spaced as opposed to isolated cavitating facets, the effect of grain
(Section

12.6.3).

Also,

if

boundary

sliding

is

very

much

only a compressive transverse stress,

a;

stronger

< 0, is

applied, sliding plays a special role to be discussed in Section 14.2.


Finally, we remark that also non-equilibrium, crack-like cavity

growth,

which

was described in Section 11.2, may be constrained under certain conditions. The
volume growth rate of the facet for crack-like growth is

still

given

by eq.

(12.1), so that eq. (12.3) for ab remains valid, but now the stress at the tip
of the cavities, ao ' becomes dependent on the cavity growth rate. The further
analysis is complicated unless it is confined to the case where ao can be
negelected. This corresponds to the limiting range in which the
rate

increases

with

the

cavity

growth

power of stress. If ab is substituted from eq.


(12.3) with ao 0 into eq. (11.47), there results a cavity growth rate which
varies as A &3/2 a3n/2 in the constrained limit, and which exhibits the
3/2

transition to unconstrained crack-like growth at the characteristic strain rate


given in eq. (12.6).

12.2

Time to Cavity Coalescence

181

12.2 The TIme to Cavity Coalescence on an Isolated Boundary Facet


If cavity nucleation occurs early
coalescence,

tc'

can

be

in

the

creep

life,

the

time

to

cavity

obtained by integrating the growth law, eq. (12.4),

from time t = 0, when R = Ro' to t = t c ' when R = A/2 (cavity coalescence). If


is neglected, time integration of eq. (12.4) is elementary by separation of
0

the variables Rand t. If the radius at nucleation, Ro'

is

neglected

against

the cavity spaCing, A, there results


(12.11)

The first term represents the time to coalescence in the absence


straint,

of

the

con-

i.e. for diffusion being rate controlling. The second term represents

the constraint, and leads to greatly prolonged times to

coalescence

at

small

strain rates. At first sight, this seems to provide the desired explanation for
the prolonged rupture lifetimes of commercial alloys compared to
calculated

from

the

lifetime

unconstrained diffusive growth. However, an interpretation of

tc as the rupture time is incorrect, as the following section shows.

12.3 On the Irrelevance of Constrained Cavity Growth


for Rupture Lifetimes
In the previous literature, the time
identified

with

the

time

to

cavity

coalescence

has

often

been

to rupture. This assumption implies that, once the

cavities have coalesced thus forming grain-size microcracks, rupture occurs

by

some fast mechanism which joins the microcracks.


In the unconstrained limit, this assumption is not unreasonable. The cavitating
facets retain their strength, i.e. they transmit the full applied stress, up to
the point of cavity coalescence. Therefore the final joining of the

facet-size

microcracks will only commence after coalescence has occurred. In this case,
the time to cavity cOqlescence can be expected to control the lifetime, if the
additional time which is required for the joining of the facet-size microcracks
is small.
In the constrained limit, however, the situation is different as Riedel (1985c)
pointed

out.

Constrained cavity growth necessarily implies that the stress on

the cavitating boundary facet is relaxed to the level of the

sintering

stress

12.

182

which

is

often

negligible.

In

other

cavitation behaves mechanically like a


although

metallographically

Constrained Diffusive Cavitation

words, a facet undergoing constrained


practically

traction-free

microcrack,

it may still look like a boundary which is barely

cavitated. It appears logical that the mechanism which leads to final


whatever

its

nature might be, will respond to the loss of mechanical strength

of the facet rather than to cavity coalescence on the


crystalline

failure,

constraint

may

facet.

Thus

the

poly-

well retard cavity growth, but once this happens,

further cavity growth and coalescence is irrelevant for

the

rupture

process.

The constrained growth model does not lend itself to the calculation of rupture
lifetimes. Rather, one is led to the conclusion that, in the constrained limit,
it

must be the joining process of the cavitating facets which is rate-control-

ling for the rupture lifetime, possibly together with the continuous nucleation
of

new

cavities.

Interaction

effects between the facets will playa central

role for the joining process. Lifetime estimates based on

this

idea

will

be

developed in Section 12.7.


Although the time to cavity coalescence is not obviously related to the rupture
lifetime, it will be found in the next section that the value calculated for tc
agrees well with the measured lifetimes of pre-cavitated materials.

12.4 Comparison of Calculated TImes to Cavity Coalescence on


Isolated Facets with Measured Rupture Lifetimes
of Pre-Cavitated Materials
12.4.1

Rupture lifetime of prestrained Nimonic 80A

Figure 12.5 shows rupture lifetimes measured by Dyson

and

Rodgers

(1977)

on

Nimonic 80A at 750 0 C. The material had intentionally been pre-cavitated by prestraining at room temperature followed by a suitable

high-temperature

anneal.

This treatment provides a uniform density of small grain boundary cavitie! from
the beginning of the creep test with a cavity spacing of

A = 1.5

~m.

The

ex-

perimental results are compared with the time to cavity coalescence calculated
from eq. (12.11). The following parameters were used: = 2.10- 17 0 4 2 (in MPa
3
and sec), 6D bo = 10 -14 mis,
Qb ~ 170 kJ/mole, d = 50~m, A - 1.5~m (from Dyson,
1979) and h(~) = 0.61. Note that the grain boundary diffusion coefficient deviates

from

that of pure nickel given in Appendix A. The dashed lines in Fig.

12.5 represent the limits of fully constrained and


viously

the

unconstrained

growth.

Ob-

data points fall within the range of constrained growth and agree

well with the theoretical curve.

12.4

Comparison with Measured Lifetimes

183

Nirnonic 80 A
750C

70 6

eq. (12.11)

70 5

-..:::::-..:::::

II)

s.....

70~

......

unconstrained
10 3

700

750

"

"""""- ""--

200
(J

""-

------~

in MPa

300

1,00 500 600

--

Fig. 12.5. Rupture lifetime of pre-cavitated Nimonic 80A


measured by Dyson and Rodgers (1977).

Dyson and Rodgers (1974) varied the cavity spaCing by applying varying

amounts

of prestrain at room temperature. Dyson (1979) evaluated the results and showed
that the strain to failure, E f , increases in proportion to the cavity spacing.
This is predicted by the constrained growth model, but not by other models.
12.4.2

Rupture lifetime of prestrained Inconel alloy X-750

Pandey, Mukherjee and Taplin (1984) measured rupture lifetimes


specimens

ing amounts of room-temperature prestrain gave


strain-induced

prestrained

decreasing

bracketed

spacing

of

the

cavities. Figure 12.6 shows the rupture lifetimes as a function

of cavity spacing. The data pOint for material which had not
is

of

of Inconel alloy X-750 at a stress of 400 MPa and at 700 oC. Increas-

been

prestrained

since this material failed by plastic instability rather than by

cavity coalescence. The solid line represents the time to coalescence calculated

from eq. (12.11) in the constrained limit. In this case, the first term can

be neglected. The second term of eq. (12.11) can be

approximately

written

as

t f = 0.32(A/d)/E.

The grain size is d = 160~m in this material and the minim~


creep rate is E = 7'10- 9 /s at 400 MPa, practically independent of prestrain. As
Fig.

12.6 shows, the model overestimates the rupture life by a factor two, but

the linear dependence on cavity spacing is qualitatively confirmed.

184

12.

..c:

(.)

300

0%

250
200

L /

3% prestrain

..? 5%

.!; T50

........

Constrained Diffusive Cavitation

TOO

Inconel

~ 744%

50

9.5%

X-750,700e

0
0

cavity spacing

Ain I'm

TO

Fig. 12.6. Rupture lifetime of Inconel X-750. Various degrees of


prestraln give different spacings of pre-existing cavities.
Data from Pandey et al (1984). Solid line: eq. (12.11).

12.4.3

Rupture time of a-brass with implanted water vapor bubbles

Svensson and Dunlop (1981, 1982) have reported results of creep rupture tests
on a-brass at 700K (427 oC). Before the creep test, the material was given a
heat treatment which produced a distribution
typical diameters of 2R

0.3

~m

of

water

and a spacing of A = 6

vapor
~m

bubbles

having

in grain boundaries.

ct.-brass

T07

700K

10 6

I/)

.5::
........ 105

10'

10

20

40

fJ

uinMPa - -

Fig. 12.7. Rupture time of a-brass with implanted water vapor bubbles, from
Svensson and Dunlop (1981). Dashed lines: Unconstrained limit, eq. (11.18).
(After Riedel, 1985c).

12.4

185

Comparison with Measured Lifetimes

About 20% of the grain boundaries were cavitated. Some creep tests were
rupted

and

inter-

the specimens were sectioned in order to investigate the evolution

of the bubbles. Other specimens were crept to rupture giving the rupture
shown

times

in Fig. 12.7. Since the number of bubbles, or cavities was constant dur-

ing the tests, the lifetime should be given by eq. (12.11). All relevant material

parameters

and the strain rate are given by Svensson and Dunlop (1982) so

that eq. (12.11) can be evaluated. The result is


Svensson

and

shown

in

Fig.

Dunlop (1982) give void tip half angles, ~, between

both angles are inserted. Obviously the constrained model

12.7.
350

describes

Since

and 500
the

data

very well except at the highest stress, a = 63.7 MPa. The unconstrained diffus-

ive growth model leads to the dashed lines in Fig.

12.7,

which

underestimate

the rupture lifetime significantly at lower stresses.

12.5 Constitutive Behavior of Creeping Materials Containing Widely


Spaced Cavitating Grain Boundary Facets
Our primary interest so far has been to determine
function

of

the

cavities influences
called

the

applied

stress

the

macroscopic

and

growth

be

is

extended

which

response,

which

is

boundary

cavitation

is

described.

The

based on a theory proposed by Hutchinson (1983), which is valid


to

higher

dena

es

of

12.6,

the

theory

facets using Simple arguments of

self-consistency. This is a first approach to deal with


ions,

as

strain rate. Conversely, the growth of


stress/strain-rate

for a dilute concentration of cavitating facets. In Section

will

rates

constitutive behavior. In the following, the constitutive behavior

of solids undergoing diffusive grain


discussion

cavity

facet-facet

interact-

will control the final rupture process. Other constitutive models

are those of Harris, Tucker and Greenwood (1974), Skelton (1975), Crossland and
Harris (1979), Beere (1980a, 1981) and Raj (1983).
12.5.1

The constrained limit (Hutchinson's model)

In the limit of accommodation-controlled cavitation, the normal stress


cavitating

facets

has

dropped

to

the

order

on

the

of magnitude of the sintering

stress. Thus, if the small sintering stress is neglected, the cavitating facets
can

be considered as microcracks with traction-free crack surfaces. The effect

of microcracks on the overall moduli of elastic bodies

has

been

investigated

extenSively in the past (see, e.g., Mura, 1982, Horii and Nemat-Nasser, 1983).

186

12.

Constrained Diffusive Cavitation

The constitutive behavior of nonlinear viscous solids


microcracks

has

been

studied

by

Hutchinson

containing

(1983).

He

penny-shaped

assumes

that the

cavitating facets are well separated from one another, so that interactions can
be neglected, and that they are aligned normal to the principal tensile stress,
or. For a matrix material obeying Norton's power law, eq. (3.6), Hutchinson
derives the following macroscopic stress/strain rate relation for the material
containing a dilute concentration of microcracks:

0i'J

p_
+ __
n+l

[( n-1) 0l!j ( -r)2 + -3


4 or m.. J}
0e
lJ

( 12.12)

The tensor mij is defined such that its components in the co-ordinate system of
the prinCipal axes of stress are zero except in the maximum pricipal stress
direction where mr = 1. The term containing the factor
the

microcracks

to the overall strain rate;

is the contribution of

is a measure for the density of

the microcracks and is defined by:


p

(1/2) d3 N (n+1) (1+3/n)-112,


mc

(12.13)

where d is the diameter of the microcracks and Nmc is the number of microcracks
per unit volume.
According to eq. (12.12), the presence

of

the

microcracks

adds

dilatant

component to the strain rate (the second term in brackets), but the response to
deviatoric stresses is also modified (the first

term

in

brackets',

and

the

latter modification is often the more important one.


For the following discussion, it suffices to consider Hutchinson's constitutive
model for uniaxial tension in which case eq. (12.12) reduces to

( 12.14)

Here,

is the steady-state strain rate which would prevail in the absence

the microcracks, s
12.5.2

Bon.

The unconstrained limit

Outside the constrained limit, cavitating grain boundaries


tensile

of

stress,

0b'

which

was

given

in

eq. (12.3).

transmit

finite

Hutchinson's

(1983)

12.5

Constitutive Behavior

187

constitutive model, which was originally developed for ab = 0, can be extended


to finite ab by replacing the maximum principal stress, aI' by aI-abo The unconstrained limit implies that ab + aI' so that [(aI-ab)la e ]2 can be neglected
compared to (aI-ab)la e Inserting ab from eq. (12.3) into eq. (12.12) and
taking the unconstrained limit leads to the constitutive law
( 12.15)

which is valid in the unconstrained limit, i.e.


tension, this reduces to

for

oD b

small

In

uniaxial

(12.16)

where, additionally, the sintering stress, ao ' was neglected, and p was replaced by the volume density of cavitating facets, Nmc ' according to eq. (12.13).
The result given in eqs. (12.15)
additional

extensional

and

(12.16)

above

simply

growth rate of the cavities contained in the unit volume of


constrained

limit

means

that

the

strain rate due to the cavities is given by the volume


material.

In

the

this was not so: the presence of the cavitating facets also

affected the deviatoric response of the material.


12.5.3

The effect of cavitation on diffusion creep

To explore the effect of cavitation on the rate of diffusion creep, we model


cavitating

grain boundary facet as a penny-shaped crack embedded in a linearly

viscous continuum.
continuum

T~is

process

ignores

the

fact

that

diffusion

creep

advantage

not

this

has

that diffusion creep becomes a special case of power-law creep,

and the constitutive equations for cavitating solids


preceding

is

on the size scale of single grain facets. If, nevertheless,

the continuum-mechanical description is adopted as an approximation,


the

sections.

Whereever

can

be

taken

from

the

the stress exponent, n, appears it must be set

equal to unity, while B is replaced by ~noDb/(kTd3) according

to

eq.

(1.5).

Equation (12.12) then becomes


(12.17)

12.

188

Constrained Diffusive Cavitation

Apparently, the effect of cavitation on linearly viscous


augment

the

creep

is

simply

to

extensional strain rate by the volume growth rate of the cavities

contained in the unit volume of material. In eq. (12.17) above,

0b

is

under-

stood to be given by eq. (12.3) with n = 1, and p is defined in eq. (12.13).

12.6 Interaction Between Closely Spaced Cavitating Boundary Facets


If cavitating facets are relatively closely spaced, such that
from

one

their

distances

another are comparable to, or not much greater than, their diameter,

then the creep flow fields around the facets start to interact. In general, the
interaction

will

enhance

cavity

growth. The interaction can be viewed, very

roughly, as a consequence of the reduction of the load carrying area by

cavit-

ating facets, which, as we have seen, behave like microcracks in the constrained limit. Such a net-sect ion-stress argument, however, does not lend itself

to

a quantitative description of interaction effects. As the results derived below


indicate, a net-sect ion-stress argument overestimates the
substantially.

It

will

also

interaction

effects

be found that the importance of the interaction

effects depends greatly on whether or not the grain boundaries between the cavitating

facets

can

slide.

As an extreme example, consider a two-dimensional

hexagonal array of grains as shown earlier in Fig. 7.4. If in such a configuration,

all boundaries that are inclined to the tensile axis slide freely, while

all normal boundaries cavitate, it is obvious that there exists


on

no

constraint

cavitation since the volume increase of the cavitating facets can be accom-

modated by grain boundary


models,

sliding

alone.

More

realistic,

three-dimensional

however, exhibit a polycrystalline constraint even if grain boundaries

slide freely. Such models are described in Section 12.6.3.


12.6.1

Self-consistent models for constrained cavitation

An approximate way of including interaction effects is the well-known self-consistent method, which is applied here to the constrained limit, i.e. the cavitating boundary
crack

is

f~cets

analyzed

are idealized as penny-shaped

microcracks.

An

isolated

as in Hutchinson's (1983) model (Section 12.5.1). The sur-

rounding material, however, is not given the properties of the parent material,
but is described by the as yet unknown constitutive law of the solid containing
a not necessarily dilute concentration of microcracks. From the
deformation

field,

the

so

calculated

contribution of the crack to the overall displacement

rate can be obtained. Finally, the self-consistent argument is

closed

by

re-

12.6

Interaction Between Facets

Quiring

that

this

189

overall material response must be identical to that of the

cracked material assumed initially.


The self-consistent method is applied

now

in

way

which

is

not

exactly

correct. We ignore that the presence of microcracks introduces a dilatancy into


the overall material response and additionally an anisotropy since
are

the

cracks

aligned in Hutchinson's model. This means to describe the cracked material

as an incompressible, isotropic von Mises solid, which


cracked

parent

material

differs

from

the

un-

only by the coefficient B in Norton's creep law. The

self-consistency argument then implies that the contribution of the microcracks

PE

to the uniaxial strain rate is

instead of

PE s ,

E is

where

the overall strain

rate of the cracked material. With this substitution, eq. (12.14) resolved

for

~ takes the self-consistent form:

(12.18)

e:

For small p, eq. (12.18) reduces to the dilute limit, eq.


increasing

p,

(12.14),

cracked

mat"erial

described as an incompressible von Mises solid. The accuracy of eq. (12.18)

can be

assessed

(1984a,b)
p

for

eq. (12.18) can only be an approximation, since self-consistent

methods are necessarily approximate and, moreover, since the


is

while

has

by

comparing

carried

out

for

with

numeric

periodic

calculations,

arrays

of

which

cracks.

Tvergaard

For n

5 and

0.38, Tvergaard (1984b) reports a macroscopic strain rate which exceeds the

value predicted by eq. (12.18) by only 1%.


An alternative self-consistent approach is the differential scheme described by
McLaughlin

(1977)

and

by Duva (1984). If again the simplifying assumption is

made that the cracked material is an isotropic von Mises solid, one
from

the

can

start

uniaxial form of the material law for the dilute limit, eq. (12.14).

Differentiating eq. (12.14) with respect to p leads to


( 12.19)
If an increment in crack density, op, is added to an already
characterized

by a density p and a strain rate

E,

must have the form of eq.

(12.19) but with e:


s
substitution and with the boundary condition that
of eq. (12.19) yields the macroscopic strain rate

cracked

material

the increment in strain rate


replaced

E=

E.

by

for p

With

this

0, integration

12.

190

Constrained Diffusive Cavitation

(12.20)
Beere (1980a) USing a different

method

analyzes

configuration of square
grains with all normal boundaries cavitating, i.e. with Nmc d 3 = 1. His result
(12.21)

is nearly identical with eq. (12.20). For n = 5, eq. (12.21) predicts that

the

strain rate is enhanced by a factor 12.6 due to cavitation .


The transverse strain rate,
from

ET,

occurring in a tensile test can

be

calculated

the condition that the trace of the strain rate tensor equals the rate of

relative volume increase, ~/V:


(12.22)
Inserting ~f from eq. (12.2) using the strain rate of the
gives

cavitating

material

(12.23)
The second term in the brackets accounts for the dilatancy

of

the

cavitating

material.
12.6.2

The penny-shaped crack in a finite cylinder

Cavitating facets are again idealized as penny-shaped microcracks. For


ience,

only

th~

completely

constrained

conven-

limit, 0b = 0, is considered. As an

approximate model for a distribution of cracks with an average center-to-center


spacing

A we consider a penny-shaped crack having diameter d and lying in the

center of a circular cylinder of finite


analyzed

diameter

A.

This

configuration

was

by He and Hutchinson (1983) numerically. Their results for the volume

growth rate of the crack can be approximated to within better than 10% by:

Vf

(facet in finite cylinder)

~f

(facet in infinite body)

exp [0.59 (n+1) (1+3/n)-1/2 (d/A)2j. (12.24)

The form of this approximation formula was


differential

self-consistent

scheme,

eq.

motivated

by

the

result

(12.20). The expression

of

(d/A)2

the
was

12.6

191

Interaction Between Facets

chosen instead of d 3Nmc ,since it is expected that


cylinder is a model for a co-planar array of

the crack in a finite


cracks rather than for a

three-dimensional array of cracks. The numerical factor 0.59 results from a fit
of the formula to the numerical results.
Numerically, the enhancement factor for the volume growth rate of the facet due
to

interaction effects is moderate. From He's and Hutchinson's (1983) results,

which are approximated by eq. (12.24), one obtains enhancement factors of 1.14,
1.45,

1.97

and

4.47 for n

1, 3, 5 and 10, respectively, and for d/A

i.e. for closely spaced facets. A net-sect ion-stress

argument

would

lead

1/2,
to

substantially greater enhancement factors. It is therefore considered to be an


inappropriate argument.
12.6.3

Interactions between closely spaced facets in the presence of grain


boundary sliding

Anderson and Rice (1985) consider the regular, three-dimensional


of

configuration

grains shown in Fig. 12.8. The grain boundaries are assumed to slide freely

and all boundaries which are normal to the maximum prinCipal

stress

cavitate.

The crack density in this configuration is Nmc d 3 = 0.5. The grains are modeled
as nonlinear viscous as described by eq. (3.6), which reduces to = Ban in
uniaxial

tension.

Based on a variational prinCiple, Anderson and Rice develop

the following approximation formulas for the volume growth rate of a cavitating
facet and for the macroscopic extensional and transverse strain rates

Strain rate

stress u

Fig. 12.8. Grain structure considered by Anderson and Rice (1985). First
orientation: hexagonal facets H are perpendicular to the applied stress and
are cavitated. Second orientation: square facets S are perpendicular to stress
axis and are cavitated.

12.

192

. 3.1 n d3
. 3.34 . 3.1 n .
. - 0.23

Vf

3.4

Constrained Diffusive Cavitation

(12.25)

ES

(12.26)

ES

(12.27)

E.

ET

Here, Es is the strain rate of the polycrystal with sliding grain boundaries,
but with no cavitation, i.e. ~ = B (fa)n. The stress-enhancement factor, f,
s
due to grain boundary sliding was defined in Section 1.5.2. (Anderson's and
Rice's model can also be applied to the case of sliding boundaries with no cavitation giving f
with

the

1.3 1.25 1/n ). It is interesting

corresponding

expression

for

to

compare

eq.

(12.25)

the isolated facet, eq. (12.2). From

this, the enhancement factor due to interaction effects is found to be

Vf

(closely spaced facets)

Vf

(isolated facet)

(12.28)

For n = 5, the interaction enhancement


greater

factor

is

now

1228,

which

is

much

than the interaction enhancement factor for no grain boundary sliding,

which is about 2 according to eq. (12.24).


Anderson and Rice (1985) analyze the configuration shown in Fig. 12.8 also
an

orientation

in

which

the

facets) cavitate and are aligned normal to the stress axis.


ration,
N

mc

d3

the

crack

density

for

square grain facets (rather than the hexagonal


is

smaller

than

in

the

In

this

first

configu-

case,

viz.

0.125. Instead of eqs. (12.25) to (12.27) the result is now:

Vf
E

ET

. 1.7n 3 .
.
= 1.68 . 1.7 n
.
- 0.39
2.9

(12.29)

ES

(12.30)

ES

(12.31)

E.

The stress-enhancement factor due to grain boundary sliding

for

uration

is

not

reported by Anderson and Rice except for n

From eqs. (12.29) and (12.2) the


obtained as 52 if n = 5.

interaction

enhancement

this

config-

1 where f = 2.9.

factor

for

~f

is

12.7

Time to Rupture for Interacting Facets

193

12.7 TIme to Rupture for Interacting Facets


There is no obvious fracture criterion available which terminates

the

opening

of the facet-size microcracks and leads to final failure. Cavity coalescence on


the facets has already been ruled out as a feasible criterion in Section
Other

possibilities

will

12.3.

be examined in the following sections. Their common

feature is that they all lead to a stress-independent Monkman-Grant product. In


Section

12.7.1,

the plastic collapse of the cracked polycrystal at strains of

order unity is considered. Further, necking of the specimen combined


creasing

microcrack

damage

with

in-

may terminate the creep test (Sections 12.7.2 and

12.7.3). Finally, cavities can be nucleated on

inclined

boundaries

near

the

facet-size microcracks. The microcracks can then grow by coalescence with these
cavities, a process analogous to creep crack growth on a small scale.
an

application

However,

of the concepts of creep crack growth described in Part III to

the special situation at microcracks is still lacking.


12.7.1

Failure by large strains

When the opening of the facet-si 7o


spacing,

the

strength

min~n,~~nVQ

)~ycrysla~

honnmos

of the

order

of

oreaKS aown since the load-carrying

area vanishes; the microcracks link up to form a crack. In the case of


spaced

microcracks

such

as

in

criterion

closely

Fig. 12.8, the opening of the crack, 6, must

become of the order of the grain size, d, to achieve final


this

their

failure.

To

apply

to Anderson's and Rice's (1985) model (Section 12.6.3), we use

the geometrical relation Vf = (~/4) d 6 and require 6 = d. For the first configuration analyzed by Anderson and Rice, Vf is given by eq. (12.25) and
failure is found to occur at the following value of the Monkman-Grant product:

s f

= 0.23/3.1 n

( 12.32)

Recall that s is the strain rate of the uncavitated parent material

subjected

to the stress o. Equation (12.32) gives extremely small values for the MonkmanGrant product, e.g. Est f = 0.08% if n = 5. For the second configuration analyzed by Anderson and Rice (in which the square boundary facets cavitate), Vf is
given by eq. (12.29) and the Monkman-Grant product becomes 1.9% for n = 5.
Anderson and Rice (1985) have carried out their

analysis

within

small-strain

theory of deformation neglecting geometry changes. Hence, the gradual weakening


of the polycrystal as the contact area between grains becomes

smaller

is

not

12.

194

Constrained Diffusive Cavitation

taken into account. This deficiency is avoided by Tvergaard (1985) who performs
a large-strain finite element analysis of an
cavitating

facet

is

aXisymmetric

conf"iguration.

The

modeled as a penny-shaped crack and the inclined, freely

sliding grain boundaries are conical surfaces through the crack front. A weakness

of this approach is that an axisymmetric model cannot accurately describe

a three-dimensional, more or less periodic array of cavitating facets. A creep


curve

computed by Tvergaard (1985) for uniaxial tension is shown in Fig. 12.9.

Up to strains of 40 to 80%, which corresponds to crack openings around 6


the

strain

rate

does

not

increase

beyond its initial value, i.e. geometry

changes do not reduce the strength of the


even

predicts

configuration

markedly

(the

within

model

small strengthening effect, i.e. a decreasing strain rate in

Fig. 12.9). At a certain strain, however, the structure starts to collaps


idly

d/2,

rap-

a short time, but with an additional increment of strain of 10 to

40%. Whereas the strain to fracture is rather large in Tvergaard's

model,

the

Monkman-Grant product calculated with the strain rate of the uncavitated material is relatively small. In the example shown in Fig. 12.9 is ~stf
12.7.2

6.6%.

Rupture lifetimes for continuous nucleation of caVitating facets

The number of grain boundary facets having relaxed normal stress is usually not
constant during a creep test. Lacking preCise information, we assume that their
number, increases in proportion to some power of strain

l.Or-----...,..------.----...........----,

a2

-----

O~~~~~------~--------~~

a02

aol,

a06

normalized time tst Fig. 12.9. Creep curve computed by Tvergaard (1985) for the axisymmetric
model shown in the insert. The dashed line characterizes creep of the
parent material with no grain boundary sliding and no cavitation.

12.7

Time to Rupture for Interacting Facets

or

195

a' Y

"

(12.33)

where p was defined in eq. (12.13). The dimensionless factor 6 and the exponent
Y reflect

the cavity nucleation kinetics; and 6' = (6/2) (n+1) (1+3/n)-1/2 is

an abbreviation. These parameters are related, albeit not uniquely, to the nucleation rate, a', of individual cavities which was introduced in Section 5.7.1.
The result of the
fracture

when

p =

self-consistent

model,

eq.

(12.18),

apparently

predicts

1. According to eq. (12.33), this condition is reached when

the strain reaches the critical value Ef = (1/6,)1/Y. By combining eqs. (12.18)
and (12.33), the creep curve for a solid with continuous cavity nucleation and
constrained growth can be determined. Integrating eq. (12.18) with eq.
inserted for constant
the form t = t(E):

(12.33)

ES ' i.e. for constant stress, gives the creep curve in

(12.34)

~ t

Figure 12.10 shows this result in the form E(t). The Monkman-Grant product is
Y

(12.35)

Y+1
Thus,

the critical strain and the

Monkman-Grant product are determined solely

0.12

t'
'S"

0.1
0.08
0.06

.!::

U)

00t.
0.02
0.01

0.02

0.03 QOt.

normalized time

i, t

0.05

---

FIg. 12.10. Secondary and tertiary stages of the creep curve


according to eq. (12.34) for various sets of parameters.

Constrained Diffusive Cavitation

12.

196

by the parameters
is

clear

that

and Y, which characterize cavity nucleation.

~'

the

described by a physical criterion such as cavity coalescence.


occurs
p

when

= 1.

the

Further,

it

occurrence of failure in the self-consistent model is not

denominator

Rather,

failure

in eq. (12.18) becomes equal to zero, i.e. when

Qualitatively, Fig. 12.10 describes tertiary creep of real materials.

The differential self-consistent


(12.20)

is

method

leads

to

combined with eq. (12.33), setting Y

similar

results.

If

eq.

1 for simplicity, the creep

curve is obtained as

In this model, rupture


tf

(~';s)-1,

the

occurs

when

+~.

The

rupture

time

is

given

by

strain to failure is infinite (i.e. necking will intervene

in practice), and the Monkman-Grant product is Est f = 1/~'.


12.7.3

The combined effects of necking and continuous nucleation

Hoff's (1953) analysis of necking which


extended

to

material

which

was

develops

(12.33) taking Y = 1 for simplicity. The


material

described
microcrack

in

Section

2.2.3

is

damage according to eq.

stress/strain-rate

response

of

the

is described by the differential self-consistent scheme, eq. (12.20).

Then the strain rate in uniaxial tension is:


(12.37)

where ES is the initial strain rate of the uncavitated material. Using the
factor (1 + E) implies that incompressibility is assumed. Since the material
develops dilatancy as p increases in

proportion

dilatancy

the

introduces

an

error

of

order

to E, the neglect of the


2 .
E
ES ln eq. (12.37). Since E

generally remains small over substantial fractions of the lifetime, this


appears

to

be

tolerable.

A proper

treatment

error

of the dilatancy renders the

analysis much more complicated.


Equation (12.37) can be integrated by separation of the

variables,

and

t,

giving the rupture lifetime, or the Monkman-Grant product as:


(12.38)

12.1

Time to Rupture for Interacting Facets

where En (8') is the exponential integral. Abramovitz


useful upper and lower bounds for En (8') such that

191

and

Stegun

(1968)

give

(12.39)
The upper bound is particularly simple illustrating how

necking

and

internal

microcrack damage combine to cause fracture. Recall that for necking alone, the
Monkman-Grant product is 1/n, while for internal damage alone, it is 1/8'.

12.8 Conclusions on Constrained Cavitation


1) Cavities in commercial materials grow at a rate which is compatible with the
constrained diffusive growth model.
2) Cavitating boundary facets behave like microcracks when

cavity

growth

be-

comes constrained. Therefore the further growth and coalescence of cavities,


after the constraint has become effective, are irrelevant

for

the

rupture

process.
3) Nevertheless, measured rupture lifetimes of pre-cavitated materials approximately agree with calculated times to cavity coalescence.
4) Despite this coincidence, other rate-controlling processes for

rupture

be-

sides cavity coalescence must be identified. Interactions between cavitating


facets are likely to play an important role in the failure process.
found

was

that interaction effects between closely spaced cavitating facets en-

hance constrained cavity growth rates by typically a


grain

It

factor

1.5

to

if

boundaries do not slide. Grain boundary sliding increases that factor

to values often greater than 100 the precise value depending strongly on the
stress

exponent

n and on the grain configuration analyzed. Increasing num-

bers of cavitating facets and their continuing opening,

sometimes

together

with necking, finally lead to a collapse of the polycrystal.


5) If it is true that the polycrystalline constraint is the cause for
observed

the

low

cavity growth rates, the number and size of cavities on cavitating

boundaries is not an appropriate measure for the state of damage in the material.

Rather the number of relaxed facets is important. For remaining life

assessments, one should attempt to determine this quantity experimentally.

13 Inhibited Cavity Growth

In this chapter, we explore the possibility that the low


observed

in

many

cavity- growth

alloys can be explained by an inhibition of stress-directed

diffusion, a phenomenon which is closely related to inhibited


(cf.

diffusion

creep

Sections 1.4 and 7.3.6). The inhibition of cavity growth may be caused by

the fact that vacancies which aggregate in the cavities are not
enough

rates

replaced

fast

by vacancy sources, as has been assumed so far. Second, experiments in-

dicate that high-melting particles in grain boundaries inhibit diffusion creep,


and may therefore also inhibit cavity growth. This is ascribed to a conjectured
difficulty to plate the diffusing atoms onto

the

particle/matrix

interfaces.

Both mechanisms will be treated by a common theory of inhibited cavity growth.

13.1 Inhibited Cavity Growth Rates


Ishida and McLean (1967) suggest that
cavity

growth,

as

well

the

vacancies

needed

in

of

the

grains

and

the boundary for a certain distance. Dislocation climb is associated

with the production (or annihilation) of vacancies.


thus

diffusional

as for diffusion creep, are produced by dislocations

which enter the grain boundaries during dislocation creep


climb

for

generated

The

number

of

vacancies

per unit time is assumed to be proportional to the strain rate

of the grains, ~. In terms of the thickening rate of the

grain

boundary,

un'

the maximum rate that can be maintained by this type of vacancy production is
u

(13.1)

and the volumetric and linear growth rates of a cavity that collects all vacancies from a grain boundary area A2 (A

2
V = A' A E.

cavity spacing) are


(13.2)

13.1

Inhibited Cavity Growth Rates

199

(13.3)

It remains to determine the constant of proportionality, A', which reflects the


nature

of

the vacancy production process. From models on a dislocation level,

Dyson (1979) concludes that A' should be


(1980b)

predicts

that

A'

should

of

order

grain

grain boundary. Since these microscopic models remain


prefer

to

relate

size,

while

Beere

scale with the spacing of particles in the


somewhat

ambiguous,

we

A'to the measured threshold stress for Coble creep, which

probably has the same microscopic origin as the inhibition of cavity growth

Free Coble creep is associated with a rate of grain boundary thickening, ~,


which is given by eq. (7.48). If that rate exceeds the rate allowed by vacancy
production, Coble creep starts to be inhibited. In analogy to Section 7.3.6, we
identify

the

threshold

stress for Coble creep, 0th' (although the inhibition


(13.1)

does not set-in discontinuously in this model) by equating un from eqs.


and (7.48). This gYves the desired expression for A' in terms of 0th:

(13.4)

AI

Here, the diffusion length,

~,

is calculated from

eq.

(4.17)

with

om = 0th.

Since 0th is nf the order of only 1 MPa, ~ is rather large, and A' is found to
be greater than the grain size in most cases.
If the inhibition of diffusion arises from the difficulty of plating atoms onto
the interface between high-melting particles and the matriX, and if the accommodation is achieved by power-law creep around the particles,
rate

is

the

displacement

given in eq.' (7.49). It has the same form as eq. (13.1), viz.

Iin

ex

e.

Hence, the above equations retain their validity. Further, the threshold stress
in eq. (13.4) above can be expressed by eq. (7.51), which gives:
AI

3 (~/4)2 (1 +3/n)-1 12 P/w~

= 1.5 p/w~

The stress range in which diffusive cavity growth is inhibited is


equating

the

free

diffusive

growth

(13.5)
obtained

by

rate, eq. (11.11), with the growth rate

allowed by vacancy production, eq. (13.2). This leads to a transition stress

(13.6)

200

13.

The sintering stress is neglected here and A' is

Inhibited Cavity Growth

expressed

by

the

threshold

stress through eq. (13.4). Taking ot = 50, w = 0.2, d/A = 10, Gth = 1.3 MPa and
n = 5, the stress at which cavity growth becomes inhibited is found to be
3 MPa.

Below that stress, growth rates are inhibited and they are proportional

to strain rate as in eqs. (13.2) and (13.3),


diffusive

growth

prevails

as

while

above

that

stress,

free

described by eqs. (11.11) and (11.14). If this

example is typical, the inhibition plays no role in most creep rupture tests.

13.2 Time to Cavity Coalescence and Time to Rupture


for Inhibited Growth
If all cavities are present from the beginning of the test, the time to
coalescence, t c ' is obtained by integrating eq. (1303) from R

0 to R

cavity
=

A/2:
(13.7)

Here it was assumed that rupture occurs once the cavities have

coalesced.

The

result exhibits the desired property that the Monkman-Grant product, ~tf' is
independent of stress.
Equation (13.7) is similar to the result of the constrained growth

model

[the

second term of eq. (12.11)]. The main difference is that eq. (13.7) contains A'
instead of the gr'ain size d. Since eq. (13.4) showed that A' is generally greater than d, we conclude that inhibited growth gives much shorter lifetimes than
does constrained growth. On the other hand, it was demonstrated in Section 12.4
that

eq.

(12.11) describes measured lifetimes approximately correctly. Hence,

the inhibited model will generally underestimate rupture lifetimes.


Despite the similarities between constrained and inhibited cavity growth
and

times

to

coalescence,

there

are

also

important

differences.

constrained growth, the stress on cavitating grain boundaries

is

rates
During

relaxed

but

this is not so during inhibited growth. Therefore the argument that constrained
growth cannot control the rupture lifetime (Section 12.3) does not
apply

to

inhibited

growth.

Another

difference

is

that

cavity

analogously
growth in

bicrystal specimens should not be constrained, whereas the inhibition should be


equally

effective in bicrystals and in polycrystals. Hence, the observation of

Raj (1978a) that cavity growth in copper bicrystals obeys

the

free

diffusive

growth law, while growth is greatly retarded in polycrystals, supports the constrained growth model, but is not explicable by the inhibited growth model.

14 Cavity Growth by Creep Flow of the Grains


or by Grain Boundary Sliding

Dimpled fracture at room temperature generally occurs by the initiation, growth


and

coalescence

of

voids at second-phase inclusions in the grains. The holes

grow by plastic flow of the surrounding matrix. Hancock


creep

cavities

(1976)

proposed

an idea is attractive at first sight since strain-controlled cavity


expected

to

lead

products

growth

is

to a stress-independent Monkman-Grant product. It will turn

out, however, that the plastic hole growth mechanism gives too
Grant

that

on grain boundaries could grow analogously by creep flow. Such

compared

to

measured

large

Monkman-

values. Nevertheless, the mechanism is

described in Section 14.1 below. In Section 14.2, a cavity growth mechanism

by

grain boundary sliding, which seems to have some relevance in metals with easily sliding boundaries, will be presented.

14.1 Hole Growth by Creep Flow of the Grains


To begin with, we consider an isolated hole in an infinite block
viscous

material

obeying

eq.

(3.6),

which

reduces

to ~

of

nonlinear

Bon in uniaxial

tension. Interaction effects between the voids will be analyzed later.


As a special case of a void growing in viscous material, the penny-shaped crack
was

analyzed

in

connection with constrained cavity growth. The volume growth

rate was given in eq. (12.2), which is a special case of eq. (14.1)

below.

In

fact, the volume growth rate of arbitrarily shaped voids in homogeneous stress
and strain-rate fields, for reasons of dimensional consistency, must have the
form:
~

A (411/3)

.CX>
R3 e

( 14.1)

Here, e is the equivalent tensile strain rate at large distances from the
void, R is the void radius as defined in Fig. 11.1, the factor 411/3 was intro-

202

14.

Cavity Growth by Creep Flow or Sliding

duced for convenience, since for spherical cavities


volume,

V,

and

(4~/3) R3

is

the

cavity

A is a dimensionless factor which depends on the current void

shape and on the ratios between the stress components. There are,

in

general,

five independent ratios for six independent stress components.


These dependencies are described in the following sections for various
cases.

The

aspect

of

solutions for the linearly viscous case are discussed.


closed-form

solution

special

the void shape is brought out in Section 14.1.1, where


In

Section

14.1.2,

for a two-dimensional void geometry is worked out, which

illustrates the effect of the stress exponent n. In Section 14.1.3, results for
spherical

voids

embedded

in nonlinear material and subjected to axisymmetric

stress fields are reported and compared with solutions for penny-shaped cracks.
Finally, in Section 14.1.5, a few remarks on void interaction effects are made.
The reader who is only interested in the void growth rate for the
rupture tests

usual

creep

under uniaxial tension is advised to proceed to the result given

in eq. (14.18) and the subsequent discussion. The results for triaxial

loading

of spherical voids and cracks are summarized in eqs. (14.17) and (12.2), and in
Fig. 14.1.
The shape of the voids is determined by the
growth

process

itself

and

surface

competition

diffusion.

maintain the equilibrium, lenticular void shape

between

Surface

the

diffusion

conserving

the

void

plastic
tends to
volume,

whereas creep flow leads to other void shapes and generally changes the volume.
Apparently, equiLibrium shapes will predominate for slowly growing
coupled

problem

of

plastic

void

growth

and

voids.

The

simultaneous shape changes by

surface diffusion will not be investigated, since it turns out

that

the

void

shape either has no great influence on the growth rate (as in uniaxial tension)
or that plastic flow does not lead to severe deviations

from

the

equilibrium

shape (as in the high triaxiality limit where growing voids remain approximately spherical).
14.1.1

The growth of isolated holes in linearly viscous materials

A comprehensive analysis of void growth in linearly viscous material


carried

out

has

been

by Budiansky, Hutchinson and Slutsky (1982). Exploiting the visc-

ous/elastic analogy, they apply Eshelby's (1957) solutions for the

deformation

fields around ellipsoidal voids in infinite elastic bodies. An important result


of Eshelby's analysis is that an ellipsoidal void in linear

material,

if

de-

14.1

formed

203

Hole Growth by Creep Flow

in a remotely uniform, but otherwise arbitrary stress field, retains an

ellipsoidal shape. However, the axes of the ellipsoid may rotate and
extend

at

different

rates.

This

implies

they

may

that the general solution for the

ellipsoid applies throughout the whole void growth history.


Budiansky et al (1982) consider the special case of a

spheroid

under

axisym-

metric loading with a~ being the tensile stress and a; being the transverse
stress as in Fig. 12.1. The remote strain rates are related to the stresses

-2E; =

through E~ =
described by

B (a~ - a;)

E; =

and

IE~I. The surface of the spheroid is

(14.2)
where R1 and R are the semi-axes in

the

tensile

and

transverse

directions,

respectively, and wand r are coordinates in these directions. The aspect ratio
of the spheroid is defined as A = R1 /R, i.e. A = 1 corresponds to a sphere.
The principal results for the volumetric void growth rate
change as expressed by

--=

Aare

and

for

the

(3/2) (a:la"") (2A2 +1) + 1 - A2

~"" V

2A2

shape

(14.3)

+ 1'>( 1-A2 )

(14.4)

where a:

tion for a""

(a;

2a;)/3 is the mean (or hydrostatic) stress, a is an abbreviaI a""I = ae'" , and I'> is defined as

"" i.e.
ar"" - aT'

if A <
I'>

if A

Because of the symmetFY of the void shape and


rotation

of

of

the

(14.5)

> 1.

loading

conditions

no

the void axes occurs. Note that eq. (14.3) is compatible with the

general form of eq. (14.1) with A being given by the

right-hand

side

of

eq.

(14.3)

and

(14.3) multiplied by A.
For a momentarily spherical void is A
(14.4) reduce to

1 and

I'>

= 2/3. Hence eqs.

Cavity Growth by Creep Flow or Sliding

14.

204

90:
40

ce V

...

...
...

01 + 20;

3
II

....

( 5/2) c

(14.6)

...

01 - T

(14.7)

Hydrostatic tension or compression is

degenerate

case

ce = O. The result for the spherical void then becomes

VIV

since

A=

0 and
(14.8)

(9/4) B

The limiting case of a penny-shaped crack corresponds to A

O. In

0, Il

this

case, eq. (14.3) gives


~.
v

[ c...e

(41[R3
13) ]

...
= ( 3 0...1 /1[0),

which agrees with eq. (12.2). For uniaxial tension


which

is

by

in

this

specializes

to

3/1[,

factor 4/1[ larger than the growth rate of a sphere having the

same radius Ceq. (14.6) with 0:/0'" = 113 for


that

(14.9)

uniaxial

uniaxial

tension].

We

conclude

tension the aspect ratio has no great influence on the void


V/[E; (41[R 3 /3)]. For hydrostatic tension, the com-

growth rate if normalized as


parison

of

eq.

(14.8)

with eq. (12.2) shows that the effect of the shape is

greater, i.e., the volume of a spherical void grows by

factor

31[/4

faster

Budiansky et al (1982) have studied the evolution of the volume and the

aspect

than that of a penny-shaped crack with the same radius.

ratio of an initially spherical void by integrating eqs. (14.3) and (14.4) over
the time. They find that, depending on the stress ratio 0;/0;,

the

voids

can

deform into prolate or oblate spheroids, including the asymptotic forms of long
cylinders, needles and cracks. In uniaxial tension, for example,
spherical

void

is

pulled

out

initially

into a needle. Its volume approaches a finite

value for large strains asymptotically, namely 1.264 times the


of

an

initial

volume

the undeformed sphere. If an arbitrarily small transverse tensile stress is

superimposed, the asymptotic void volume

tends

towards

infinity,

while

the

asymptotic shape is still a needle as long as 0;/0~ < 1/4.


14.1.2

An isolated circular-cylindrical void in nonlinear viscous material

For nonlinear material, Eshelby' s (1957) solution does no longer apply. Direct
integration

of

the

equilibrium

and

compatibility

conditions together with

14.1

205

Hole Growth by Creep Flow

Norton's power law is possible for circular-cylindrical voids under axisymmetric loading (Budiansky et aI, 1982). The radius of the cylinder is R. To be
compatible with the notation in the preceding section, the stress along the
axis is denoted by o~, while the radial, or transverse, stress is

cylindrical
CD

aT' According to Norton's law, eq. (3.6), the corresponding remote strain rates
.. ''I n- 1
....
...
ICDI
are e:...I - - 2'"
e: T B I aI-aT
(aI-aT)
and e:e
= e: I
In a cylindrical coordinate system (r,
axis

e, z) with the z-axis aligned with the

the cylindrical void, the only non-zero displacement rates are ur and
z The strain rates are calculated from the displacement rates according to

of

(14.10)
Incompressibility requires:
( 14.11)
In cylindrical coordinates, the equilibrium condition takes the form
(14.12)
Together with the material law, eq. (3.6), and the boundary conditions (orr

0;

at r = R, and orr =
at infinity) these equations can be integrated by
elementary algebraic operations to give the stress and strain-rate fields in
the

Whole

block.

In

particular,

radial displacement rate at r

the

void growth rate is obtained from the

R as:
(14.13)

where A is implicitly given by the relation


(14.14)

For linearly viscous material (n 1), the integral

can

be

evaluated

easily

giving
( 14.15)

14.

206

Cavity Growth by Creep Flow or Sliding

where the hydrostatic (or mean) stress om and 0


following

eq.

(14.4).

For

perfectly

plastic

were defined in the

paragraph

material (n = m), eq. (14.14)

reproduces McClintock's (1968) result:


(14.16 )
m

As the second form shows, the hydrostatic component of the stress, om' enhances
the

void

growth

rate

exponentially

in

the

perfectly plastic limit. Thus,

although the hydrostatic component of stress does not affect the remote
rate

(only

the

stress

deviator

strain

enters into the creep flow law), it greatly

enhances the void growth rate in nonlinear materials.


14.1.3

Spherical voids in nonlinear material under axisymmetric loading.


Comparison with penny-shaped cracks

A spherical void with radius R is embedded in an infinite


viscous

material

and

is

subjected

block

of

power-law

to axisymmetric loading, i.e. one of the

remotely applied principal stresses is 0; and the two others are 0;. This case
has been investigated by Budiansky, Hutchinson and Slutsky (1982). Following
Hill (1956b), they cast the constitutive equations

of

the

nonlinear

viscous

material into a variational principle and seek approximate solutions. Guided by


physical intuition they assume an analytical expression
rate

field

containing a set of unspecified parameters

for
~

with the requirement of incompressibility. In the spirit of

the

displacement

and being consistent


the

Rayleigh-Ritz

method,

the parameters a k are then determined by minimizing the functional of


the variational principle with respect to the parameters. In general, the mini-

mization procedure must be carried out numerically.


Numerical results for the normalized void growth rate are given
as

function

of

(14.4).

Uniaxial

implies

0"'/0'" = '"
m

represent the

0"'/0'"

to
and pure shear means o'"

tension

in

Fig. 14.1a

with aCIJ and at: as defined in the text following eq.

corresponds

om/o'"
m

1/3,

hydrostatic

tension

O. The dashed lines in Fig. 14.1a

limit (0"'/0'" ~ m). As this limit is approached,


m
spherically symmetric term, ~ ~ a Ir 2 , in the displacement rate

hi~h-triaxiality

all but the


r
0
field become negligbly small so that the minimization needs to be

carried

out

with

respect to a o only. This implies that the voids preserve their spherical
shape during growth in highly triaxial stress fields. The minimization can be
carried

out,

making

another algebraic approximation, in closed analytic form

(Budiansky et aI, 1982) with the result that

14.1

Hole Growth by Creep Flow

207

30

'e"

3'----"---'---r"T"""J"7"""T7JI

bJ

aJ
spherical
void

25
20

<1>'

..c::
~

15

c:
0

.iii

c:

10

uniaxial tension

c:

o~--~----~----~----~

1.5

05

aCIJ/a(X)= aoo/aOO +213 _


I
e
m

triaxiality, am(X)/ aOO_

Fig. 14.1. Normalized void growth rate A = ~/[(4~R3/3)~J as


a function of stress triaxiality for a) spherical void, and b) penny-shaped
crack. Dashed lines are analytical results, eqs. (14.17) and (12.2).

~ { 3 1a;1 + (n-1) (n+0.4319) }n .

2na

This result is valid for large positive a:/a~.


range

of

validity

depends

(14.17)

n2
(Recall

that

a; = la~I).

on n (cf. Fig. 14.1). For n = 1, it is exact. For

large negative a~/a~, the sign of the right-hand side must be inverted and
constant

0.4319

must

Fig.

14.1

bring

out

the

replaced by 0.4031. Depending on the sign of ~; and

be

a~/a~, eq. (14.17) describes growing or shrinking voids. Equation


m

Its

the

(14.17)

and

strong influence of the triaxiality of the applied

stress on the void growth rate for large n.


The limiting case of purely hydrostatic tension or compression is a
case

since

the

remQte

strain

rate

is

degenerate

zero. Equation (14.17) must then be

replaced by the following closed-form solution which was derived

by

Budiansky

et al (1982) for the spherically symmetric case:


(14.18)
In contrast to the other results described here, eq. (14.18) was derived for

14.

208

spherical

Cavity Growth by Creep Flow or Sliding

cavity in the center of a finite, spherical body having the diameter

A. To make contact with the rest of the section, in which infinite


considered, the limit AIR

bodies

can easily be taken. A finite sphere was analyzed

+ ~

by Budiansky et al as a model for interacting voids having an inter-void


ing

A.

Equation

(14.18)

shows

volume

fraction

of

voids

as

0.001 enhances the void growth rate by 37 per cent if n = 3

as (2R/A)3

and by a factor 4.2 if


hydrostatic

spac-

that interaction effects may be large in the

hydrostatic limit even for small RIA-ratios. A


small

are

n =5

compared

to

an

isolated

void.

However,

the

limit is an extreme case, and interaction effects are much smaller

under uniaxial tension, as will be shown in Section 14.1.5.


For comparison the volume growth rate of a penny-shaped crack is shown in

Fig.

14.1b. The dashed lines represent the analytic approximation, eq. (12.2), which
has been obtained by He and Hutchinson- (1981)
whereas

using

linearization

method,

the solid lines are numerical solutions developed by the same authors.

It is apparent that for uniaxial tension the volume growth rates

of

voids

Thus a Simple

and

cracks

of

having

the

same

radius

are

similar.

spherical

approximation formula, which is valid to within 35% in uniaxial tension for all
n and for all void shapes between a sphere and a crack, is
(14.19)
where E is the tensile Strain rate. In particular, this formula will be approximately valid for lenticular, equilibrium void shapes.
The case of a spherical void in a remotely uniform,
stress

but

otherwise

by Rice and Tracey (1969), but only for perfect plasticity (n


loading

case

for

that

for

spherical

irrelevant). The most convenient


however,

consists

The general

~).

spherical void can be characterized by three independent

stress components, for example, by the


(Note

arbitrary,

field, as distinct from the axially symmetric case, has been considered

of

the

void

remotely
the

triple

equivalent

applied

direction
of

of

independent

tensile

strain

principal

stresses.

the principal axes is


loading
rate

.=
Ee'

applied hydrostatiC stress to applied equivalent tensile stress,


so-called Lode variable

parameters,
the ratio of

0:/0:,

and the

(14.20)
.m

.~

where EI > Ell> EIII are the principal components of the remote

strain

rate.

14.1

Hole Growth by Creep Flow

209

The value of vL lies between +1 and -1. For aXisymmetric fields is vL

1 with

the plus sign for negative ;1 and the minus sign for positive ;1' In terms of
these three loading variables, the growth rate of a spherical void in a
perfectly plastic matrix is given by Rice and Tracey (1969) in the form

The result shows that under most loading conditions, the Lode variable

has

no

great influence on the void growth rate. For uniaxial tenSion, for example, the
first term of eq. (14.21) is 32 times greater than the second one

and

in

the

high-triaxiality limit this ratio is 1.674/0.024 = 70. It can be concluded that


the growth of a spherical void in a perfectly
determined

0:/0:.

by

axisymmetric

plastic

material

is

primarily

and by the ratio


In other words the analysis of the
is sufficiently general for most purposes. The third

case

variable, vL' can be neglected except for cases in which the hydrostatic stress
component is very small, say if
< 0.05. Also for arbitrary n, the Lode
variable
will have no great influence on the growth rate, since the

1:1/:

axisymmetric results for positive and negative o:/o m , which corresponds to


extreme values vL
14.1.4

the

1, are only slightly different.

Strain to failure neglecting void interaction effects

The volume gruwth rate of a void by creep flow is


equilibrium

void

shapes

(11.13), so that the

~inear

is

(4w/3) heW) R3

given
with

by

eq.

(14.1).

For

heW) as defined in eq.

growth rate becomes:


.m

R A Ee 1 [3 hew)] .

(14.22)

This can be integrated in time, which leads to


(14.23)

Here EO is the strain at which the void is nucleated and Ro is the void radius
at nucleation. The second form of eq. (14.23) applies if A is time-independent.
A time-independent A obtains for proportional loading and
does

not

if

the

void

shape

vary during growth. Rupture occurs when neighboring voids touch each

other, i.e. when 2R = A, where A is the void spaCing. Inserting this


into eq. (14.23) gives the strain to failure

condition

14.

210

Cavity Growth by Creep Flow or Sliding

(14.24)
In the second form, the strain to failure is expressed in terms of

the

voided

area fraction at nucleation, Wo0


If the strain EO' at which voids are

nucleated,

is

neglected,

the

critical

strain, E f , is independent of the stress level (A depends on stress ratios


only). This agrees qualitatively with the empirical Monkman-Grant rule. However,

the

absolute

values of the critical strain calculated from eq. (14.24)

are far too large compared with measured values. Assuming that A/2Ro
10 and
= 0.61, and using A = 1.1 for uniaxial tension, the critical strain is

h(~)

predicted to be 383%. Measured strains to rupture are typically a few per

cent

under cond1tions where 1ntergranular cavitation predominates.


Several hundred per cent strain to failure as predicted here
tests

in

are

typical

for

the superplastic regime. These high ductilities are achieved even in

the presence of voids (see, for example, Stowell, Livesey

and

Ridley,

1984).

The relevant difference to creep rupture tests is probably that the strain rate
in superplastic deformation, although low compared to usual tensile
higher

than

in

tests,

is

creep rupture tests, so that diffusion does not contribute to

cavity growth substantially. Since also necking does not occur

in

the

super-

plastic regime, it is conceivable that high strains are achieved.


Also dimpled fracture at room temperature is
growth.

This

fracture

mode

is

brought

about

by

plastic

void

often associated with high strains, and they

would be even higher did necking not intervene. Thus, the analysis presented in
this chapter possibly applies to superplastic materials and to dimpled fracture
at room temperature, apart from the necking instability, but

probably

not

to

typical creep rupture conditions.


14.1.5

Void interaction effects

As was indicated above, the growth of isolated voids by creep of the


ing matrix leads to ductilities which are too high compared to
in

creep

those

surroundobserved

rupture tests. One might suspect that void interaction effects could

explain the discrepancy. However, several analyses described next indicate that
void interaction effects are moderate under typical creep rupture conditions.
This point of view is supported by a finite element

analysis

carried

out

by

14.1

211

Hole Growth by Creep Flow

Needleman

and

Rice

(1980).

They model a planar distribution of axisymmetric

voids under uniaxial tension by imposing symmetric boundary

conditions

on

an

appropriate unit cell. Diffusive void growth is included in that model, but the
limit of void growth by creep flow alone can readily be evaluated. The
indicate

that

up

to

an area coverage of at least w

results

0.1 the interaction is

negligibly small. Thus the strain to fracture can hardly deviate markedly

from

eq. (14.24).
A simple, although not very well defined, approach to account for
action

effects

is

void

inter-

based on a net-section-stress argument. Here it is assumed

that, owing to the reduced load carrying area, the stress on a cavitated
boundary is enhanced by the factor 1/(1-w) where w
ion of voids. Therefore, after replacing
eq.

(14.28),

can

be

integrated

Wo

(2Ro/A)2, to coalescence, w

from

i em

an

initial

Wo

for

integer

n.

the void growth

law,

cavitated area fraction,

(14.25)

dw .

The integral on the right-hand side can be evaluated


functions

(2R/A)2 is the area fract-

i em /(1_w)n,

1, giving the Monkman-Grant product:

3 he lP) /1 (1-w)n
2 A

by

grain

Figure

in

terms

of

elementary

14.2 shows the result as a function of the

voided area fraction Wo at void nucleation.

Although the Monkman-Grant product

1.5 .------"T1rr-'C'"""T---,---,--.--""T'""-.,...-----,

L-~_-L_~

002

__

0.06

~_L-~

0.10

_ _~_~

a It,

Fig. 14.2. The integral shown is proportional to the Monkman-Grant product

according to eq. (14.25). The factor of proportionality is unity if


A = 1.1 and helP) = 0.73.

Cavity Growth by Creep Flow or Sliding

14.

212

can be reduced considerably by the net-sect ion-stress


the

effect

example, if n = 5, the initially voided area


Wo

(corresponding to 2Ro/A

0.16

argument

applied

here,

is insufficient to explain the low observed creep ductilities. For


fraction

must

as

be

large

as

0.4) if the Monkman-Grant product is to be

25~. If more realistic values for Ware chosen, say W < 0.01, then the
preo
0
dicted Monkman-Grant product is well above 100~. Again this is a result which

is appropriate for superplastic materials but not for creep rupture conditions.
Incidentally, self-consistent estimates of interaction
Chen

and

Argon

(1979)

lead

to

effects

like

that

by

similar result as the net-section-stress

argument employed above.


As a last analysis which shows
finite

element

analysis

that

of

Burke

interaction

effects

creeping

matrix.

Burke

size-to-spacing ratio of 2Ro/A


ally

high,

least

45~.

they

find

moderate,

the

and Nix (1979) is quoted. They consider a

square grid of cylindrical voids with initially circular


power-law

are

and

cross

section

in

Nix start their calculation at a void

0.25. Although this appears to be unrealistic-

that interaction effects are small up to strains of at

This is so although the two-dimensional model is expected

to

over-

Finally, it should be mentioned that the formation of shear bands, which

some-

estimate interaction effects.

times

ter~inates

the

deformability of cavitating metals at room temperature,

probably plays no role under creep rupture conditions, since shear bands do not
develop

until

the strain reaches values of typically more than

(see, for

100~

example, Pan, Saje and Needleman, 1983).

14.2 Cavity Growth by Grain Boundary Sliding


As is illustrated in Fig. 14.3, grain boundary sliding
growth

is obvious that sliding sheds load on a cavitating


remarked

contributes

to

cavity

in two different ways. The first possibility is shown in Fig. 14.3a. It


in

connection

with

constrained

boundary,

as

was

already

diffusive cavity growth in Section

12.1.3. Under normal tensile stresses, this only means that cavity growth rates
are

enhanced, but under a transverse compressive stress the cavities would not

grow at all unless sliding took place. Sliding transposes the transverse stress
into a wedging force pushing the grains adjacent to the cavitating facet apart.
This mechanism may have operated in the exper1-ments of Davies and Dutton (1966)

14.2

213

Cavity Growth by Grain Boundary Sliding

I O'l =0

0)

b)

d
boundary
Fig. 14.3. Cavity growth by grain boundary sliding.
a) Compressive transverse stress. b) Growth at a tensile ledge.

and Davies and Williams (1969). They applied tensile stresses to copper and
copper/aluminum alloys until the specimens developed cavities and reached the
tertiary stage. Then the stress was either reversed, which led to at least
partial closure of the cavities, or a compressive stress was applied at 90 0 to
the original tensile direction. In the latter case, the cavities

continued

to

grow, which is expected from Fig. 14.3a. To quantify cavity growth under transverse compression, we assume that diffusive growth occurs so easily that it

is

constrained. Then the analysis of the problem shows that the cavity growth rate
has the same form as the constrained growth rate in tension, eq. (12.5),
from

numerical

factor,

and

a~

apart

must be replaced by the shear stress IT~I

acting on the inclined boundaries.


The second mechanism by which sliding contributes to cavity growth is shown

in

Fig. 14.3b. A cavity located at a tensile ledge is pulled out by grain boundary
sliding. If surface diffusion is slow (which we assume), the cavity growth rate

is directly equal to the rate of sliding, 2R = U


b , and the cavities coalesce
when the sliding offset reaches the spacing between tensile ledges. This mechanism is expected to predominate on boundaries which are inclined to the tensile
direction. It was proposed by Gifkins (1956), Chen and Machlin (1956) and Evans
(1971,

1984).

Evidence for its occurrence results from the bi-crystal experi-

ments of Chen and Machlin (1956) andfrom the observation of


(1980)

on

copper

Sklenicka

et

al

showing that the average cavity growth rate and the average

sliding rate are very nearly equal.

14.

214

Cavity Growth by Creep Flow or Sliding

The kinetics of this cavity growth mechanism is determined

by

that

of

grain

boundary sliding. A grain boundary embedded in a power-law creeping polycrystal


slides at a rate given by eq. (7.11). If, further, the shear stress transmitted
nUb' with the viscosity n,
by the boundary obeys the linear viscous law Tb
the sliding rate, and hence the cavity growth rate, are found to be:
(14.26)

= 10,

The numerical factor 0.89 is valid for n = 5; for n


is

the diameter of the sliding boundary facet,

on the facet,
valent

T~

0.84;

is the resolved shear stress

a: is the applied equivalent tensile stress, and : is the

equi-

strain rate of the polycrystal. Equation (14.26) has two limiting rang-

es. For large strain rates, the second term in the


sents

it would be

the

polycrystalline

constraint

on

denominator,

sliding,

which

repre-

is negligible, while the

intrinsic viscosity of the boundary, n, dominates. For small strain rates,

the

constraint beeemes effective and the intrinsic viscosity becomes negligible. In


this limiting case, the sliding rate is controlled by the strain rate:
1.1

0.55

(14.27)

d.

The second form is valid for uniaxial tension and T~/a:

1/2.

It is not surprising that the result resembles that for the constrained
growth

cavity

mechanism described in Chapter 12. Both arise from polycrystalline con-

straints. Comparison of eqs. (14.27) and (12.5) shows that constrained

diffus-

ive cavity growth rates are larger by a factor of the order (A/2R)2 than growth
rates by constrained sliding. Therefore constrained diffusive
be

the

dominant

fracture

cavitation

will

mechanism except when no cavities are nucleated on

boundaries that are oriented normal to the applied stress.

15 Creep-Enhanced DitTusive Cavity Growth


and Elastic Accommodation

Diffusive cavity growth was analyzed in Chapter 11 under

the

assumption

that

the adjacent grains were rigid. Now the consequences of grain deformability for
the cavity growth rates are explored. In Sections

15.1

and

15.2,

power-law

creep and elastic deformation, respectively, are considered.

15.1 Cavity Growth by a Coupling of DitIusioD and Power-Law Creep


In the rigid-grain limit
Chapter

11,

atoms

of

leaving

diffusive
the

cavitation,

which was

described

in

growing cavities must diffuse halfway to the

neighboring cavities since only a uniform thickening of the grain

boundary

is

compatible with the assumption of rigid grains. If the grains are deformable by
creep, the diffusion distance may be shorter, since the then nonuniform

thick-

ening of the grain boundary can be accommodated by larger creep deformations in


the vicinity ot the cavity. The dimensional considerations described in Section
4.5

suggest

that

the

diffusion

distance in the presence of creep is of the

order of the diffusive length R. which was defined in eq. (4.1-7). If R.


smaller

than

the

cavity

half-spacing,

growth substantially, and the limit R./A


flow,

which

was

is

much

A/2, creep enhances diffusive cavity

0 corresponds to hole growth by creep

despribed in Chapter 14. If, on the other hand, R. is greater

than A/2, the possibility of creep does not reduce the diffusion

distance

and

therefore does not enhance diffusive cavity growth markedly.


15.1.1

Models for the interactive growth mechanism

A quantitative analysis of the combined processes of

diffusion

and

creep

in

cavity growth is not simple. Beere and Speight (1978) were the first to present
an approximate model of the coupling which turned out to yield reasonably accurate

results when compared with later numerical studies to be described short-

ly. Other approximate models were proposed by Edward and Ashby

(1979)

and

by

216

15.

Cocks

and

Ashby

(1982),

Diffusive Growth with Creep or Elastic Accommodation

which,

however, do not reproduce the limit of hole

growth by creep flow correctly and therefore deviate

from

the

numerical

re-

sults.
Such numerical finite-element solutions were developed by
(1980)

and

by

Sham

Needleman

and

Rice

and Needleman (1983) who analyzed an axisymmetric, equi-

A
prescribing appropriate boundary conditions on the surface of the cylinder.
Chen and Argon (1981c) found the following surprisingly simple interpolation
formula, which accurately reproduces Needleman's and Rice's finite element
results in the whole range from diffusive cavity growth to hole growth by creep
flow: the cavity growth rate is given by eqs. (11.11) or (11.14) (for Vor R,
respectively) with the only modification that the cavitated area fraction w is
interpreted as
librium-shaped cavity of radius R in a finite circular cylinder of

diameter

w = maximum of { (2R/A)2 and [R/(R+~)]2 }


with

rate.

(15.1)

calculated from eq. (4.17) using the remotely applied stress and
For

large

~,

eq.

(15.1) is identical with the definition of w as the

cavitated area fraction, eq. (11.9), and the growth rate is then the
growth

rate

obtained

in

the

rigid

grain

diffusive

limit. Surprisingly, and this is

fortuitous, the use of eq. (15.1) in eq. (11.11) leads to the correct
in

the

strain

behavior

cr.'eep-controled limit as well. A power series expansion for small

the function q(w) defined in eq. (11.10) with w from eq. (15.1) shows that

of
the

leading term is of order ~3. This means that V - R3 e is obtained. Moreover, the

coefficient is such that the factor A introduced in eq. (14.1) becomes A = 9/8,
which happens to agree with the result of the hole growth analysis given in eq.
(14.19).
For multiaxial loading it cannot be expected that the same accidental
of

q(w)

behavior

leads to the correct growth rates in the plastic hole growth limit as

well. In this case Sham and Needleman (1983) obtain an excellent representation
of their finite element results by adding the cavity growth rates by diffusion
Ceq. (11.11)] and by creep Ceq. (14.17)] in the following way:
(n-1)(n+O.4319)

n2
Here, w has the following meaning:

}~

(15.2)

15.1

217

Creep-Enhanced Diffusive Growth

maximum of { (2R/A)2 and [R/(R+l.51)]2 },


which deviates from eq. (15.1) by the

factor

1.5

(15.3)

multiplying

1,

and

is

and the
calculated from eq. (4.17) using the von Mises equivalent stress, om,
e
corresponding strain rate, ~m = B (om)n, which under axisymmetric loading is
e
e
equal to ~: = ~;. Since the high-triaxiality limit of plastic hole growth is
used in eq. (15.2), its validity is restricted to

0:/0:

corresponds to

0:/0:

> 1. (Uniaxial tension

= 113).

It should be noted that a linear superposition of cavity growth rates calculated

for

the

diffusive

and

the

creep

flow

mechanisms

separately,

with w

interpreted simply as the cavitated area fraction, also reproduces the limiting
cases and is accurate to within a factor of about 2 in the transition range.
From the preceding formulas it is clear that the transition from diffusive
cavity growth to hole growth by creep flow occurs when 1 becomes smaller than
the cavity radius, R. Typical cavity sizes are 2R
the

melting

temperature,

one obtains 1 = R

0.5

~m.

~m

In pure iron at

half

if the stress is 50 MPa.

For higher stresses growth occurs predominantly by creep flow, whereas at lower
stresses diffusion predominates. In creep resistant materials, owing to the relative smallness of ~/o, 1 is usually so large that diffusive growth

predomin-

ates over the whole range of conditions generally applied in creep testing.
Another remark on the coupled growth mechanism seems to be in
the

literature,

order

here.

In

hole growth by creep flow has been combined with non-equilib-

rium, crack-like diffusive cavity growth, which was described in Section

11.2.

The interaction was treated in complete analogy to the case of equilibrium void
shapes, i.e., in eq.
eq.

~11.44)

to (11.47) w was given the

meaning

specified

in

(15.1). Such a procedure, however, appears to be inappropriate in the case

of crack-like growth. The reason that a cavity may grow in a crack-like mode is
the

localized

drain

of atoms out of the cavity into the grain boundary. This

localized drain of atoms happens only in grain-boundary-diffusion controlled


growth, whereas growth by creep flow does not tend to deform the cavity into
crack-like shapes. Hence, the use of eq. (15.1) in eqs. (11.44) to (11.47)
leads

to

totally

wrong behavior in the creep controlled limit. Whether it

makes any sense near the diffusive


R;/l = 0), remains to be examined.

limit,

i.e.

for

small

R/l

(except

for

15.

218

15.1.2

Diffusive Growth with Creep or Elastic Accommodation

Comparison with experiments

Creep-enhanced diffusive cavity growth plays no great role in commercial

creep

resistant materials at their usual operating temperatures. In these cases it is


commonly found that the diffusive length

is greater than the cavity size

and

spacing. Hence, diffusion is the predominant growth mechanism. Often the strain
rate is even so low that diffusive cavitation is constrained, which, in terms
of ~, means that ~3 > A2d, where d is grain size [cf. eq. (12.6)J. Therefore,
there are only a few experiments in which the creep-enhanced
can

growth

mechanism

be expected to dominate. Two examples, both referring to pure metals, will

be descr i bed.
Cane and Greenwood (1975) measured the size of
700 0 C.

cavities

growing

in

iron

For the largest cavity observed in each sample they find that the radius

expands according to R

oc 0 3/2

t 1/2 ,

which in terms of growth rate means


(15.4)

where lengths are in m, time is in sec and stress is in MPa. This


result

at

is

experimental

shown in Fig. 15.1 together with theoretical predictions. Diffusive

growth gives the straight line with slope 1 if the sintering stress is neglected,

which

is

justified

for the large voids considered by Cane and Greenwood

(1975). The absolute values of the diffusive growth rate were

calculated

from

eq. (11.14) using the material parameters for ~-iron at 700 0 C given in Appendix

10-1'

E
.~ 10-12

15

9 10
(J

20

inMPo - -

Fig. 15.1. Growth rate of largest cavity measured by


Cane and Greenwood (1975), compared with various models.

15.1

Creep-Enhanced Diffusive Growth

219

A, and setting w = 0.05. Then eq. (11.14) gives


units as above. In Fig. 15.1 R = 5

R= 5

10-25 a/R2

the

prediction

the

same

was assumed. The diffusive growth rate is

too small by a factor 250 to 1000 compared to the measured values.


is

in

Also

shown

of the plastic hole growth mechanism given in eq. (14.19).

Since the strain rate increases with the 7.3'th power of stress in the material
considered, one obtains a line with slope 7.3. The interactive growth mechanism
is treated in the simplest possible way by adding the growth rates of

the

two

basic mechanisms, which leads to the dashed line in Fig. 15.1.


The third-power of stress dependence of
suggest

non-equilibrium,

crack-like

the

experimental

diffusive

growth

growth
as

rate

might

described

by eq.

(11.46). However, Cane and Greenwood (1975) show that their cavities have equilibrium

shapes.

It

could

be argued that the cavities grew in the crack-like

mode and assumed the equilibrium shape only after the end of

the

creep

test.

However, since surface diffusion preserves the cavity volume, the volume growth
rate would have been measured correctly even if the cavities were

R,

after the creep test. Now, although the linear growth rate,

rounded

out

is higher in the

crack-like mode than it is in the equilibrium mode, the volume growth rate ~ is
smaller.

This

is clear from eq. (11.11) which is valid for both modes. In the

crack-like mode, ao is necessarily larger than in the equilibrium mode, and


hence the growth rate would be even lower than the straight line with slope 1
in Fig. 15.1 indicates. In spite of this conclusion, Miller (1979) claims
agreement

of

Cane

good

and Greenwood's data with the crack-like growth model, but

this is probably incorrect.


A polycrystalline constraint on diffusive growth should
Cane

and

not

be

if

con-

were effective, it would only reduce the calculated growth rate, which

is already too low compared to the experiments. Growth


ledges

in

Greenwood's experiments, since the creep rate, which they report, is

higher than the characteristic rate given in eq. (12.6). Moreover,


straint

effective

rates

of

cavities

at

by grain boundary sliding Ceq. (14.270] are even lower than constrained

growth rates.
In summary, the rates at which the largest cavities in a-iron are
grow

are

This may be due to the fact that the largest cavities tend to
~re

observed

located

at

be

those

junctions

which

stress concentrations. Cane and Greenwood (1975) report grain

boundary sliding in their material. Therefore it is conceivable that at


grain

to

higher than any of the growth mechanisms discussed here can explain.

the

stress

is

concentrated.

Because

triple

of the strong stress

220

15.

Diffusive Growth with Creep or Elastic Accommodation

dependence of the plastic hole growth mechanism a stress

concentration

factor

of 1.4 to 1.8 suffices to raise the calculated growth rates to the level of the
observed ones.
Svensson and Dunlop (1981) evaluate Cane and Greenwood's (1975) data for the
growth rate of the average, rather than the largest, cavity. Since nucleation
was continuous in these experiments, the average size may not be indicative of
the cavity growth kinetics as will be pointed out in Chapter 16. Nevertheless,
Svensson and Dunlop report reasonable agreement of the data with the
growth model.

diffusive

The second set of experiments which refers to the range of creep-enhanced diffusive cavity growth are the tests of Wang, Martinez and Nix (1983) on copper
containing implanted water vapor bubbles. They compare the measured lifetimes
with lifetimes calculated from various models of creep-enhanced cavitation,
e.g. that by Chen and Argon (1981c). They find

that

the

model

exhibits

the

correct dependence of t f on stress and temperature, but the measured lifetimes


are two orders of magnitude smaller. This corresponds to the observation reported above that the cavity growth rates measured by Cane and Greenwood (1975)
exceed the predicted values substantially. As a possible

explanation

for

the

discrepancy, Wang et al offer the effect of grain boundary sliding.

15.2 Diffusive Cavity Growth with Elastic Accomodation


Just as creep deformation, also elastic deformation of the grains allows a nonuniform deposition of atoms on the grain boundary near the growing cavity. However, whereas creep strains can increase indefinitely, elastic strain remains
finite. Hence, only a limited amount of matter can be accommodated elastically.
It will in fact be found that elastic accommodation plays no great role under
typical creep rupture conditions in metallic materials, whereas in ceramics
crack-like growth with elastic accommodation might playa role.
15.2.1

Elasticity effects in the growth of equilibrium-shaped cavities

Consider a cavitating grain boundary between elastic grains. Suddenly, at


t = 0,

remote

tensile

stress is applied. In the first instant, the stress

distribution around the cavity is the elastic field with stress


near

the

void

tip.

In

time

concentrations

the stress gradients along the grain boundary, grain

15.2

Diffusive Growth with Elastic Accommodation

221

boundary diffusion commences as described by eq. (4.4). Diffusion converts

the

initial elastic stress distribution into the diffusion-controlled, steady-state


distribution, which exhibits no stress concentration at the
the

transient,

void

tip.

During

the cavity growth rate decreases from an initially high to the

steady-state diffusive growth rate given by eq. (11.14).


The transient problem was analyzed quantitatively by Raj (1975). He
two-dimensional,

cylindrical

cavities

with

diameter

2R

solved the coupled elasticity/grain-boundary diffusion

considered

and spacing A, and

problem

emplDying

the

Fourier transformation method. His numeric results indicate that the transition
from the initial elastic behavior to the final steady

state

occurs

within

characteristic time of

td

kT A3

0.01

(15.5)

E n &Db

Such a characteristtic time had already been obtained in


relaxation

of

elastic

stress

concentrations

at

connection

triple

with

the

grain junctions or

particles. Comparing this relaxation time with the rupture time for diffusional
cavity

growth, eq. (11.18), shows that elastic accommodation plays a role only

during a small fraction of the life since td/tf

30/E

and

alE

is

typically

less than 1/1000.


A slight deficiency of Raj's analysis is that he keeps the cavity size constant
while

the

stress distribution relaxes to the rigid grain limit. Vitek (1980),

on the other hand, considers an isolated cylindrical cavity growing at


stant rate,

R,

tions. Trinkaus (1978, 1979) considers an isolated axisymmetric cavity


at

a constant volume growth rate

V.

con-

growing

He notes that in this case, if the sinter-

ing stress 00 is neglected, the stress and displacement


the

and solves the problem numerically after making a few approxima-

fields

expand

around

cavity in a self-similar manner in proportion to the cavity radius,


1
t / 3 However, these solutions for the transient problem of a growing void

have

little

practical

relevance,

since

the transient occupies such a small

fraction of the lifetime.


15.2.2

Crack-like cavity growth with elastic accommodation

Vitek (1978) and Chuang (1982) modeled the growth of flat cavities by prescribing

traction-free

boundary

conditions on a mathematically sharp crack rather

222

15.

Diffusive Growth with Creep or Elastic Accommodation

than on the actual cavity surface. This is a valid approximation if the


thickness,

2w,

is

small

cavity

compared to the zone ahead of the void tip in which

substantial diffusive activity takes place [see the length

defined

in

eq.

(15.7) below]. The coupled elasticity/grain-boundary diffusion problem is formulated in terms of an integro-differential equation for the normal stress an on
the

grain

boundary,

which is then solved numerically for steady-state cavity

growth. Here, steady state means that in a coordinate system


cavity

tip,

time

moving

with

the

derivatives can be replaced by alat = - Ra/ax. Far from the

void tip, the stress is required to approach the elastic singular field asymptotically, an

+ KI /(2nx)1/2. This remote condition is justified as long as grain


boundary diffusion is essentially confined to a zone which is small enough

compared

to

the

cavity

diameter 2R (small-scale diffusion), but still large

compared to the cavity thickness as already mentioned. The boundary

conditions

at the void tip for an (denoted by 0 0 at the tip) and for don/dx (denoted by o~
at the tip) are chosen differently by Chuang (1982) and by Vitek (1978), but as
long

as

00

and

we follow the
surface

o~

are not specified the two analyses essentially agree. Here,

~pproach

diffusion

of Chuang who chooses

problem

described

00

and

o~

in accordance

with

the

in Section 11.2. The resulting relation

between the stress intensity factor and the growth rate can be written

in

the

implici t form:
0.75 a

L1/2

where the growth rate is contained in


E

11

0.60 a' L3/ 2 ,

(15.6)

and

00

and in the abbreviation

1/2

cDb

4 (1-v ) kT

0'

(15.7)

] ,

which has the meaning of the diffusion zone size ahead of


linear

dependence of KI on

only numerical result. If

00

00

and

and

o~
o~

the

void

tip.

The

in eq. (15.8) is an exact, rather than an


are

taken

from

the

surface

diffusion

problem as indicated above, eq. (15.6) finally becomes


K = 0.5 Kmin [(R/R . )1112
I
I
mIn

(R/R . )-1/12].
mIn

(15.8)

For convenience the result was written in terms of the quantities


(1

5.9)

15.2

223

Diffusive Growth with Elastic Accommodation

[l

6D 4

8.13 ....:...-.::...
kT y2s

(15.10)

(1-v ) 6Db

which have the meaning of the minimum stress intensity factor and cavity growth
rate, for which eq. (15.8) has solutions. For smaller stress intensity factors,
the steady growth of isolated cavities by elastic accommodation
possible.

Fig.

15.2

shows the relationship between KI and

(15.8). Apparently the slope in a log-log plot varies from 12


intensities to

alone

Aaccording
at

high

is

not

to eq.
stress

near the minimum stress intensity.

As mentioned earlier, it is unlikely that the theory described above is applicable

to

metals.

Lewis and Karunaratne (1981), however, applied the theory to


their experiments on slow crack growth in Sialon ceramics at 1400 oc. Although

the crack length was much larger than what is usually considered as typical for
cavities, the micromechanism of crack growth in one of the ceramics tested

may

have been diffusive growth as assumed in the present theory. The pertinent data
are included in Fig. 15.2, using values for the unknown diffusion
and

surface

energies

which lead to a best fit of the data to the theoretical

curve. The slope of the curve, however, is independent of that fit


with the predicted slope 12.

a6

Near the minimum

t
10-2

lOB

.....,.

'0:

'uQ
~

.
~

.5

:'l::

where

3'

..

agrees

Kr inMPa.m"}_

.!:;

.~

and

stress intensity factor,

00E

coefficients

10-6

10'

....,
..
u

.,

..E
Q

c:

K/K min _

3'

r r

Fig. 15.2. Crack growth rate in Sialon ceramic from Lewis and Karunaratne
(1982) fitted to eq. (15.8).

224

15.

Diffusive Growth with Creep or Elastic Accommodation

the present model ceases to be valid and some other mechanism

is

expected

to

become prevalent, the data start to deviate from the theoretical curve.
The assumptions of small-scale diffusion (L
(2w

Rand L)

are

R) and of a very flat void shape

not made in a numerical investigation by Vitek (1980). He

prescribed traction-free boundary conditions on an elliptical cylinder in order


to

approximate

the actual cavity shape more closely. The numerical result for

the steady-state cavity growth rate can be represented by the formula


( 15.11)

for (O-Oo)/E
Roberts

0.2 w/R, i.e. for relatively thick

(1978)

obtain

essentially

the

same

voids.
result

analytical model. Taking the cavity thickness, 2w, from


stress

at the void tip,

a relation between

Rand

00 ,

Speight,
from
eq.

an

Beere

and

approximate

(11.35)

and

the

from eq. (11.5) with eq. (11.36), then one obtains

stress

which can be resolved for

{( 1 +

1 .75 cSDb

)1/2

E cSDs sin(1/I/2)

R to
1

give:

I6 .

(15.12)

For ocSD b EcSD , eq. (15.12) gives R oc 0 3 while in the opposite limiting case
~
6
there results R oc 0 Correspondingly the rupture time exhibits inverse stress
dependences with exponents between 3 and 6.
Takasugi and Vitek (1981) compare the growth rate predicted by eq. (15.12) with
the

growth

rate obtained for rigid grains. They observe that under almost all

practically interesting circumstances, elastic accommodation does not


ate

of cSDs/cSD b , does elastic accommodation predominate. For example, in


of

acceler-

cavity growth markedly. Only at very high stress or extremely small ratios
Goods

and

the

tests

Nix (1978) on water vapor bubbles in silver, which were already

discussed in Section 11.2.7, elasticity effects might make a small contribution


R oc 0 3 . 7 , can be appro-

to the growth rate. The experimental stress dependence,

ximated by eq. (15.12) more closely than by the rigid grain model,
an

exponent

of

which

gave

3. However, the best fit of eq. (15.12) with the experimental


-5
0
data requires that cSDs/cSD b = 310
and 1/1 = 8 Both are unusually small values
which can only be accepted if it is admitted that the water vapor in the cavities affects the surface properties of silver strongly.

16 The Cavity Size Distribution Function for


Continuous Cavity Nucleation.
Rupture Lifetimes and Density Changes

As was pOinted out in Section 5.7, cavities usually nucleate continuously


substantial

fractions

of

the

creep

rupture

over

life of metals and engineering

alloys. Attempts to calculate the rupture lifetime must

take

this

fact

into

account.

16.1 The Cavity Size Distribution Function


A link between experimental data on one hand and theories on cavity

nucleation

and growth on the other is the cavity size distribution function. It is denoted
by N(R,t), where NdR is the number of cavities per
having

radii

unit

grain

boundary

between Rand R+dR. By agreement, we refer the cavity density to

those boundary facets only which are 'essential' for the rupture
failure

process.

For

by diffusive cavity growth, for example, it is convenient to refer the

density to boundaries. which have orientations


principal

area

stress

axis,

since

between

60 0

and

90 0

to

the

these boundaries cavitate preferentially. This

freedom in definition does not affect the results.


Cavities can pass from one size class, R, to the next,
growth,

R+dR,

only

by

cavity

if cavity coalescence during the late stages of the rupture process is

ignored. Then the distribution function must obey the continuity

condition

in

size space (Riedel, 1985c),


( 16.1)
where the superposed dot denotes the time derivative and R(R,t) is
rate

of

cavities

having

the

growth

a radius R at time t. The growth rate will be taken

from any of the models in the preceding chapters.

Its

time

dependence

stems

from its dependence on the cavity spacing, which decreases continuously as more
cavities are nucleated. Cavity nucleation enters into the problem in

the

form

16.

226

of

boundary

condition

to

Continuous Nucleation

eq. (16.1): the flux in size space, NR, at some

small cavity radius, which will be set equal

to

zero,

must

be

equal

to

prescribed nucleation rate,


NR

= J*

at R

O.

(16.2]

The nucleation rate is taken either from nucleation

theories

or,

as

in

th.e

following, from the observed nucleation kinetics.


Equations (16.1) and (16.2) have steady-state solutions,
the

nucleation

&= 0,

provided

that

rate is constant and R is not time dependent. In this case the

distribution function is directly given by N = J*/R. However, the prerequisites


for

the steady-state solution are usually not satisfied. Correspondingly, Chen

and Argon (1981b) remarked that the steady-state solutions are


with

the

size

distributions

observed

by

Cane

not

compatible

and Greenwood (1975) and by

Needham and Gladman (1980).


A more general class of solutions can be obtained if the cavity growth rate and
the nucleation rate have the the power-law forms
(16.3)
(16.4)

J*

where A1 , A2 , a, Band Y may possibly depend on stress and strain rate, but not
on time nor on the cavity size. Because of the power-law forms of the
equations,

one

solutions,

which one seeks in the form N(R,t) - tPf(R/t Q). The exponents P and

may

expect

that

eqs.

(16.1)

Q and the functioh f are obtained by inserting


eqs.

(16.1)

and

(16.2).

The

and
the

(16.2)

have

similarity

similarity

solution

into

resulting equation for f indeed depends on the

co-ordinate R/tQ only, but not on Rand t separately,

if

Q = (a-1)/(B+1)

and

P = BQ+a+Y. The ordinary differential equation for f(R/t Q) can be solved by


separation of the variables. The final result is:
1-a

RB+1

------]

(a+Y)/(l-a)

1+B A t 1-a

(16.5)

This solution is shown in Fig. 16.1. If a < 1, the distribution


be cut

off at a maximum R,

which is obtained by setting

function

must

the term in brackets

16.1

227

The Cavity Size Distribution Function

ex. = -5/4
~ = 2
1=0

c:n

ex. = 2

ex. = 0
~ = 2
1=1

30

P= 2

1 =0

10
2:

<D

t=1

C"")

10
R-

R-

Fig. 16.1. Evolution of the cavity size distribution function for A1

1.

equal to zero. If a > i, the distribution extends to infinite

the

first

cavities

nucleated

R,

because

infinitely fast if a > 1 is assumed. The case

grow

a = 1 is a degenerate one. Recalling the formula (1

= e for x ~

+ 1/x)x

gives

00

(16.6)
for a = 1. It is worth noting that in this last example,
size

distribution,

the

maximum

of

the

as well as the average size, remain at a fixed value of R,

while in the general case represented by

eq.

(16.5),

the

maximum,

and

the

average, can move to larger or to smaller sizes depending on the exponents a, S


and Y. Hence no conclusions regarding

cavity

growth

kinetics

can

be

drawn

directly from the observation of the evolution of the average cavity size.

16.2 The Cavitated Area Fraction and the Rupture Lifetime


The cavitated area fraction,

w,

of the

grain

boundaries

having

orientations

and
to the tensile axis is obtained by integrating the areas,
~R2, occupied by the individual cavities times their density, NdR:

between

60 0

90 0

f ~R2 N(R,t) dR
I(

a,,,,
0

y)

(16.7)

A A2/ (S+1) t a +Y+(1-a)(S+3)/(S+1)


2

where the second line was calculated with

N(R,t)

from

eq.

(16.5),

and

the

16.

2~

Continuous Nucleation

dimensionless factor I(a,B,Y) is the definite integral


w (1+B)(B+3)/(B+l) ~ xB+2 [1 - (l-a) x B+1 ](a+Y)/(1-a)dx

with U

(16.8)

= ~ if a> 1 and U = (l_a)-l/(B+l)

expressed

if

< 1. This integral can be

by the Beta function (Abramowitz and Stegun, 1968, pg. 258). Here it

suffices to remember that I is a number which is independent of A1 , A2 and t.


Rupture is assumed to occur when the area coverage- attains
denoted

by

chosen as wf

W=

wf

The

critical

value,

numerical value of wf is somewhat arbitrary, and will be

= w/4, since regularly spaced round cavities touch each other if

w/4. With the failure criterion,

W=

wf ' inserted, eq. (16.7) can easily be

resolved for t, which gives the rupture lifetime, t f


considered next.
16.2.1

Special

cases

will

be

Lifetimes for diffusive cavity growth and continuous nucleation

unconstrained diffusive cavity growth rates are given by eq. (11.14). Unfortunately

neither the sintering stress nor the function q(w), which was defined in

eq. (11.10), are exactly compatible with the

power-law

form

of

eq.

(16.3),

which is required for the similarity solutions to be valid. In order to be able


to apply the similarity solutions, we neglect the sintering stress and approximate

q(w)

by

some power funtion of w. The simplest choice is q(w)

approximates q(w) in an average, though admittedly


With

this

choice,

eq.

(11.14)

not

well

eq.
=

(5.1),

the

nucleation

1, which
sense.

assumes a power-law form compatible with eq.

(16.3) with a = 0 and B = 2. If, in agreement with the observations


in

defined,

rate

summarized

is assumed to be time-independent, i.e.

0, then the rupture lifetime follows to be


0.33 ( h(~)kT )2/5
06D b O

(16.9)

~ 1/0(3n +2)/5.

The second line shows the stress dependence of the lifetime predicted for the
case in which the nucleation rate is given by J* = a'E = a'Bo n according to eq.
(5.1) and to Norton's creep law. The activation energy is then predicted to

be

(2Qb+3Q)/5. where Qb and Q are the activation energies for grain boundary
diffusion and for power law creep. respectively. The result will be compared

16.2

229

Cavitated Area Fraction and Rupture Lifetimes

with experiments in Section 16.3.


Instead of approximating q(w) by unity, q(w) can be replaced by a power

funct-

ion, which is always greater than q(w). This leads to a lower-bound estimate of
t f In the rigid-grain limit, which is considered here, the volume growth rate
rather than on
of a cavity depends on the average cavitated area fraction,

w,

w = (2R/A)2 calculated with the individual cavity radius. This means that

should

be

approximated by a power function of

w.

q(w)

If, for example, one chooses

q(w) = 0.64/w1/2 , the requirement that the resulting

wmust

be consistent

with

(16.7) leads to a = -(3Y+5)/4. Then t f is obtained in a form like eq.


(16.9), but with a numerical factor 0.25 instead of 0.33 and with w~/5 instead

eq.

w~/5. Apparently, the result is relatively insensitive to the approximation

of

made on q(w).
This insensitivity of the result to changes in the cavity growth
why

Lonsdale

and

Flewitt

(1979,

1981)

law

explains

obtain a result very similar to eq.

(16.9), although they start from an incorrect cavity growth rate. Also
and

Gladman

(1980)

do

not

use

Needham

a correct expression for the growth rate of

axisymmetric cavities. Further, they assume that a critical volume fraction


cavities,

rather

Consequently,

than

they

critical

arrive

at

slightly

area

fraction,

different

leads

stress

to

and

of

fracture.
temperature

dependences of the rupture time and they obtain a grain size dependence.
In some cases, the cavity spacing is found to decrease in inverse proportion to
the

creep

strain,

A = 1 1(1ii"e:) , where a" is an empirical factor having the


physical dimension 11m2 This means that the nucleation rate has the form
2 a"

J*

~2 t,

(16.10)

which is a power law in time compatible with eq. (16.4)


with

the

diffusive

growth

rate,

with

and approximating q(w) by

Y = 1.

Together

as before, eq.

(16.7) then leads to a lifetime:


t

0.56 ( kT h(ljI) )1/4 (


QIiDb o

w~2 )3/8.

(16.11)

aile:

Together
tf

with Norton's law,


this
predicts
a
stress
dependence
(3n+1)/4
.
1/0
A comparison with experiments follows in Section 16.3.3.

Instead of employing an empirical relation for the

nucleation

rate,

of

Trinkaus

16.

230

and

Ullmaier

(1979)

take

Continuous Nucleation

the theoretical expression for J* derived from the

theory of thermally activated vacancy condensation.

They

use

eq.

(6.28)

to

calculate the nucleation rate, with c neglected against cmax . This assumption
requires that the possible nucleation sites are not exhausted during the rupture

lifetime,

and

it

gives

a constant nucleation rate, i.e. Y

cavity growth rate they use a result for the diffusive growth
cavity

with

elastic

accommodation.

exactly time-independent, so that R R


fact

that

In

-2

of

O. For the
an

isolated

this case, the volume growth rate is


,i.e. a = 0 and B = 2.

Ignoring

the

the validity of that solution is restricted to very short times, as

was pointed out in Section 15.2.1, one obtains the rupture time in the form
(16.12)
where NL is the Loschmidt number andt fo is the following abbreviation:

Here, the spacing of the potential nucleation sites, A ,is related to their
2
nuc
area density, c max ' by 1/Anuc = cmax For high stresses the first term in the
exponential of eq. (16.12) is small. Then the rupture time behaves similarly as
in

eq.

(11.18)

where

nucleation was assumed to occur instantaneously at the

beginning of the test, i.e., t f 1/(a6Db ). When the stress becomes smaller,
the first term in the exponential function starts to dominate. That means that
nucleation becomes difficult,
decreasing

and

the

rupture

time

increases

sharply

for

stress. At the same time, the effective activation energy for creep

rupture becomes strongly stress-dependent


(16.111)
For the effective stress exponent appearing in t f

1/am, eq. (16.12) predicts

m = -3(lnt f )/3(lna) = 1 + 24 y3 f (111)/(5


s v

kT).

(16.15)

For high stresses, m approaches unity, while it rises sharply for

stresses

the

Under

order

of,

and

smaller

than,

the

nucleation

stress.

of

which

circumstances such a stress dependence is observed experimentally remains to be


examined.

Certainly

the

predictions

compatible with the Monkman-Grant rule.

of

eqs.

(16.12)

to

(16.15)

are not

16.2

Cavitated Area Fraction and Rupture Lifetimes

Raj and Ashby (1975) analyze a problem

of

231

continuous

cavity

nucleation

and

diffusive growth similar to that considered by Trinkaus and Ullmaier (1979). In


contrast to the latter authors, Raj and Ashby use an expression for the
sive

growth

rate

of

diffu-

cavities having a finite, as distinct from an infinite,

inter-cavity spacing. For the nucleation rate by vacancy condensation they


ploy

eq.

em-

(6.17), which neglects the Zeldovich factor, but they admit that the

potential nucleation sites may be exhausted, i.e. they do not

neglect

com-

pared to cmax . Otherwise their model, which they treat numerically, is the same
as that of Trinkaus and Ullmaier (1979). They find that for stresses which
exceed the cavity nucleation stress markedly, the rupture lifetime exhibits the
inverse stress dependence of eq. (11.18), while at

lower

stresses

nucleation

becomes difficult and the lifetime increases drastically.


16.2.2

Crack-like diffusive growth and continuous nucleation

The general result for the growth rate in the


cavity

growth

was

given

in

limit

of

crack-like

diffusive

eq. (11.44). Again, the growth rate has not the

power-law form assumed in eq. (16.3), which allows for similarity solutions for
the

cavity

size

diffusion where

R~

distribution function. However, in the limit of slow surface


0 3 , eq. (11.46), the growth rate is independent of Rand t

if the dependence on (1-w) is neglected, i.e. a

a=

O. For a time-independent

nucleation rate, the rupture time is found by solving


setting

W=

eq.

(16.7)

for

and

wf :
(16.16)

If J* ~

= Bon, the rupture time exhibits a stress dependence t f ~ 1/o2+n/3.

16.2.3

Constrained diffusive growth and continuous nucleation

The analysis of constrained cavitation in Chapter 12 has shown that constrained


cavitation

is

synonymous

Therefore it is

unlik~ly

continues

boundaries

on

with

small

stresses on the cavitating boundaries.

that cavity nucleation, which requires high


that

undergo

constrained

cavitation.

nucleation then means that new boundaries, which were not

previously

stresses,
Continuous
damaged,

develop cavities. The continuous nucleation of practically traction-free facets


was considered

in

Section

12.7.2

and

12.7.3,

and

that

is

probably

the

appropriate way of dealing with continuous nucleation in the constrained-growth


regime. Nevertheless, the less convincing approach,

which

assumes

continuous

Continuous Nucleation

16.

232

cavity nucleation on already cavitated boundaries, is briefly described next.


In the limit of creep-constrained cavity growth, the growth rate
eq.

is

given

by

(12.5). If the sintering stress is neglected, the growth rate has the pow-

er-law form,

(A/R)2. The cavity spacing, A, is determined by the number

of

cavities already nucleated,

1//

(16.17)

f J*dt

so that B = 2 and a = Y+1. For time-independent nucleation, Y = 0,

eq.

(16.7)

gives the time to cavity coalescence on isolated boundary facets, tc

where r(2/3)
the time to

3n(1+3/n)
c

J*

1/3
)

a; hew)

(a~

e; d)

2/3

wf
2r(2/3)

( 16.18)

1.354. It should be emphasized that tc cannot be expected to

r~ture

be

if cavity growth is constrained. An alternative form of the

result is obtained if the cavity spacing

at

coalescence,

introduced instead of the nucleation rate:

Ac = (J*t c )-1/2

is

(16.19)

wf = n/4 and uniaxial tension.


Comparison with the second term of eq. (12.11) shows that continuous nucleation
The second line is for

n = 5,

heW) = 0.61,

compared

to

instantaneous

1.7 w~/2

wf = n/4)
nucleation of all cavities at the beginning of the

increases the time to coalescence by a factor


test. Therefore, it is almost irrelevant

n~~erically

(=

1.2

for

whether cavity

nucleation

on boundaries undergoing constrained cavitation is continuous or instantaneous.


Another alternative to present the time to coalescence is
nucleation rate by J* = a'E. Equation (16.18) then becomes

to

express

the

(16.20)
with n = 5, hew) = 0.61 and a:la~ = 1 as above. The dependence on
nucleation rate, a', resembles the empirical relation given in eq. (5.2).

the

16.2

233

Cavitated Area Fraction and Rupture Lifetimes

For later use we note that the whole problem, starting at eq.
formulated

(16.1),

sis parallels that in terms of time, but now no assumption as to the


of

the

can

be

using strain instead of time as an independent variable. The analy-

strain

constancy

rate is needed. The results remain the same if in eqs. (16.19)

and (16.20) the product Etc is replaced by the time-integral on E from 0 to tc.

J* = 2a"E 2t as in eq. (16.10), and if


cavity growth is constrained, the time to cavity coalescence is
If the cavity nucleation rate is given by

(16.21)
16.2.4

Inhibited cavity growth and continuous nucleation

Inhibited cavity growth rates as treated in Chapter 13 exhibit the same dependencies on cavity spacing and cavity size as constrained growth rates. Therefore
the results of the preceding subsection can be transferred to inhibited growth.
In particular, the conclusion remains true that continuous nucleation extends
the time to coalescence by some 20% compared to instantaneous nucleation. Thus
the

time to coalescence, which can be identified with the rupture lifetime for

inhibited cavity growth, is given by eq. (13.7) multiplied by 1.2.


16.2.5

Plastic hoie growth and continuous nucleation

The cavity growth rate by creep flow of the matrix is given by eq. (14.19),

if

R = [0.37/h(1jI)]RE, I.e. a

0,

B = -1. Unfortunately, the case B = -1 is a degenerate case in eq. (16.5),

but

only
for

uniaxial

tension

is

considered. Then is

time-independent J*, a steady-state solution for the cavity size distribu-

tion function is possible, N = J*/R. Then eq. (16.7) can be integrated from

Ro

(the

size at which cavities are nucleated) to Rmax ' which follows by integrating the cavity growth law to be R
R exp[0.37E/h(1jI)]. Resolving for strain
max
0
and setting; = wf leads to the strain to failure
(16.22)
where h(1jI)
of

0.61 was inserted. If J*

E, the strain to failure is independent

stress, but eq. (16.22) usually predicts an absolute value of several hund-

red per cent. This is too large to be applicable to creep rupture conditions.

234

16.

Continuous Nucleation

16.3 Comparison of Calculated Rupture Tunes with Experiments


Involving Continuous Nucleation
16.3.1

Rupture lifetimes of ferritic steels

Needham (1983) reports rupture lifetimes of several ferritic steels along


measured

cavity

with

nucleation rates and creep strain rates, so that the data can

be compared unambiguously

with

subsection.

example from that investigation, Fig. 16.2 shows the

As

first

the

lifetimes

calculated

in

the

preceding

rupture lifetime of 2 /.Cr-1Mo steel austenized at 950 0 C having a grain size of


1

18

~m

(steel No.1 in Needham's notation). Measured cavity growth rates in that

material were given in Fig. 12.2, which also showed that the constrained growth
model

agrees

with

the

observed

growth

rates.

The solid line in Fig. 16.2

represents the time to cavity coalescence, tc' calculated for constrained


growth from eq. (16.18), whereas the dashed line is the time to rupture for
unconstrained growth as calculated from eq. (16.9). Numerical values for J* and
~ were taken from Needham's report, Q6Db is given in Appendix A, and for wf and
the following plausible values were chosen: wf = ~/4, h(~) = 0.61. In the

h(~)

stress

range

considered, the constrained model leads to greater rupture times

than the free diffusive model. Since the slower of the two processes determines
the rate, the constrained model is the appropriate one here .

constrained

10'

uncon- ,
strained "

100

"-

150
(5

, ......
2(1J 250

in MPa --

Fig. 16.2. Rupture lifetime of 2 1 /.Cr-1Mo steel. Solid line: eq. (16.18).
Dashed line: eq. (16.9). (From .Riedel, 1985c).

16.3

Comparison with Experiments

235

Interestingly, the constrained growth model


to

factor

1.5,

although

the

predicts

the

data

to

within

time to cavity coalescence is not obviously

related to the time to rupture, if cavity growth is constrained as was pointed


out in Section 12.3. Apparently, the joining of the cavitating facets, which is
rate controlling in the constrained-growth regime, obeys similar

kinetic

laws

as constrained cavity growth. The curve for unconstrained growth underestimates


the lifetime, but by not more than by a factor 4. This shows that unconstrained
diffusive

growth,

in

conjunction with continuous nucleation, may lead to the

correct order of magnitude for t f , although the cavity growth rate is sUbstantially overestimated as was shown in Fig. 12.2.
Figure 16.3 shows the same type of comparison for a 1Cr-1/2Mo steel in two
differently heat-treated conditions. The data were taken from Needham's (1983)
report. Cavity growth rates for the material austenized at the ordinary austenizing
the

temperature
circular

experimental

of

symbols
data,

930 0 C were given previously in Fig. 12.2. In Fig. 16.3,


refer

while

to

the

that
open

material:
symbols

closed symbols

bound

the

represent

ranges in which the

theoretical formulas were evaluated. Square symbols represent data for the
mater.ial austenized at 1300 oC. The high austenizing temperature was chosen in
order to simulate the conditions in the neat-affected

zone

of

weld.

This

special heat treatment leads to the precipitation of finely dispersed MnS-part-

D. Gust. 1300e

10'

t
<:
.5;

Gust. 930e

0....
lQ1

0.....

....

1~

,
"-

, ......
"- . . . . I ....... ,
........
"II ........

"'-

"-

. "

550e
"'""'1er-'izMo steel

10
100

150
(j

in MPa

'"0

--

200

250 300

Fig. 16.3. Rupture lifetimes of two heats of 1Cr- 1/ 2Mo steel.


Solid lines: constrained growth, eq. (16.18). Dashed lines:
unconstrained growth, eq. (16.9). The open symbols bound the
ranges in which the equations were evaluated. (From Riedel, 1985c).

236

16.

Continuous Nucleation

icles on prior austenite grain boundaries. Since the sulfides nucleate cavities
easily,

the nucleation rate increases, and therefore the rupture lifetimes are

reduced. In addition, the high austenizing temperature increases the grain size
88 ~ compared to 18 ~m in the material austenized at 930 oC, and the coarse

to

grain size affects the time to cavity coalescence


constrained

the data is poor in this example.


predictions

adversely

in

the

case

of

growth. Apparently, the agreement of the calculated lifetimes with


of

the

The

data

lie

about

halfway

between

the

constrained and the unconstrained models, and differ from

both by factors between 2 and 10.


Figure 16.4 shows rupture
(1979).

lifetimes

of

steel

21/~Cr-1Mo

measured

by

Cane

Measured cavity growth rates from that investigation were shown previ-

ously in Fig. 12.3. The material exhibits relatively poor creep rupture properties since it was given the heat treatment involving austenization at 1300 oC,
which results in an unfavorable distribution of sulfides. Figure 16.4 illustrates

the

effect of continuous as compared to instantaneous nucleation. For un-

constrained growth, the rupture lifetime is substantially


are formed continuously than if all cavities

~ere

longer

if

cavities

present from the start of the

test. This is demonstrated by the two dashed lines in Fig. 16.4, which are calculated from eqs. (11.18) and (16.9), respectively.

10CX1J

--!\............
. . . . . -..

lCX1J

.
contmuous
nucleation

unconstrain~" "-

1 j
100

(The curve for instantane-

'\~

instantaneous "
_____ ::~~~n~~
2'4 Cr-1Mo steel 565-C

10~--------~----~----~

50

100

150

200

oinMPa __
Fig. 16.4. Rupture lifetime of overheated 21/~Cr-1Mo steel after Cane (1979).
The theoretical curves correspond to eqs. (11.18), (16.8), (12.11) and (16.18)
(from bottom to top). (From Riedel, 1985c).

16.3

Comparison with Experiments

237

ous nucleation and unconstrained growth


spacing,

A,

increases

is

slightly

from approximately 4.5

bent

to 5.2

since

the

cavity

at higher stresses

according to Cane's observations). On the other hand, if cavity growth is

con-

for the"time to coalescence whether all


cavities are nucleated at the beginning of the test or continuously throughout
the test. This is shown by the solid lines, which represent eqs. (12.11) and
(16.18), respectively. In this case, the agreement between the calculated time
to coalescence by constrained growth and the measured rupture lifetime is
excellent considering the fact that no adjustable parameters are used.

strained,

it

is

almost

irrelevant

The effect of the sintering stress on the rupture lifetime, which


in

was

studied

Section 11.1.3, was examined in relation to the data of Cane (1979). It was

found from eq. (11.22) that regarding the sintering stress

raises

the

bottom

Needham and Gladman (1980) tested Type 347 stainless steel at 550 0 C and

650 0 C.

curve in Fig. 16.4 by at most 5%.


16.3.2

Lifetimes of austenitic steels

Measured

cavity

growth

growth rates in Fig.


nucleate

rates

12.4.

continuously

at

In
a

in this material were compared with calculated


these
constant

experiments,

cavities

were

found to
11
rate J* = a'E with a' = 8.10 /m2 . The

or
growth were calculated from eqs. (16.9) and (16.18), respective-

rupture lifetimes resulting from this nucleation rate and from


unconstrained

lY. The results are shown in Fig. 16.5 together with

constrained

Needham's

and

Gladman's

data.

The agreement of the constrained growth model with the data is good, but
to a lesser extent at 650 0 C than at 550 oC. Thus again one observes that the
time

to

coalescence by constrained growth is about equal to the time to rupt-

ure, although there is no obvious reason that the two quantities should be
rectly

related.

with continuous nucleation, also gives the


qualitatively

di-

Interestingly, the unconstrained growth model, in combination

correct

dependence

on

right

order

of

magnitude

and

stress. Obviously, if nucleation occurs

continuously, the rupture lifetime is sometimes relatively insensitive to the


cavity growth mechanism. Conversely, this example illustrates the difficulty of
investigating growth kinetics by lifetime measurements.
figure 16.6 shows rupture lifetimes of
specimens

tested

Type

316

stainless

steel

bi-crystal

by Gandhi and Raj (1982). The diffusive growth model in com-

bination with a constant number of cavities leads to a lifetime, given


(11.18),

which is

represented by the dashed line.

in

eq.

This theory underestimates

16.

238

10000

347 stainlessstee/

'\

, 1000

--<:

.s;:

10 6

10 5

\I)

'\

10'

'\

100

10J
50

"-

200

150
(fin

977K
bicrystals
316 stainless st.
Gandhi and Raj

-- ---

---

300

MPa

100

(j

400

in MPa

the data by more than two orders of magnitude. In bi-crystal


be

due

to

---

200

Fig. 16.6. Rupture lifetime of Type


316 stainless steel bicrystals
(Gandhi and Raj, 1982). Solid line:
eq. (16.9). Dashed line: eq. (11.18).
(From Riedel, 1985c).

Fig. 16.5. Rupture lifetime of Type


347 stainless steel (Needham and
Gladman, 1980). Solid lines: eq.
(16.18). Dashed lines: eq. (16.9).
(From Riedel, 1985c).

cannot

A=2!J.m

unconstr.~:
100

Continuous Nucleation

experiments

this

polycrystalline constraint, but continuous, rather than

instantaneous, nucleation can increase the rupture time. The solid line in Fig.
16.6

is calculated for diffusive growth and continuous nucleation as described

by eq. (16.9). Unfortunately, Gandhi and Raj do not provide all data
to

evaluate

necessary

eq. (16.9). Therefore the nucleation rate was assumed to obey the

relation J* = a'E; a' was chosen to fit the solid line to the data as
12 2
a' = 310 1m, which is slightly higher than the value for a' observed by
Needham and Gladman
B = 3'10-25 MPa- 8 1 Is

(1980).
by

The

creep

extrapolation

coefficient
of

was

assumed

to

be

Needham and Gladman's data to the

testing temperature of 977 K. Owing to these uncertainties the

absolute

value

of the theoretical rupture time should not be taken seriously, but the slope is
predicted as -(3n+2)/5 = - 5.26 independent of these uncertainties, which is in
reasonable agreement with the data.
The prolonged lifetimes of the bi-crystal specimens compared to
cavity

the

diffusive

growth model could possibly be a consequence of an inhibition of cavity

growth as described in Chapter 13. Inhibited growth models, however, predict


slope

of

- n = - 8.1

observed slope.

in

Fig.

16.6, which is significantly steeper than the

16.3

Comparison with Experiments

16.3.3

239

Rupture lifetimes of astroloy

Shiozawa and Weertman (1983) performed creep rupture tests on the nickel-base
superalloy 'astroloy' at 750 oC. By giving the material appropriate prestraining
and annealing treatments, they were able to generate cavities preferentially on
grain

boundaries

having a special orientation. The pre-cavitated material was

then creep tested with the cavitated boundaries


transverse

oriented

times. Material which was pre-cavitated on grain


direction

either

parallel

or

to the applied stress. Figure 16.7 shows the measured rupture lifeboundaries

parallel

to

the

of the creep strain behaves similarly to material which was not pre-

strained at

all,

whereas,

not

surprisingly,

pre-cavitation

on

transverse

boundaries leads to much shorter lifetimes.


Figure 16.7 also shows calculated lifetimes. The bottom two lines relate to the
material pre-cavitated on transverse boundaries. Equations (11.18) and (12.11),
which are valid for instantaneous cavity nucleation,
fixed

cavity

spacing

of

A= 2

~m,

were

a grain size of d

evaluated
90

~m,

using

the diffusion

coefficient, oD b , reported in Section 12.4.1 for Nimonic 80A, and heW) = 0.61.
Strain rates have been given by Shiozawa and Weertman (1983) as E = 1.4.10-6 /s
at a = 500 MPa and E = 3.6.10-6 /s at 650 MPa. With these values, the time to
cavity coalescence

predicted by the constrained

growth

model,

eq. (12.11),

astroloy

1000

750C

1(Xl

250

300

400
(J

500

650

inMPo -

Fig. 16.7. Rupture lifetimes of astroloy for virgin material (triangle) and
for material pre-cavitated on boundaries parallel to stress axis (open
squares) and transverse to stress axis (solid squares), from Shiozawa
and Weertman (1983). Theoretical curves for constrained and unconstrained growth and for continuous and inst'antaneous nucleation.

16.

240

approximately agrees with the solid squares, while

Continuous Nucleation

unconstrained

growth,

eq.

(11.18), underestimates the lifetime.


In material which was not prestrained, the cavity spacing was found by metallographic inspection to decrease inversely proportional to strain as described by
eq. (16.10) with the empirical factor ex" having the value
rupture

lifetime

ex" = 3'1 012/m 2

The

is then given by eq. (16.11) for unconstrained growth and by

(16.21) for completely constrained growth. In addition to the numerical

values

given above, wf = ~/4 was used in preparing the top two curves in Fig. 16.7.
Interestingly, unconstrained growth leads to longer lifetimes than does
completely constrained growth. Among these two growth mechanisms the slower one
should determine the lifetime, i.e. the constraint should not be
the

cavitation

of

the

effective

in

virgin material, and eq. (16.11) should be valid. The

agreement of eq. (16.11) with the data represented by the triangle and the open
squares

in

Fig. 16.7 is satisfactory. Final conclusions cannot be drawn since

the stress ranges in which rupture lifetimes have been measured


the equation can be evaluated do not overlap, because data for

and

in

which

are lacking at

smaller stresses.

16.4 Density Changes During Cavitation


Density measurements have often
investigate

its

kinetics.

From

been
the

applied
cavity

to

detect

cavitation

and

change in specific volume can be calculated as an integral over the volumes


the

individual

cavities.

For

cube-shaped

to

size distribution function, the


grains

of

with cube axis d and with

cavitation on all boundaries normal to the applied stress the

relative

volume

change by cavitation is:


l1V/V = (lid) J N(R,t) (4~/3)h(IjJ)R3 dR,

(16.23)

where an equilibrium cavity shape was assumed. With N(R,t) from eq. (16.5)
intgral can be evaluated and leads to:
l1V IV

[4~h(IjJ)/3dJ A A3/ (B+1) t Y+1 +3(1-ex)/(1+6) IV I (6+1),


2

where the dimensionless factor Iv is given by

the

(16.24)

16.4

Density Changes

241

for Cl '" 1
(16.25)
for Cl

The second argument of the Beta function, w, depends on the sign of 1-Cl, viz.,
for Cl < 1

w = (1 +y) I ( 1-Cl)

(16.26)
w

(1+Y)/(Cl-1) -

for Cl ) 1.

3/(6+1)

The numerical value of IV is of no interest here. Rather the dependencies on


time and on the nucleation and growth kinetics will be discussed, which are
contained in A2 and A1 , respectively. As an example, we consider a constant
nucleation rate, J* = A2 = Cl'E, together with diffusive cavity growth. With the
approximations made in Section 16.2.1, the diffusive cavity growth
the

rate

takes

power-law form of eq. (16.3) with Cl = 0 and 6 = 2. Then, from eq. (16.24),

the volume change becomes


AVIV

(16.27)
2

A linear dependence of the volume change on Et


but

usually

= Et has often

been

observed,

with an additional stress dependence, as in the following example

on constrained growth.
In the limiting case of creep-constrained cavity growth a
calculate

direct

approach

to

the volume change is simpler. The volume growth rate of a cavitating

grain facet was given in eq. (12.2). If Nmc is the number of cavitating facets
per unit volume (the subscript means 'microcracks', since facets undergoing
constrained
increases

cavitation

behave

mechanically like microcracks) and if Nmc


in proportion to strain, Nmc d3 = 6E, as assumed in eq. (12.33), then

the total volume change by the cavitating facets is given by


AVIV

(16.28)

This can be written as AVIV ~ Eta n exp(-Q/RT), which describes data by


and

Gladman

(1980)

for

Type

energy for power-law creep).

347

Needham

stainless steel well (Q is the activation

17 Summary of Results on Cavity Nucleation


and Growth

Grain-boundary cavitation is the dominant failure mode of many


and

ceramic

metals,

alloys

materials at temperatures between about a third and two thirds of

the melting temperature and at low


associated

with

relatively

applied

low

stresses.

Cavitation

the

~ucti1ity,

is

fracture

usually

path

being

intergranular. The regime of cavitation failure is bounded at high stresses

by

transgranular dimpled fracture or, in some materials, by transgranular cleavage


fracture.

At

high

recrystallization

temperatures,
prevent

the

grain

cavitation

boundary
of

migration

grain

and

dynamic

boundaries. At very low

stresses and lew temperatures, cavitation certainly proceeds

very

slowly

and

possibly stops when no stable cavities can be nucleated.

17.1 Nucleation
If a critical stress for cavity nucleation
10 MPa

in

pure

metals

exists,

it

certainly

these stresses cavity formation Is still observed to occur. The


cavity

nucleation

sites

lies

below

and well below 100 MPa in engineering alloys since at


in

metals

most

frequent

and alloys are grain boundary particles,

triple grain junctions and ledges in grain boundaries. Another important observation is that cavity nucleation occurs continuously over substantial fractions
of the creep life, the number of cavities being approximately

proportional

to

creep strain.
The theories of cavity nucleation by the rupturing of atomic bonds or by vacancy

condensation

described in Chapter 6 are not directly compatible with these

observations. Cavity nucleation by vacancy condensation would be


occur

at

stresses

above

typically

5,360 f~/2 MPa,

function fv characterizes the shape of the cavity

predicted

to

where the dimensionless

nucleus.

(For

spherical

nucleus Is fv = 4w/3). It is unlikely that fv has values very much less than
unity in all materials which cavitate. Therefore, there is a gap between the

17.1

243

Nucleation

observed and the calculated nucleation stress.


In order to bridge the gap, local stress concentrations
grain

junctions

at

particles,

triple

and grain boundary ledges were considered in Chapter 7. Grain

boundary sliding plays a role by contributing to the stress concentrations, but


it

will usually not be the overriding factor for the cavity nucleation process

in engineering materials. It was found that the local stress


grain

are oriented favorably for sliding. On boundaries


stress,

there

is

less

correspondingly smaller.
preferentially.

(The

or

no

sliding

Nevertheless,

apparent

and

transverse

the

stress

by

transverse

localized

slip

bands

such

boundaries

as

the

nickel

base

which

cavity

nucleation.

the

applied

cavitate

If

cavity

tend

to

superalloys,

concentrations at the intersection of slip bands and grain


to

to

concentrations are
even

preference may be due to cavity growth, but at

any event, nucleation must occur first). In materials

lead

concentration

boundary sliding can hardly exceed a factor 10 or 20 on boundaries which

develop

the

boundary

stress

particles

nucleation is triggered by slip band

formation, it is conceivable that nucleation occurs

continuously,

since

slip

bands are also formed throughout the creep life.


If a material contains nonadherent or weakly bonded particles on

grain

bound-

aries, such as sulfides in steels, cavities are easily nucleated at these particles. For some types of precipitates,
interface

is

affected

by

the

the

strength

segregation

of

of

the

particle/matrix

trace impurities. Oxides and

carbides, for example, tend to nucleate cavities preferentially in the presence


of

(even

very

smal~)

quantities of sulfur. Whether cavity formation at these

interfaces occurs by decohesion, i.e. by the rupturing of atomic bonds,

or

by

vacancy condensation is not yet clear. Observations and theories related to the
segregation of impurities to interfaces were described in Chapter 8.
The precipitation of gases under high
assist

pressure

into

the

cavity

nuclei

may

cavity nucleation under conditions where hydrogen attack, oxygen attack

or helium embrittlement can take place (Chapter 9). It

is

unlikely,

however,

that these effects playa key role under ordinary creep rupture conditions.

17.2 Cavity Growth Rates and Rupture Lifetimes


for Instantaneous Nucleation
The growth of cavities by stress-directed grain

boundary

diffusion

has

been

analyzed in Chapter 11 for the limiting cases of equilibrium void shapes (rapid

17.

244

Summary on Cavity Nucleation and Growth

surface diffusion) and of crack-like void shapes (slow surface diffusion).


models

for

The

diffusive cavity growth lead to growth rates and rupture lifetimes

which differ substantially from what is usually observed.


The discrepancy can, at least partly, be resolved by noting that diffusive cavity growth rates should be constrained by the material surrounding a cavitating
boundary facet, if the strain rate is below a certain
(Chapter

12).

transition

strain

rate

In fact, it has been found that most tests on commercial alloys

are performed in the constrained regime. Observed

cavity

growch

rates

agree

with the constrained growth model.


However, the constrained growth model is not
rupture

lifetimes

for

suited

for

the

calculation

reasons pOinted out in Section 12.3. Nevertheless, the

time to cavity coalescence calculated from the constrained growth


agrees

model

often

with measured rupture lifetimes. In the limit of completely constrained

growth, the model predicts a constant


often

of

Monkman-Grant

product,

Et f ,

which

is

be

the

observed~

Theoretically, the joining of cavitating grain boundary facets


rate-controlling

step

under

constrained

should

growth conditions. This process has

been investigated using self-consistent models and other models for facet-facet
interactions in Section 12.6.
The inhibition of diffusive cavity growth by a lack of vacancies near the growing

cavities or by difficulties to accommodate atoms in particle/matrix inter-

faces has been considered in Chapter 13. The importance of the effect has
estimated

been

by linking it to the phenomenon of inhibited Coble creep. The analy-

sis suggests that inhibition should become effective at

stresses

below

those

usually applied in creep rupture tests.


The continuum plastic hole growth mechanism leads to a
product,

but

it

growth

Monkman-Grant

dominates only at high stresses (Chapter 14). In the experi-

ments evaluated, there was no


Cavity

constant

indication

that

this. mechanism

predominated.

at grain boundary ledges by grain boundary sliding was found to

be too slow to account for observed growth rates, since sliding in polycrystals
is usually constrained (Section 14.2). Similarly, diffusive growth with elastic
or creep accommodation played no

role

in

the

analysis

of

the

except, maybe, in the growth of cracks in ceramics (Chapter 15).

experiments

17.2

245

Cavity Growth Rates and Rupture Lifetimes

continuum hoe grO'Nth


eq. (11.1,7)

constrained or inhibited, with


creep accommodation

constrained with diffusive accommodation

eq. (11./,)

log (stress) ___

Fig. 17.1. Calculated cavity growth rates (schematic).

Figure 17.1 gives a somewhat simplified picture


calculated

from

of

the

cavity

growth

rates

various models and it shows schematically the ranges in which

different mechanisms predominate. Of course, real materials do not

necessarily

exhibit all of the mechanisms indicated depending on material parameters and on

constrained, diffusive accommodation

constrained or inhibited, creep


accommodation, eq. 02.11)

Q,;

....

eq. (11.23)

-~

fluctuations near sintenng

stress

'---~

crack-Ii ke diffusive
continuum hole growth, eq.

log (stress) - - - -

Fig. 17.2. Calculated lifetimes for instantaneous nucleation (schematic).

246

17.

Summary on Cavity Nucleation and Growth

the cavity size and spacing. If, for example, the surface diffusion coefficient
is

large, crack-like cavity growth may have no range of validity between equi-

librium diffusive growth and plastic hole growth. Figure


times

to

17.2

summarizes

the

cavity coalescence calculated for instantaneous cavity nucleation at

the beginning of the test. In many cases, the time


approximately

equal

to

the

time

to

cavity

coalescence

is

to rupture, but in the case of constrained

growth, this relation is questionable.

17.3 Rupture Lifetimes for Continuous Nucleation


In engineering alloys, cavity nucleation usually occurs continuously,
the

rupture

time

is

affected

by

the

nucleation

complicated manner. A formalism to deal with


Chapter

16.

The

theory

is

based

on

this

the

so

that

and growth kinetics in a

problem

was

developed

in

evolution law of the cavity size

distribution function, which incorporates both, nucleation and growth kinetics.


The

importance

of

continuous

growth mechanism. If
nucleation

leads

diffusive

to

much

versus instantaneous nucleation depends on the


cavity

greater

cavities had been present from the


growth,

on

the

other

hand,

growth

lifetimes
beginning

the

is

unconstrained,

than
of

would be obtained if all

the

test.

For

Fig.

16.4.

Using

illustrated

the models for free and constrained diffusive growth in

cOnjunction with continuous nucleation leads to rupture lifetimes


well

constrained

time to cavity coalescence is not strongly

influenced by the kinetics of cavity nucleation. This behavior was


in

continuous

which

agree

with measured data for many materials. However, it should be kept in mind

that the use of the


questionable.

constrained

growth

model

in

lifetime

calculations

is

18 Grain Boundary Cavitation Under


Creep-Fatigue Conditions

18.1 Micromechanisms of Creep-Fatigue Failure


Engineering structures are often subjected to variable, rather than merely constant,

loads. Cyclic loading causes a degradation of materials which is called

fatigue. Fatigue damage can be subdivided into two broad categories.


The first one consists in the formation and propagation
mechanism

predominates

at

room

of

microcracks.

This

temperature and is frequent also at elevated

temperatures. Its discussion is postponed until after

the

stress

and

strain

fields at cracks have been described (Chapter 28).


Second, fatigue failure may also occur by more or less
of

homogeneous

cavitation

grain boundaries over the whole cross section of a specimen, rather than by

the growth of a dominant crack. Cavitation is preferred at higher


and

temperatures

by loading wave-shapes which involve slow tension-going and fast compress-

ion-going strain rates, i.e. long hold times at tensile stresses (Fig. 18.1).

a.

a.

Fig. 18.1. (a) Balanced wave-shape. (b) Slow-fast, strain-reversed loading.


The strain is assumed to be prescribed, the stress response
of the material is shown schematically.

18.

248

Cavitation under Creep-Fatigue Conditions

In some cases, however, symmetric loading cycles may also cause cavitation. The
strong

effect of the wave-shape was demonstrated, for example, by Majumdar and

Maiya (1979), Sidey and Coffin (1979), Hales (1980) and Baik and Raj
1983).

While

the

crack

(1982a,b,

nucleation and propagation mode of fracture is often

very sensitive to the chemical environment (Coffin, 1972, Ericsson, 1979),

the

cavitational mode is relatively insensitive (Maiya, 1981). The transition between the cavitational mode and microcrack propagation depends strongly on material and wave-shape and is very sensitive to environment.
There are materials like the aluminum alloy studied by Baik and
which

do

Raj

(1982a,b)

not fail by cavitation in unidirectional creep but do so under slow-

fast cyclic loading. The reason for this behavior is that, under constant load,
failure

by

transgranular

hole growth or necking intervenes when the cavities

are still very small. In fatigue, however, plastic


which

are

hole

growth

and

necking,

strain-controlled processes, are reversed within each cycle. (Tests

are usually conducted under complete strain reversal). Diffusive cavity growth,
on

the

other

hand,

is

a stress-controlled process. Hence, the tensile mean

stress associated with slow-fast cycles leads to an overall enlargement of

the

cavities in each cycle.


The fatigue maps introduced by Collins, Sidey and Taplin (1979) are an
to

attempt

display the ranges of stress, temperature, loading frequency and wave-shape

within which different micromechanisms predominate. It is hopeless to


the

extensive

evaluate

literature on creep-fatigue-environment interactions and to in-

clude the results in a brief chapter of this book. Therefore only a few

examp-

les will be selected which highlight the most important aspects. There are more
comprehensive reviews on various aspects available, for example those by Coffin
(1977),

Taplin

and

Collins

(1978), Ericsson (1979), Tomkins (1981), Snowden

(1981) and the b06k edited by Skelton (1983). The cavitational mode of
failure

will

be

fatigue

described and analyzed in some detail next, while microcrack

propagation will be studied in Chapter 28.

18.2 Theories o{Cavitational Failure for Slow-Fast Fatigue Loading


If fatigue failure is brought about by the

growth

and

coalescence

of

grain

boundary cavities, one expects that the models for cavity nucleation and growth
described in the preceding sections should also be applicable to calculate

the

number of cycles to failure. There is, however, a remarkable observation which,

18.2

Theories of Cavitation Failure

249

at least at first sight, cannot be reconciled with any


discussed

so

far.

of

the

grain boundary cavitation even if the load cycles are balanced,


tensile

part

of

growth

models

A few materials, for example copper and magnesium, exhibit


i.e.,

if

the

the load cycle is identical with the compressive part except

that the sign is reversed. In this case, the cavity growth models would predict
that

the

average

cavity growth rate is zero (or even negative because of the

sintering stress). Nevertheless, copper cavitates readily under balanced cyclic


loading

both,

at

high

loading

frequencies

and low amplitudes (Saegusa and

Weertman, 1978), and at low frequencies and high amplitudes (Sidey and

Coffin,

1979, Baik and Raj, 1983). This problem will be addressed in Section 18.4.
Compared to balanced
unbalanced,

loading,

slow-fast

the

fatigue

life

drops

significantly

under

loading. This drop in fatigue life will be explained on

the basis of the cavity nucleation and growth models described in the preceding
sections.

Beere

and

Roberts

(1982) have examined this question earlier, but

with little success. They assumed the cavity density to be constant and,
for

this

maybe

reason, obtained poor agreement with low-cycle fatigue tests. In the

following it will turn out that continuous cavity nucleation is important.


18.2.1

Cycles to failure for unconstrained diffusive cavity growth

Here, fatigue failure by unconstrained diffusive cavity growth


The

cavity

is

growth rate is given by eq. (11.14). Neglecting possible transient

effects, for example by elastic accommodation, the result is valid


loading

as

considered.

well.

If

all

for

cyclic

cavities are already present at the beginning of a

test, integration of the cavity growth law leads to the rupture time, t f , given
in

eq. (11.18), where now the stress is replaced by its time average, o. If

v~

is the loading frequency, the number of cycles to failure is


(18.1)
This result will not have great practical

cr,

relevance

since

it

predicts

that

Nf ~ v~ /
whereas in the empirical Manson-Coffin law the freqency dependence
is much weaker and the main dependence is that on the applied nonelastic strain
range, i.e.,
(18.2)
with the empirical exponents a

0.2 to 0.4 and fl

1 to 4. (Many of

the

data

18.

250

Cavitation under Creep-Fatigue Conditions

which underly the Manson-Coffin law correspond to failure by crack propagation.


Therefore care must be exercised when these data are compared with models based
on

cavitation.

However, there are also well documented cases where eq. (18.2)

applies, while cavitation is the dominant failure mode).


In fatigue tests, just as
usually

nucleate

in

unidirectional

creep

rupture

tests,

cavities

continuously rather than instantaneously at the beginning of

the test. In creep rupture tests, it has been found empirically that the number
of cavities often increases in proportion to creep strain. Although the reasons
for this behavior are not quite clear, it is likely that it has to do with slip
band

formation

or

other

strain-related

processes, which may trigger cavity

nucleation. These processes do not depend on the sign of straining.


in

the

absence

Therefore,

of direct experimental evidence, we assume that, under cyclic

straining, too, the cavity nucleation rate is related to the magnitude

of

the

nonelastic strain rate by


J*

where

11'

11'

I Ene I ,

(18.3)

is an empirical factor which should ideally be of similar magnitude as

that found in creep rupture tests. Of course, cavity nucleation cannot continue
at a constant rate up to very high accumulated
cavity

density

would

become

strains,

unrealistically

relation, eq. (18.3), certainly ceases to be

since

high.

valid

otherwise

Therefore

when

the

the

cavity

growth

leads

the

simple
density

approaches a certain saturation value.


Continuous cavity nucleation combined with
rupture

lifetime,

diffusional

to

the

t f , given in eq. (16.9). For cyclic loading, the stress and

the nucleation rate must be replaced by their time averages, Ci and


over a cycle). From eq. (18.3) is

J* =

J*

(averaged

2 1l'~neVt' where ~ne is the nonelastic

cyclic strain range. This, together with eq. (16.9), gives the number of cycles
to failure for continuous, strain-controlled cavity nucleation,
(18.4)
Experimental results are often plotted as a function of
rate

in

related to

the

Et

tension-going

part

of

~ne

and of the

strain

the cycle, ~t' The loading frequency is

and to the compression-going strain rate,

Ec '

by
(18.5)

18.2

Theories of Cavitation Failure

251

Inserting this into eq. (18.4) gives the cycles to failure in terms

of

strain

range and strain rates:


h(~) kT
(1+Et /IE c

l)

Et

]2/5.

(18.6)

Q6DbG

The results given in eqs. (18.4) and (18.6) have a form similar to the

Manson-

Coffin law. (A more detailed comparison with experiments will follow later). In
addition to the dependences contained in the Manson-Coffin law, eqs. (18.4) and
(18.6) also predict a dependence on the average stress 0; Nf goes to infinity
if 0 goes to zero, i.e., if the load cycles are balanced. In this limiting
case, a new cavity growth mechanism must become effective in certain materials,
for example magnesium or copper, which fail by cavitation under balanced cyclic
loading, albeit with a much longer fatigue life than under slow-fast loading.
18.2.2

Cycles to failure for plastic hole grOwth

At high strain rates, the plastic hole growth mechanism described in Chapter 14
leads to higher growth rates than does the diffusive mechanism. Whether or not
plastic hole growth predominates depends on the magnitude of the characteristic
diffusive

length,

defined

i,

in

eq.

spacing. The author has calculated the

(4.17)
value

of

compared with cavity size and


i

for

the

experiments

of

Majumdar and Maiya (1979), Sidey and Coffin (1979) and Baik and Raj (1983), and
has found that i ranges from 0.4
and

is

~m

to 1.2

~m

for the tests on stainless

steel

an order of magnitude smaller for copper. Values of i in the submicron

range indicate that plastic hole growth plays on important role,

since

cavity

size and spacing are usually of the order of a micrometer.


Usually, fatigue tests are done such that the strain is fully reversed in

each

cycle. Since plastic hole growth is strain controlled, the net growth increment
in a cycle is zero. In reversed strain tests, the plastic hole growth mechanism
can

only

contribute

to

continuing

growth, if the cavity surface created in

tension is not completely removed in compression. Such an irreversibility could


be

caused

by

residual

elements

segregating

to the fresh surface. Since no

quantitative information is available on these irreversibilities,


be

said

irreversible surface formation. In this case, the cavity surface


proportion

nothing

can

about the cycles to failure, except in the extreme case of completely


to

6E

increases

in

in each cycle. During the compressive part of the cycle the

cavity is deformed into a flat crack-like shape, but with no rewelding

of

the

18.

252

crack

faces,

and

in

the

Cavity coalescence occurs

next
when

Cavitation under Creep-Fatigue Conditions

cycle another increment in surface is added.


the

cumulative

tensile

strain

reaches

the

critical value given in eqs. (14.24) or (16.22) for instantaneous or continuous


nucleation, respectively. Hence, in this limiting case, the number of cycles to
failure is

(18.7)
The strain to failure is predicted by the plastic hole growth mechanism in
order

of

few

hundred

per

the

cent, i.e. under cyclic loading conditions the

cumulative tensile strain to failure should have such a value,

if

indeed

the

formation of surface by plastic hole growth is irreversible. There is no reason


why the mechanism should not work for balanced cycling as well.
the

effect

of

the

irreversibilities,

which

Unfortunately,

are essential for a net cavity

growth increment per cycle, cannot be quantified easily.

18.2.3

Cycles to failure for constrained growth

For balanced cycles, the constrained cavity growth model (Chapter 12)
no

net

predicts

growth increment per cycle. Even unbalanced cycles, if they are strain

reversed and if cavity growth is fully creep-constrained in both parts


cycle,

ing ramp is so fast that the rate of cavity shrinkage is limited


rather

than

the

the

by

diffusion

by the polycrystalline constraint, a net growth increment results

in each cycle.
during

of

do not lead to a net growth increment. If, however, the compression-go-

I~

the extreme case of

compressive

part

of

the

very

fast

cycle

is

compression,

the

shrinkage

negligible,

and

only

the

tension-going ramp contributes to (constrained or unconstrained) cavity growth.


Here we assume that the tension-going ramp is so slow
creep-constrained

that

cavity

growth

is

during that part of the cycle, while cavity shrinkage during

rapid compression is assumed to be negligible. Then cavity

coalescence

occurs

the cumulative tensile creep strain, NfAE ne , reaches the critical value,
Ef ~ 0.37 Aid aC90rding to eq. (16.19). If cavity coalescence implies failure,
the fatigue life is given by

when

(18.8)
which is independent of loading frequency and temperature.
this

In

agreement

with

result, but without reference to a particular experimental investigation,

18.2

Theories of Cavitation Failure

253

Collin, Sidey and Taplin (1979) remark that the product NfAEne
low

as

the

become

as

tensile ductility, i.e. of the order of 10%, at very low frequen-

cies. In the majority of cases, however, the observed fatigue


stantially

may
lives

are

sub-

greater than predicted by eq. (18.8). This is understandable, since

most tests are done outside the range of constrained growth.


18.2.4

Summary of fatigue lifetimes for different cavity growth mechanisms

The results of the preceding three subsections are shown schematically in


18.2.

The

fatigue

life

Fig.

is plotted as a function of temperature, with strain

range, frequency and the ratio Et/lEcl < 1) held constant. As the figure
shows, diffusive cavity growth prevails at intermediate temperatures. As long
as the accumulated strain 2AEneNf is not too large, cavities nucleate
continuously throughout the test and eq. ( 1 8.4) applies. The effects of
Nf ~ (vtT/6Db) 2/5
Nf , on the other hand, the cumulative strain becomes so large

frequency and temperature in this range are


For

very

large

described

by

that the cavity density saturates early in the fatigue life, so that it can
considered

as

being

effectively constant. Then the dependencies on frequency

and temperature are given by Nf

vt TI6D b according to eq. (18.1).

The range of diffusive growth is bounded at lower temperatures by


hole growth mechanism

the

plastic

which is independent of temperature and frequency if the

plastic hole
growthJl8.7 )

be

trans granular
crack
growth

instantaneous nucleation (18.1)


diffusive growth
nucleation,
constrained
(18.1.)
. . . . . . growth in
tension and
I
compression
constrained
growth in tensionJI8.8)
I

Fig. 18.2. Fatigue lifetime vs. temperature for slow-fast tests


predicted by theories for cavitation failure (schematic).

254

18.

Cavitation under Creep-Fatigue Conditions

formation of cavity surface is completely irreversible. Of course, the irreversibility

in surface formation, which is necessary for accumulated hole growth,

may depend on frequency and temperature in an as yet unknown way. In most

mat-

erials, the formation and propagation of fatigue cracks will intervene at lower
temperatures before plastic hole growth leads to failure
At high temperatures, diffusive growth becomes so fast that the imposed

strain

rate (which is kept constant in Fig. 18.2) can no longer accommodate the cavity
volume. Growth in the slow tension-going ramp becomes

constrained

first.

The

fatigue life is then approximately given by eq. (18.8), which is independent of


frequency and temperature.
At even higher
compressive

temperature,

part

the

shrinkage

of

the

cavities

in

the

short

of the cycle becomes significant compared to the constrained

growth during tension. In the limit of very fast diffusion (i.e., at very
temperatures),

cavity

growth

becomes

constrained

high

in either direction which

results in zero net growth and prolonged fatigue lifetimes.

18.3 Comparison with Results of Slow-Fast rests


18.3.1

Low cycle fatigue tests on AI-5%Mg

Baik and Raj (1982a,b) performed creep-fatigue tests on


high

an

AI-5%Mg

vacuum with a loadig frequency of 0.2 cycles per minute and a

strain range of 8ne = 1.6%.

Under

balanced

cyclic

loading,

alloy

in

on-elastic

they

observed

failure by transgranular microcrack propagation. Of particular interest for our


purposes are their unbalanced slow-fast tests with a tension-going strain rate
of t = 7.5'10-5 /s and a hundred times greater compression-going strain rate.
For this wave-shape, the fracture mode

was

intergranular

cavitation

in

the

temperature range 450 K to 550 K, and the number of cycles to failure exhibit a
behavior which is compatible with that predicted in Fig. 18.2.
It should be noted that Baik and Raj (1982a,b) offer a slightly
planation

ed. Following Min and Raj (1979a,b) they


plays

different

ex-

for their observation than ours, but the two explanations are relatcentral

role

in

argue

that

grain

boundary

sliding

cavitation failure by opening up wedge cracks. Now

grain boundary sliding in a polycrystal is constrained just as diffusive cavitation is constrained. The similarities are evident if the shear crack model for

18.3

Comparison with Results of Slow-Fast Tests

255

sliding (Chapter 7) and the tensile crack model for cavitation (Chapter 12) are
compared.

Therefore it is clear that, if the behavior can be explained by con-

strained cavity growth, it can also be explained

by

constrained

sliding.

If

sliding is constrained in the tension-going ramp and is diffusion-controlled in


the (faster) compression-going ramp, then a sliding
may

offset

accumulates.

This

open up wedge cracks and may lead to failure. The primary difference betw-

een the explanations by constrained cavitation and by

constrained

sliding

is

that the constrained sliding model predicts a sliding offset not only for slowfast tests, but also for fast-slow tests, whereas constrained cavitation

takes

place only in slow-fast tests. Since slow-fast cycles are generally observed to
be far more damaging than fast-slow cycles, the explanation based on constrained cavitation is preferred here.
18.3.2

Low-cycle fatigue tests on nickel

Lim and Raj (1984) tested pure nickel in vacuum with a frequency of 1 cycle per
= 310 =4 Is, IEC I = 1.510 -2 Is, and AEne = 1.35%. Up to a temperature
minute, E
t
of 573 K, failure occurred by transgranular crack propagation, whereas above
700 K,

cavitation

predominated.

The measured fatigue lives are shown in Fig.

18.3 together

with the prediction of eq. (18.4) using the following numerical


values: a' = 1.2.10 11 1m2 (fitted since unknown), oD bo = 3.510- 15 m3 /s,

la'

x
10 3

.r\ \.M
\ \
"\, .M
\
\

eq. (18.1)

x\
fvf

Ni ~=1,35%
x Cu L1=1,65%

Icpm

;;::,-

la'
I

lntergronular
.!::fixed

Iransgranular
10

3(1)

to)

5(J)

6ClJ
7!lJ 8I1J
T inK _ _

9JJ

1a:rJ

Fig. 18.3. Fatigue life of pure nickel and pure copper vs. temperature. Data
from Lim and Raj (1984b) and Baik and Raj (1983). Comparison with models.

256

18.

Cavitation under Creep-Fatigue Conditions

3 = 115 kJ/mol, wf = n/4, h(~) = 0.61, g = 1.0910 -29 m,


a = 400 MPa at 573 K
and a = 200 MPa at 873 K (estimated). The fitted value of the cavity nucleation
Qb

rate per unit strain, a', is within the

range

of

values

observed

rupture tests of other materials. It leads to a cavity spacing of A =

in

creep

l~m

after

310 cycles, which is compatible with the final spacing observed by Lim and Raj.
The

fatigue

life

in

the

whole range where cavitation prevails is less than

310 cycles so that continuous nucleation, which

underlies

eq.

(18.4),

is

valid assumption at temperatures above 650 0 K.


When the fatigue lifetime exceeds 310 cycles substantially, the cavity
saturates

and

eq.

(18.1),

denSity

which is based on instantaneous nucleation at the

beginning of the test, becomes the appropriate equation. The dashed-dotted line
in

18.3 represents eq. (18.1) with A = l~m,

Fig.

It has a very limited range of validity


temperatures

trans granular

crack

in

growth

this

a = 400
case,

intervenes,

MPa and h(~)


since

towards

while

towards

0.66.
lower
higher

temperatures continuous nucleation starts to playa dominant role.


The temperature at which the constraint starts to be effective can be estimated
from eq. (12.6). With A = 1.5 ~m, d = 150 ~m, q(w) = 1 and q' = 12.5, eq.
(12.6) predicts that cavitation in the tension-going ramp
strained

starts

be

con-

at around 1200 K. The constraint on the compression-going ramp starts

only beyond 1600 K, i.e., far outside the experimental range.


an

to

increase

Correspondingly,

in fatigue life at high temperatures is not observed, which would

be expected if cavitation were constrained in both parts of the cycle.


18.3.3

Low-cycle fatigue tests on copper

Figure 18.3 also shows data on OFHC-copper measured by Baik and Raj (1983). The
= 3.2 10 -4 Is, 1 Ec =
loading
frequency
was v t = 1 cpm, aEne = 1.65%, E
l
t
1.6 010-2 /s. The material response in terms of stress was
= 80 MPa at 573 K
0

and

a=

60 MPa at 673 K. Above 473 K failure occurred by cavitation. Not shown

in Fig. 18.3 are the experimental results on balanced cyclic loading for
the

fatigue

lives

were

factor

which

3 higher than for slow-fast loading above

373 K. At and below 373 K, where transgranular crack growth leads

to

failure,

the wave shape had no effect.


The theoretical curve in Fig. 18.3 represents eq.

(18.4) with the following


numerical values substituted for copper: oD bo = 005 010- 14 m3 /s, Qb = 104 kJ/mol,
g = 1.18 10-29 m3 , a' = 1.5 10 12 /m2 (fitted). This nucleation rate leads to a
0

18.3

257

Comparison with Results of Slow-Fast Tests

cavity

spacing

of

within

0.3~

100 cycles.

The agreement of the predicted

number of cycles to failure with the observed data


with
d

other

= 75~

models

and A =

(12.6),

the

such

0.5~

as

eq.

constraint

is

found

not

perfect,

although

(18.1) the agreement would be far worse. If

(assumed since

ion-going ramp and at 1300 K for the

is

not

reported)

are

inserted

into

eq.

to become effective at 900 K for the tenscompression-going

ramp.

At

least,

the

latter temperature is far outside the range covered experimentally.


18.3.4

Low cycle fatigue tests on austenitic steel

Majumdar and Maiya (1979) conducted low-cycle fatigue test on Type

304

stain-

less steel at 866 K for various wave-shapes and frequencies. They observed that
failure occurred by intergranular cavitation in slow-fast tests. Otherwise
fracture

mode

was

the

trans granular with well-defined striations on the fracture

surface. Table 18. 1 shows the number of cycles to failure

for

the

slow-fast

tests together with the strain rates and the associated stresses.
Table 18.1 also includes test results by Sidey and Coffin (1979)
material

on

the

same

for the same type of slow-fast fatigue loading. The fracture mode for

these tests was also intergranular cavitation.


The last column of Table 18.1 shows the theoretical
(18.4).

The

rate

of

cavity

nucleation

result

material

Qb = 167.4 kJ/mol,

Source

parameters
were
taken
29
3
1.21.10m in agreement

ne
in %

T
in

fit

eq.

data. The
oD bo = 2.10- 13 m3 /s,
the values generally

as
with

the

V.L1
in s

in MPa

Nf
(exp)

870

0.01

= 200

847

815

by

is not reported in the experimental

papers. It was assumed to be a' = 1.8.10 10 /m 2 in order to


remaining

predicted

Nf
(theor)

Majumdar and
Maiya (1979)

0.7
0.7

870

0.0004

= 190

261

230

Sidey and
Coffin (1979)

1.8

923

0.0017

150

148

152

1.9

1093

0.0022

50

52

70

Table 18.1. Fatigue lifetime of Type 304 stainless steel under


slow-fast loading cycles. Theoretical values from eq. (18.4).

Cavitation under Creep-Fatigue Conditions

18.

258

reported. With the nucleation rate


reasonably

well

over

wide

fitted,

range

eq.

(18.4)

represents

the

of temperatures and over two decades in

frequency. From eq. (12.6) it is concluded that all tests in Table 18.1
for

that

data
except

at the highest temperature (1093 K) should be well outside the range

of constrained cavity growth which begins at 1200 K, if

A=

and

2~m

d =

50~

are substituted in eq. (12.6).

18.4 Why Do Cavities Grow under Balanced Cyclic Loading?


There have been several attempts to explain why some materials
balanced

cyclic

loading

with

zero

cavitate

under

mean stress. One possibility was already

described in Section 18.2.2, that is cavitation by plastic hole growth combined


with

some irreversibility in surface formation. Several other suggestions will

be described next.
Skelton (1966) and Weertman (1974) argue
stress-directed

diffusion,

which

that

the

fundamental

equation

is linear in stress according to eq. (4.4),

should in fact be nonlinear. In tension, the production of atomic vacancies


encouraged

is

so that their density and hence the diffusion coefficient increase,

whereas in compression the


slightly

for

more

during

the

reverse

is

true.

Therefore,

the

cavities

grow

tensile part of the loading cycle than they shrink

during the compressive part, and this acts against the tendency of the cavities
to

shrink by sintering forces. However, the nonlinear effects are of the order
which is so small compared to

020/kT

that the sintering

stress

can

usually

not be overcome except for unrealistically large cavities.


Another explanation for cavity growth under zero mean stress

was

proposed

by

Weertman (1979). She argues that for some transient time after the beginning of
the test, effects of elasticity enhance cavity growth as described
15.2.

Under

in

Section

cyclic loading, the growth rate oscillates with the amplitude de-

creasing during the elastic transient. It is then clear that a compressive half
wave cannot quite compensate the growth increment of the preceding tensile half
wave, which was slightly higher. Therefore, depending on whether cycling starts
with

tensile

or

compressive

half wave, the transient results in a net

growth or shrinkage. If this mechanism is to lead to cavity coalescence it

can

do so only during the elastic transient, since afterwards no further net growth
occurs. However, the elastic transients are generally much shorter than typical
test

durations. It is therefore hard to believe that elasticity effects playa

18.4

Why Do Cavities Grow under Balanced Loading?

259

central role in cavitation under balanced cyclic loading.


Cavities might also be formed by the condensation of the vacancies which cyclic
loading

is

known to produce, for example by the motion of dislocation jogs or

by the annihilation of close edge

dislocation

occur

as

during

monotonic

loading

well,

dipoles.

(The

same

processes

but to a smaller extent since the

cumulative strain is much smaller). In fact, neutron scattering experiments


Kettunen

al

of

indicate that even at room temperature small voids of


2 nm size and a volume fraction of 10-5 are formed, however, not necessarily on

grain

et

(1981)

boundaries.

Thus,

the

old idea that cavitation is driven by a vacancy

supersaturation (Section 6.2), which had been

dismissed

under

creep

rupture

conditions, might be applicable under balanced fatigue-loading conditions.


Further, on a microscopic level, certain portions of the material may well
perience

stressing

experience a compressive mean stress, even under balanced applied


may

arise

from

ex-

cycle with a tensile mean stress, while other locations

asymmetric

loads.

This

plastic yielding as in body-centered cubic metals

which tend to activate different slip systems in tension and

compression.

The

accumulated grain boundary sliding offsets observed by Evans and Skelton (1969)
and by Westwood and Taplin (1975) are also evidence for asymmetries of

deform-

ation processes on a microscopic level. However, the magnitude of the stresses,


which result from these asymmetries, is not precisely known
fact

be

too

small

and

they

may

in

in comparison with the sintering forces to explain cavity

growth.
Finally, the possibility should be kept in mind that residual gases may be precipitated

into cavity nuclei, and the gas pressure can provide the mean stress

for cavity growth.

18.5 Discussion
The dependence of the fatigue lifetime on temperature and on loading
obtained

in a limited number of slow-fast tests could be explained by a theory

based on (unconstrained) diffusive cavity


However,

frequency

in

growth

and

continuous

nucleation.

all tests considered, the strain range was either constant or was

not varied considerably. A great number of tests has been done by other workers
showing
Nf

oc

that

the

dependence

of

fatigue

life

on strain range has the form

(~ne) -B ,where B has typical values between 1 and 1.7 ( Lloyd and Wareing,

Cavitation under Creep-Fatigue ConditionE

18.

260

1981).

On

constant

the

other

frequency

hand,

or

eqs.

strain

(18.4) and (18.6) predict

rate,

respectively.

Thus,

a = 0.6
the

or 1 for

theoretical

dependence on strain range is weaker than the observed one.


The discrepancy may be due to the fact that many of the reported data refer
failure

by

crack

propagation

rather

fracture was not investigated.


intergranular

cavitation

than

Generally,

and,

at

the

to

by cavitation. Often the mode of

longer
same

tensile

hold

times

time, reduce the exponent

favor

a but

seldom to values below 1.


In conclusion, eq. (18.4) describes the fatigue life well in
where

the

dominant

failure

mode

certain

tests,

is intergranular cavitation, as far as the

temperature dependence is concerned. The predicted


range,

slow-fast

dependence

on

the

strain

however, is weaker than that generally observed, although it is not yet


whether

the

agreement

would

be

better

if

the

tests

were

done

systematically in the range of cavitation failure.


Conclusive theories on cavitation failure under balanced cyclic loading are not
yet

available.

Finally, the reader is reminded that the subject of creep-fat-

igue failure is taken up again in Part III where failure by


is considered.

crack

prupagation

Part III
Creep Crack Growth
and Creep-Fatigue Crack Growth

19 Introduction to Part III

19.1 The Relevance of Cracks


Creep crack growth is the time-dependent extension of a

macroscopic

crack

at

elevated temperatures under more or less constant load. For cyclic loading conditions, one speaks of creep-fatigue crack

growth.

distinct

is a crack which is larger than the

from

grain

boundary

cavities,

A macroscopic

structural lengths of the material which are relevant

for

crack

crack,
growth

as
~for

example, the grain size). Failure by crack growth may predominate over homogeneous grain boundary cavitation in the whole cross section, if a crack is
iated

at

init-

pre-existing defect or a sharp notch early in the lifetime of the

component. A dominant crack may also be

nucleated

by

some

corrosion-related

process before profuse cavitation occurs within the bulk of the material. Under
cyclic loading conditions, surface microcracks are generally

formed

early

in

the lifetime of the specimen.


After the occurrence of a few premature failures in
plants,

generating

which were attributed to creep crack growth, the problem started to be

discussed in the literature in the


Harrison

electric-power

early

1970's

(Siverns

and

Price,

subject area of creep crack growth was reviewed repeatedly, for example by
Leeuwen

1970,

and Sandor, 1971, Robson, 1972, Thornton, 1972). In the meantime, the
(1977),

van

Ellison and Harper (1978), Pilkington (1979), Speidel (1981),

Sadananda and Shahinian (1981a) and by Riedel (1984a, 1985a).


If creep crack growth is a potential failure mode, design codes and

inspection

standards should provide a means to assess the relevance of cracks in high-temperature components. Such an assessment should include crack-like defects which
have

actually

been

detected

by non-destructive evaluation, and, in critical

applications, one must assume that cracks of

certain

maximum

size,

which

might have escaped detection, are present. A lifetime prediction should then be
based on the expected growth rates of such defects in future service.

19.

264

Current

codes, however, usually rely on the unrealistic requirement that high-

temperature components should not contain any


better

Introduction to Part II

crack-like

defects

at

all.

understanding of the mechanics and the mechanisms of creep crack growth

would help to improve the situation.


Incidentally, to avoid a frequent misconception, it should be remarked that the
following

analyses

are

not

aimed

at

cracks which result from the eventual

inter-linkage of grain boundary cavities in more or less homogeneously stressed


specimens

or components such as straight pipes under internal pressure. Incip-

ient crack formation in uniformly strained specimens usually indicates that the
lifetime

is

practically

exhausted. Rather, we are interested in cracks which

are present from the beginning or are nucleated early and which grow in
wise

virtually

other-

uncavitated material. The border-line between failure by homo-

geneous cavitation and failure by crack growth cannot be

easily

predicted

or

defined in general terms.


In the analysis of creep crack growth

it

is

convenient

to

distinguish

two

aspects; namely, (1) the continuum-mechanical deformation fields in the cracked


body, and (2) the micromechanisms of crack growth which operate near the
tip

(e.g.

grain

crack

boundary cavitation). These two aspects can often be treated

separately, an approach which will be adopted in the following chapters. A combined treatment of deformation and damage will be described in Chapter 27.

19.2 The First Aspect: Deformation Fields in Cracked Bodies


In a first step, we analyze the continuum-mechanical deformation
mathematically

sharp

crack

fields

at

neglecting the effect of damage on the stress and

strain distributions. Important results of such crack analyses are the asymptotic fields near crack tips and the load parameters (K I , J, C* or others depending on the material law) which determine the strength of the asymptotic fields.
As

long

as damage is confined to a sufficiently small 'process zone' near the

crack tip, the continuum deformation fields encompass and control the evolution
of

damage

and

therefore

the respective load parameters should control crack

growth. (There are possible exceptions to that rule: if, for example, the crack
growth

rate

is

atmosphere to the
obviously

controlled by the penetration of a corrosive species from the


crack-tip

region,

the

crack-tip

deformation

fields

are

not the only controlling factor). In many cases, however, it is true

that the crack-tip fields control the crack growth rates.

19.2

The First Aspect: Deformation Fields

265

The main task in the following chapters will be to identify the load parameters
which

determine the crack-tip fields, and to explore their ranges of validity.

Limitations to the applicability of the


crack-tip

blunting

finiteness of the

(Section
specimen

22.1),
thickness

various

load

parameters

three-dimensional
(Section

effects

22.2),

arise
due

deviations

from

to

the

from

the

idealized material laws employed (Chapters 23 to 26), and from the evolution of
damage in a growing process zone (Chapter 27). The practical value of the

load

parameters lies in the fact that creep crack growth rates measured in the laboratory can be transferred to structures of different size and shape
appropriate

load

parameter

the

parameter. This transfer of growth rates can be done without

explicit reference to the micromechanisms of crack growth. If


load

using

has

the

the

appropriate

same value in the cracked structure as it had in the

test specimen, the crack-tip fields will also be the same and the fracture processes leading to crack growth must respond in an identical manner.
In the early papers on creep crack growth, the stress intenSity factor, KI , and
the net section stress, Gnet , were considered as candidates for appropriate
load parameters. While KI turned out to have its range of validity, the net
section stress has no sound theoretical basis, but, nevertheless, was sometimes
successfully applied. However, the most important load
creep

parameter

by Ohji, Ogura and Kubo (1974) and by Landes and Begley (1976). The
of

C*

to

describe

crack growth macroscopically turned out to be the C*-integral introduced


for

usefulness

t!1e more ductile alloys was subsequently demonstrated by many wor-

kers. Among the first were Nikbin, Webster and Turner

(1976),

Koterazawa

and

Mori (1977), Taira, Ohtani and Kitamura (1979), Saxena (1980), and Ohji, Ogura,
Kubo and Katada (1980). The boundaries between the ranges of validity of KI and
C*

were established theoretically by Riedel and Rice (1980) and by Ohji, Ogura

and Kubo (1980).

19.3 The Second Aspect: Micromechanisms


19.3.1

Grain boundary cavitation ahead of the crack tip

Creep crack growth often occurs by the nucleation and growth of grain

boundary

cavities ahead of the main crack and their final coalescence with the crack.
This will be modeled in Chapters 21 and 23.3 by combining the laws for cavity
growth with the stress field of the growing main crack. A rather conclusive
picture of creep crack growth in the absence of corrosive effects emerges,
since the models predict the observed behavior of several materials well.

266

19.

19.3.2

Introduction to Part II

Corrosive processes at the crack tip

It is known that the chemical environment exerts


creep

crack

strong

influence

on

the

growth rate in some materials (Speidel, 1981). This is especially

so under cyclic loading conditions, which will be considered in Chapter 28. But
also

under

constant load, many of the superalloys are very sensitive to atmo-

spheres containing oxygen or sulfur

(Sadananda

and

Shahinian,

1980a,

1983;

Floreen, 1983; Stucke et aI, 1984). The crack growth rate may increase 100-fold
in air compared to an inert environment, even if unstressed
hardly

any

material

exhibits

visible oxidation under otherwise identical conditions. This is so

because structural high-temperature alloys are designed to

develop

protective

surface oxide layers (usually cr 20 3 ) in oxidizing environments. At a stressed


crack tip, however, the layer ruptures continuously and oxygen atoms can penetrate

into

the material, which they do preferentially along grain boundaries.

It is not yet known precisely how oxygen accelerates creep crack growth to
large

extent

observed.

Possibly, it reacts with constituents of the alloy to

form brittle oxides which fracture readily. Oxygen may also attack carbides
the

carbide/matrix
internal

had

oxidation and therefore they exert an opening force on the crack faces.

Finally, the grain boundaries can be embrittled by


demonstrated

segregation

of

oxygen

Ferritic

and

accelerate

austenitic

creep

crack

growth

compared

insensitive.

to

steels (Sadananda and Shahinian, 1980b) and

even some of the cast superalloys (Shahinian and Sadananda, 1984) appear to
relatively

as

by Nieh and Nix (1981) for copper. It should be mentioned that an

air environment does not always


vacuum.

up

gas pressure which promotes cavitation. Further, the oxide spikes

formed along grain boundaries usually have a greater volume than the metal
before

at

interface which leads to decohesion of the carbides. Under

special circumstances, gaseous reaction products (e.g. CO or CO 2 ) can build


an

the

be

The diverse behavior of superalloys might result from

their microstructure, which varies from material to material and has a pronounced effect on creep crack growth rates (Floreen, 1983).
Little work has been done to model the corrosive effects on creep crack

growth

rates quantitatively. In fact, the author knows of only one publication on that
subject (McClintock and Bassani, 1981). Before a more detailed understanding of
the

corrosive

processes

at crack tips emerges, this remains an area which is

necessarily dominated by metallurgical empiricism.

20 Nonlinear Viscous Materials and the Use of C*

Creep deformation of metals and alloys at elevated temperatures can oe described, to a first approximation, as nonlinear viscous. Viscous behavior is defined
by the existence of a unique relationship between the strain

rate

tensor

and

the current stress tensor, Eij = f(Oij). This corresponds to steady-state, or


secondary, creep while primary-creep effects and the elastic response of materials

are

neglected.

Although real materials generally do not exhibit exactly

nonlinear viscous behavior, the results of this section, in particular the


of

C*

are

often

use

justified as a reasonable approximation within limits to be

specified in the following sections.

20.1 Definiton ofthe C*-Integral


In fracture mechanics testing under elevated temperature creep conditions,
C*-integral

plays

the

an important role (Landes and Begley, 1976; Ohji, Ogura and

Kubo, 1974; Nikbin, Webster and Turner, 1976). For

two-dimensional,

planar

crack, C* is defined ,as the contour integral


(20.1 )
with W*
For

= JOijdC ij ; here, 0ij is uniquely related to ij by the


material law.
the definition of the symbols, see Fig. 20.1 and Section 3.4.3. In viscous

materials, the C*-integral is independent of the choice


proof

of

of

the

path

r.

The

the path-independence parallels that of the path-independence of the

J-integral in elastic. materials (Rice, 1968a,b), and will not be repeated here.
In principle, C* can be measured by testing a pair of specimens having infinit~simally

different crack lengths, a and a

da, and which are otherwise identi-

cal. From the measured load-displacement rate records of the two specimens,

C*

can be determined by differentiating the area under the load-displacement rate

Nonlinear Viscous Material and the Use of C*

20.

268

p
Xz

x,
1-----

W -----001

Fig. 20.1. Definitions.

curve with respect to crack length:


(20.2)
Here, P1 is load per unit specimen thickness and A is the displacement rate of
the load points. Landes and Begley (1976) applied eq. (20.2) to the measurement
of C* and described its use in greater detail. Alternately, C* can
ated

from

eq.

(20.1)

provided

that

be

calcul-

the stress and strain rate fields have

already been determined for the configuration considered. This usually requires
numerical

techniques

such

as

the

finite

element

method. This approach is

particularly useful for power-law viscous materials, which are described next.

20.2 Stress Fields and the C*-Integral in


Power-Law Viscous Materials
Power-law viscous materials are characterized by the special stress-strain rate
relation
(3/2) B

which simplifies to e

(20.3)

Ban (Norton's creep law) in uniaxial tension. The prime

denotes the deviator, ae is the equivalent tensile stress, and Band n are
material parameters. In agreement with the behavior of real materials, eq.
(20.3) describes incompressible deformation, since the trace of the strain rate
is zero,

E11
..

a!.

11

= 0, by definition of the deviator.

20.2

Stress Fields and C*

269

A power-law constitutive equation simplifies


strain

rate

fields

because

of

the

the

description

of

stress

and

scaling properties described in Section

3.4.2. All stress components at any point of the body increase in proportion to
the

applied

load

while

strain rate components increase in proportion to the

n'th power of the load. In configurations which are symmetrical with respect to
the

crack

plane. it is convenient to introduce the net section stress defined

as load divided by the area of the uncracked ligament ahead of the crack.

Then

the scaling properties can be expressed as


(20.4)
(20.5)
(20.6)
Here. xi represents the spatial variables, W represents the

in-plane

specimen

and a is crack length. The dimensionless functions Gij , Eij and Ui


depend on only two dimensionless parameters. a/Wand n, but not on the load nor
dimensions,

on the material parameter B.


20.2.1

The C*-integral in power-law viscous materials

If the scaling laws, eqs. (20.4) to (20.6) are inserted

into

eq.

(20.1)

the

C*-integral takes the form


C*

n+1
a B Gnet g1(a/w,n).

(20.7)

where the dimensionless function g1 is an abbreviation for the integral


(20.8)
fOijdE ij xi = xi/a and ds = ds/a. Evaluation of the integral
the knowle~ge of the field quantities. These have been computed by

with W*
requires

several workers
collection
~hich

of

using
results

the

finite

element

method.

is known as 'The Plastic Fracture Handbook', or

Their

tabulations

The

most

comprehensive

was given by Kumar, German and Shih (1981) in a report


as

the

EPRI

Handbook.

are in terms of the J-integral for time-independent elastic

materials. Due to the elastic-viscous analogy (Section 3.4.1) one can write
instead

of

J,

and displacement rate,

A,

C*

instead of ~ if their combination of

270

20.

Nonlinear Viscous Material and the Use of C*

material parameters, a o lan,


is replaced by B. Further, they chose to use a
0
notation based on the plastic limit load rather than on net section stress, but
both ways of writing the results are

connected

through

elementary

algebraic

relations. For the compact specimen in plane strain, for example, they write
n+1
(W-a) B h 1 (a/W,n) [ P1 /[1.455n(W-a)] ]
,

C*

(20.9)

where a and W were defined in Fig. 20.1, P1 is load per unit thickness, h1 is
tabulated by Kumar et al (1981) (see also Appendix C) and n is given there in
closed form as a function of a/W. Comparison with our notation

in

eq.

(20.7)

shows that g1 can be expressed by quantities computed by Kumar et al through


(20.10)
A practical drawback of the calculation of C* using eqs. (20.7)

or

(20.9)

is

that these formulas strongly rely on Norton's power law, whereas real materials
often deviate from a strict power law. In addition,
markedly

upon

specimens

of

appropriate

whether
finite

(see

plane-stress
thickness

the

discussion

it

or
is

in

the

plane-strain
not

Section

clear

which

displacement

are

depends
used.

In

case

is

limiting

22.2). For these reasons it is

advisable to measure a displacement rate, preferably at a


calculated

function
values

location

for

which

rates are available in the Plastic Fracture Handbook.

For the compact specimen, results are given for the load line deflection rate

=a

h 3 (a/W,n) [P 1 /[1.455n(W-a)] ]n

(20.11)

and for the crack opening displacement rate at the edge of the specimen
(20.12)

5e are relative
between the respective pOints above and below the symmetry

where h2 and h3 are tabulated dimensionless quantities; ~ and


displacement

rates

line of the specimen. Combining eqs. (20.9) and (20.11) leads to the

following

useful formulas for C*:


C* = g2(a/W,n) anet ~
C*

(20.13)
(20.14)

20.2

Stress Fields and C*

271

The dimensionless factors g2 and g3 are related to the notation of Kumar et


(1981) through
g3

(W/a-1)

h1 I h 31+1/n.

al

(20.15)

Plots and approximation formulas for g2 are given in Appendix C. Equation


(20.13) allows for a determination of C* which is independent of the material
parameter B and is insensitive to the choice of plane strain
and

to

or

plane

stress

the value of n. This only mild dependence is plausible in the light of

approximate, but accurate, analytic solutions for a crack in an infinite


These

solutions

were

developed

body.

by He and Hutchinson (1981) and were already

&is

used in connection with the constrained cavity growth model in Part II. If
the

relative

displacement rate between the two crack faces in the center of a

plane-strain crack subjected to a remotely applied tensile stress, o~, C*


approximately given by C* = (~/4) o~ &. This is indeed independent of n.
20.2.2

is

Crack-tip fields in power-law viscous materials

While the calculation of the whole


requires

numerical

techniques,

stress
the

field

in

finite

body

usually

asymptotic field near a crack tip can be

obtained analytically. These so-called HRR-fields were already shown in Section


3.4.4, eq. (3.25), where the symbols were also explained. The result is
0ij = (

C* )l/(n+1)
InBr

a.. (9).

(20.16)

lJ

The asymptotic field is valid for small distances from the crack
It

is

important
Load

a.

to note that the C*-integral, which can be measured far from

the crack tip, at the same time unequivocally determines the


field.

tip,

crack-tip

stress

and specimen geometry affect the crack tip field only through the

C*-integral. Therefore, in

viscous

materials,

creep

crack

growth

will

be

described macroscopically by C*. This remains true even if a sufficiently small


zone near the crack tip behaves arbitrarily

differently,

say,

due

to

grain

boundary cavitation or other processes which offset the validity of a nonlinear


viscous description. Further, although eq. (20.16) was
materials,

C*

is

valid

for

derived

for

power-law

general viscous behavior. The power law is only

especially convenient in that the near-tip field can be given in closed analytic

form.

Examples involving the use of C* in creep crack growth tests will be

described at the end of Chapter 21.

21 C* -Controlled Creep Crack Growth by


Grain-Boundary Cavitation

The continuum-mechanical analysis presented so far specifies the load parameter


which determines the crack-tip fields, but it cannot predict or explain how the
crack grows in response to the applied load

parameter

(i.e.,

C*

in

viscous

materials). The understanding of crack growth rates requires a knowledge of the


micromechanisms by which the crack grows.
As an important example, the nucleation, growth
ahead

and

coalescence

of

cavities

of the main crack will now be considered (see Fig. 21.1). As a plausible

crack growth criterion it is assumed that the crack starts growing,

and

grows

at such a rate, that at a distance Xc ahead of its tip, the condition for cavity coalescence is met. Usually Xc is equal to the cavity spacing. The analysis
is

carried

out

for

different cavity growth mechanisms separately. Corrosive

effects are neglected, which is justified in many, but not

in

all,

cases.

Fig. 21.1. Creep crack growth by grain boundary cavitation.

practical

21.1

273

Crack Growth Subject to Critical Strain

The material in which the crack grows is modeled as power-law viscous. Only the
asymptotic

term

of the stress field [i.e. the HRR-field given in eq. (20.16)J

is taken into account in the analysis to follow. Then, beyond the knowledge
C*,

the

load

and

the

outer

specimen geometry need not be specified. As an

approximation, whose validity will be shown in a later section, it


that

the

cavities

is

at

is

assumed

do not modify the HRR-field. In a viscous material with no

cavitation, it is true that the stress field


crack

of

is

independent

of

whether

the

rest or whether is grows; only the current configuration defines

the stress field. Cavitation, however, does affect the stress distribution, but
this is neglected beforehand.
The formulation of the problem in terms of integral equations follows
Riedel

(1981a),

work

Bassani (1981), Bassani and Vitek (1982), Wilkinson and Vitek

(1982) and Hui and Banthia (1984). Ohji (1977) and Kubo, Ohji and Ogura
choose

of

formal

damage

parameter

description,

which

(1979)

is equivalent to the

formulation based on cavity growth laws in important limiting cases. In a later


chapter,

continuum

damage mechanics will be applied to creep crack growth and

the results will be shown to be practically equivalent with those derived next.
Other related models were quoted in reviews by Riedel (1984a, 1985a).

21.1 Creep Crack Growth Based on a Local Critical-Strain Criterion


Failure by grain boundary cavitation as described in Part II of this book often
obeys a critical-strain criterion. Within the framework of a theory which, like
the present one, neglects effects

of

cavitation

on

the

stress

and

strain

fields, it is not necessary to distinguish between critical strain and MonkmanGrant product. For constrained cavity growth and
nucleat1"on,

the

Monkman-Grant

strain-controlled

continuous

product is given by eq. (16.19). For the ratio

0eIOI we insert the corresponding ratio in the HRR-field of the crack. Then the
critical strain at which cavity coalescence occurs takes the form:
(21.1)

The numerical factor 0.4 is approximately valid for n > 3 and


n = 5,

h(w)

0.61.

If

the ratio OelOI is equal to 0.21 in a plane-strain HRR-field. Hence for

n = 5 and Aid = 1/10, the critical strain is f = 0.78%.


Now we employ the critical-strain criterion as a local criterion at

the

crack

274

21.

C*-Controlled Crack Growth by Cavitation

tip in the following sense: the macroscopic crack must grow at such a rate that
the material at a distance Xc ahead of its current tip just reaches the
cal strain

Ef

This leads to an equation of motion for the crack tip.

In order to calculate the strain ahead of a growing crack tip one


the

HRR

criti-

stress

starts

from

field, and calculates the strain rate using the material law,

eq. (20.3). The strain follows from time integration of the strain

rate.

This

time integral is evaluated at a distance Xc ahead of the arbitrary crack tip


position, a. The integral extends backwards in time over the periods when the
crack grew and when it was stationary. While the crack is stationary, the
strain field varies in proportion to x- n/ (n+1). From this, together with the
requirement that the strain must be equal to Ef at x ~ Xc for growth initiation, it follows that the contribution of the initiation period to the strain at
.
Xc ahead of the current crack tip varies as Ef [ xc/(a+xc-a o ) ]n/(n+1) '
a dlstance
where a is the current crack length, and a o is the initial crack length. The
contribution

to

strain during the growth period is an integral over the prior

.
a

crack tip posl-tions, ai, with the

time

differential

replaced

by

da'/a(a'),

is the as yet Unknown growth rate. Then the total strain ahead of the
where
crack tip is given by the expression on the left hand side of the following
equation. To satisfy the critical-strain criterion, this strain must equal

a
)n/(n+1)

Bo~(O)

ao

C*
) n/(n+1 )_.da'
InB(a+xc-a')
a(a')

This is a linear, Volterra-type integral equation for


function

of

a.

the

Ef

unknown

1/a

Ef

(21.2)

as

It starts to be valid after the crack growth initiation time,

which elapses until the critical strain is

reached

ahead

of

the

stationary

crack:
(21.3)
where 0e(O) is the value of the normalized angular function for the

equivalent

stress in the HRR-field directly ahead of the crack.


Equation (21.2) is written in a dimensionless form by introducing a
less

crack

according to

growth

increment

dimension-

A ~ (a-a )/x and a dimensionless growth rate A


o
c

(21.4)

21.1

Crack Growth Subject to Critical Strain

275

This equation already gives the dependence of


on C*. Ef Xc and B. while A
depends on A and n only. The integral equation for A follows from eq. (21.2) to
be
A

f (1+A_A,)-n/(n+1) [1/A(A')] dA'

(21 .5)

Integral equations of this type which contain a

convolution

integral

can

be

solved by the Laplace transformation method. However. in the present case the
result takes the form of an integral over the incomplete Gamma function. which
must

evaluated numerically or by series expansion (Hui and Banthia, 1984).

be

Therefore it appears easier to


standard

solve

eq.

(21.5)

directly

The

procedure to solve Volterra-type integral equations numerically is to

approximate the integral by a finite sum. This leads to


for

numerically.

1/A

recurrence

formula

which can be evaluated step by step. Results are shown in Fig. 21.2a.

while Fig. 21.2b shows the resultant crack growth increment


integration

of

the

growth

rate.

which

follows

by

Also shown is a useful two-term asymptotic

expansion for large A (dashed lines). This was

obtained

by

inserting

A A~

with

unspecified ~ into eq. (21.5). and expanding for small 1/A. In terms of
the physical coordinates a and a-a o ' the two-term expansion for large A is:

v iin(O) (BX c )1/(n+1)


e
E f sin( va)

n/(n+1)

C*

In

[(

oj
o<{

...
~
...0.
0

10

----

.,-

"2

--- ---

xc

500

IS

2!-

a-a_
_
0 )1/(n+1)

r(a)
] (21.6)
r(2-a) r(2a-1)

b)

'00

~"3OO

......
ti'

.-..--n~86

12(1)

100

~
0
c:

00

100

2(1)

300

A=(a-aoJ/xc

--

'00

500

00

20

'0

t/ti--

60

Fig. 21.2. a) Crack growth rate normalized as in eq. (21.4) vs. crack-growth
increment. Dashed lines: eq. (21.6); solid lines: full solution of
eq. (21.5). b) Crack-growth increment vs. time; ti from eq. (21.3).

21.

276

C*-Controlled Crack Growth by Cavitation

where a = n/(n+1). The expression involving the Gamma-functions has the numerical values 0.85, 0.90, 0.95 and 1.0 for n = 4, 5, 7.5 and ~, respectively. The
initial growth rate is obtained by neglecting A against 1 in eq. (21.5). There
results

Ao (n+1)/n or, in physical dimensions,

= (1+1/n)x c Iti'

In experiments, C* usually increases with crack length unless the load is reduced in order to maintain a constant C*. The predicted growth rates for three
hypothetical constant-load tests are shown schematically in Fig. 21.3. The ini-

tial growth rates, o ' lie on the dashed line which, in a log-log plot, has the
slope n/(n+1). As Figs. 21.2 and 21.3 show the initial growth rate is rather
low, typically by a factor 10 lower than later on, for the same value of C*.
Subsequently the growth rate increases due to both, the increase in C* and the
increase in A shown in Fig. 21.2a. After the crack has grown by a sufficiently
large multiple of xc' A increases only slowly, although it does not exactly
saturate [eq. (21.6)]. Then the further increase in growth rate is almost exclusively caused by the increase in C*, and
becomes an almost unique function
of C*. A comparison with experimental results in Section 21.5 will show that

the above results agree with the observed crack growth rates, whereas the
models described in Sections 21.2 to 21.4 disagree. Therefore the reader who is
interested in the most important models only can proceed to Section 21.5.
The continuum hole growth mechanism described in Chapter 14
critical-strain

also

leads

to

criterion for local failure at the crack tip, but the critical

Fig. 21.3. Crack growth rate vs. C* for constant-load test (schematic).

21.1

Crack Growth Subject to Critical Strain

277

strain is larger than for constrained diffusive cavitation. From

eqs.

(14.24)

and (14.17) one obtains a critical strain Ef = 13.4%, if the multiaxiality of a


plane-strain HRR-field for n = 5 is inserted in eq. (14.17), if the logarithm
in

that

equation

is set equal to 4 and h(w)

0.61. This is a relatively low

ductility compared to that predicted by the hole growth mechanism for


tension,

uniaxial

but it is still higher than the 0.78% strain to failure obtained from

the constrained model for the HRR-field.

21.2 Strain-Controlled Cavity Growth and Stress-Controlled


Nucleation
Now we assume that cavity nucleation occurs

instantaneously

once

critical

stress, 0nuc' is exceeded. This is not a very realistic assumption when compared to the observations reported in Chapter 5, although it agrees with the
basic

nucleation

theories described in Chapter 6. This nucleation behavior is

combined with strain-controlled cavity growth. Hence, the crack must grow
ject

to

the

sub-

criterion that the strain increment accumulated after nucleation

must reach the critical value Ef at a distance Xc ahead of the crack tip.
In the HRR-field, eq. (20.16), a given nucleation stress

is

exceeded

over

distance
(21 .7)
ahead of the crack tip. This model gives a threshold for creep crack growth:

crack does not grow if C* is so small that rnuc < xc. Otherwise, the crack
starts growing, and nucleation has no effect on the crack growth rate as long
as

the crack grows within the zone in which nucleation had occurred initially.

At the boundary of the initial nucleation zone, the


nucleation

drops

discontinuously

strain

by the amount Ef(xc/rnuc)

accumulated after
n/(n+1)
. Therefore

the crack must stop at a-a o = rnuc-xc and wait until this defect in damage is
compensated by further straining. The waiting time is t.(x Ir
)n/(n+1).
1
c nuc
During that time, the strain increment
E

(I
)n/(n+1)[(a - r
f Xc rnuc
nuc

+ 2X c )/X c

is accumulated ahead of the waiting crack. The


reaches

same

]-n/(n+1)
happens

when

the

crack

twice, or generally k times, the nucleation distance. The integral eq.

(21.5), which governs the crack growth rate, then takes the form:

21.

278

J (1+A-A,)-a
A-R

A- 1dA'

C*-Controlled Crack Growth by Cavitation

1 - (1+R )-ka (1+A-kR )-a.


n
n

(21.8)

This equation is valid piecewise for kRn < A < (k+1)R n for any positive integer
k,

while for k = 0, the lower boundary of the integral is zero. The abbreviat-

ions a = n/(n+1) and Rn = rnuc/xc - 1 were used, while A and A are


in eq. (21.4).
Again, the integral equation is solved numerically, but
difficult

to

the

author

defined

as

found

it

continue the numerical scheme stably beyond the discontinuity at

A = Rn' Therefore, at A = Rn' the procedure was changed as follows. The


integral equation (21.8) was differentiated with respect to A and the resulting
Volterra-type integral equation was treated numerically in the same way as
original

quation.

This avoids difficulties at the discontinuities, but leads

to a gradual drift
criterion,

i.e.,

the

of
the

the

solutions

strain

away

from

the

imposed

crack

growth

accumulated after nucleation is not kept at the

desired valuef' This drift can be corrected by noting that, for large A,
curves must merge into the analytic result for A ~ ~:

the

(21 .9)

.~

Rn =20

20

40

60

80

100

Fig. 21.4. Crack growth rate normalized as in eq. (21.4) vs. crack-growth
increment for stress-controlled cavity nucleation and strain-controlled growth.

21.2

279

Strain-Controlled Cavity Growth, Stress-Controlled Nucleation

which is obtained from eq. (21.8) by letting k


In this way, the results shown in Fig. 21.4
discontinuities

in

the

and setting A = const.

+ ~

were

calculated.

by

the

growth curve become small after the crack has grown a

distance equal to a few times the nucleation distance, and


given

Apparently

eq. (21.9) is approached. The conversion of

the

A to

constant

rate

physical dimensions

was given in eq. (21.4). For the steady-state growth rate this gives, from

eq.

(21.9) :

(21.10)

.
1/(n+1)
where the second line is valld for (rnuc/x c )

1.

21.3 Diffusive Growth of a Constant Number of Cavities


Unconstrained diffusive cavity growth rates are given in eq.
cavity

spacing
law

can

If

the

is constant, i.e., if nucleation occurs instantaneously at the

beginning of the test, and if the sintering stress


growth

(11.14).

be

Integration from R

integrated

o at

by

separation

is
of

neglected,
the

the

variables

cavity

R and t.

nucleation to R = A/2 at coalescence gives


(21.11)

Obviously, failure by unconstrained diffusive cavitation occurs when


integral

on

the

maximum

principal

stress

attains

the

right-hand side of eq. (21.11). Creep crack growth is now treated by


that,

at

distance

Xc

the

time

value given by the


requiring

ahead of the crack tip, this critical value must be

maintained. The time integral

on

stress

is

calculated

in

analogy

to

the

integral on strain rate in the preceding section. This leads to an integral


equation having the same structure as eq. (21.5) except that n/(n+1) is
replaced by 1/(n+1). As an asymptotic solution, the crack growth rate for large
growth increments, (a-a )/x 1, is found to be
o
c

280

21.

11

315 l'!oD b

01(0)

sin(1I/(n+1) 2kTA 3h(1jI)

C*-Controlled Crack Growth by Cavitation

C* )1/(n+1) (a-a )n/(n+1).

I B
n

(21.12)

The maximum tensile stress component in an HRR-field ahead of the crack, 1(0) ,
is oriented normal to the plane of the crack. A discussion of the result
follows in Section 21.5.

21.4 Diffusive Cavity Growth and Stress-Controlled Nucleation


Again the sintering stress in diffusive cavity growth is

neglected.

Then

the

criterion for the growth of the main crack is that the time integral on 01 must
attain the critical value given in eq. (21.11). Since now stress-controlled
nucleation is taken into account, the integration on 01 must start only when
the principal stress exceeds the nucleation stress. In close analogy to Section
21.2,

one obtains the same type of integral equation as eq. (21.8), but with a

having the meaning a


a

= 1/(n+1),

and

Abeing

non-dimensionalized differently

in

way not shown here. Figure 21.5 shows a numerical solution for n = 5 and for

a nuc:eation distance Rn = 20. It is apparent that the discontinuities at


A = kR n are stronger than for strain-controlled cavity growth. After growing by

80
60

t
.-::(

Rn=20

40

20

100 200 300 400


A =(a-aa)/xc ~

Fig. 21.5. Normalized crack growth rate vs. crack-growth increment


for stress-controlled cavity nucleation and diffusion-controlled growth.

21.3

Diffusive Growth of a Constant Number of Cavities

about 10 times the nucleation distance, the

281

crack

approaches

h(~)

n
0nuc
In B]

steady-state

growth rate of
a = 158 (1+1/n) -n+1
01
06D b C*I [
kT3
A
This result will be discussed in
neglected compared to Rn/ (n+1).

Section

21.5.

In

eq.

(21.13),

(21.13)
unity

was

21.5 Comparison with Experiments


As was pOinted out in Chapter 20, C* is expected to be a valid parameter for
the characterization of crack growth if the material deforms viscously. Experimentally, this means that the load line displacement must be caused primarily
by
viscous
creep, while the contributions by elastic deformation and
instantaneous plasticity must be much smaller. The criteria for the validity of
C* will be detailed in later chapters. Many tests on ductile alloys reported in
the literature were done under conditions where these criteria are fullfilled.
(A few references were given in Sections 19.2 and 19.4). Considering the
experimental results one notices a remarkable similarity between different
materials over a wide range of temperatures as far as the creep crack growth
rate as a function of C* is concerned. In the following, one representative
example will be described in greater detail.
21.5.1

Tests on a 1Cr-1/2Mo steel

Riedel and Wagner (1985) have measured creep crack growth rates in a 1Cr-1/2Mo
steel in the temperature range 450 0 C to 600 oc. Some of the tests were performed
en as-processed steel which had a fine, bainitic microstructure. However, the
majority

of

the

tests

were done on material which had been in service (as a

pipe in a power station) for 103,000 h at nominally 530 0 C under a stress of 53


MPa. During that time, the carbides had coarsened and the creep resistance had
dropped accordingly. Only the results on the used material will be reported
here. The crack growth rates in new material are very similar and will be given
in Section 23.4 where they are to illustrate the appropriateness of C* for this
material. Uniaxial creep data for the material having a prior service history
were measured and fitted to Norton's creep law, E = B~. At 535 0 C, which was
the usual testing temperature, this gave B = 5.6.10-26 MPa-n/s and n = 8.6.

282

21 .

C*-Controlled Crack Growth by Cavitation

So far, only CT-specimens were tested which had thicknesses

of

12.5,

25

and

50 rom. Most of them had 25% side grooves as recommended by deLorenzi and Shih
(1983) in order to approximate a plane-strain field along the whole crack front
(cf.

Section

22.2).

Indeed,

the crack front remained nearly straight during

fatigue pre-cracking at room temperature and during creep crack


indicates

that

the

stress

growth,

which

field is uniform in the direction parallel to the

crack front .
During the tests, the crack length was continuously monitored using
tric

potential

drop

technique

the

elec-

and, sometimes, the elastic compliance during

small unloading/reloading events. Both are well established methods in fracture


mechanics,

and

they

were

described

recently

by Hollstein, Blauel and Voss

(1984) and by Hollstein (1985). Additionally, markings on the fracture


caused

surface

by load changes or changes of the atmosphere were used to calibrate the

continuous methods .
Some creep

cra~k

growth tests were interrupted and the specimens were

examined

for grain boundary damage. There is clear evidence of cavitation near the crack
tip, as the scanning electron micrograph in Fig. 21.6 shows. The main crack
appears to grow by interlinkage with cavitated grain boundary facets.
In order to see whether environmental effects playa dominant role, some of the

Fig. 21.6. Scanning electron micrograph of etched and polished section

showing cavitation in crack-tip region in a CrMo steel.


(From Riedel and Wagner, 1985).

21.5

Comparison with Experiments

283

tests were performed in a mixture of argon and 3% hydrogen.


creep crack growth rate could be detected

No

difference

in

compared to tests in air. In agree-

ment with other workers, it is concluded that creep crack

growth

in

ferritic

steel at 535 0 C is not primarily controlled by the environment.


Figure 21.7 shows measured crack growth rates as a function of C*; C*

was

de-

termined using eq. (20.13). [Application of eq. (20.7) gave the expected deviations due to the uncertainty as to plane strain or plane

stressJ.

If

plotted

against C* the crack growth rates exhibit a very small temperature dependence.
On a log-log plot of
versus C*, a slope of n/(n+1) fits the data. In Fig.

21.7,

data pOints from the early stages of the tests were omitted. These early

growth rates are generally smaller than later on, for the same value of C*,

as

was indicated schematically in Fig. 21.3. It should be mentioned, however, that


this initial behavior may
pre-cracked

occupy

large

fraction

of

the

lifetime

of

specimen since the crack grows slowly in this range. Thus the data

omitted from Fig. 21.7 for clarIty's

sake

are

very

important

for

lifetime

estimates. OtherwiSe the lifetime predictions are overly conservative.


21.5.2

Comparison with models

A consistent explanation of the data shown in FIg. 21.7 is given by

the

model

which assumes that local failure of the material at the crack tip is strain
controlled (SectIon 21.1). First. the predicted dependence
~ C*n/(n+1) is

."

'!>2

Cr /VIa steel

'"Q
~\':

E!>2

.~

0 ~

" 450C
+ 535C
o 600C

!>2

~
51

'!>2

"

"

1(;2 10' 10 10' 102 103


ClI inWlm2

Fig. 21.7. Crack growth rates in 1Cr-1/2 Mo steel. Solid line: eq. (21.6).
(From Riedel and Wagner. 1985).

21.

284

C*-Controlled Crack Growth by Cavitation

confirmed experimentally. Second, eqs. (21.4) and (21.6) explain why the temperature

dependence

is

weak:

the

strongly

temperature-dependent material

parameter B from Norton's creep law is raised to the small power 1/(n+1). This
reduces the apparent activation energy of the growth rate (n+1)-fold compared
to that of B. The other parameters in eq. (21.6) are not strongly

temperature-

dependent nor do they vary greatly from material to material. This explains why
different kinds of materials exhibit similar creep crack growth rates. Thirdly,
the

data

omitted

from

Fig.

21.7 show a dependence of

a on

the crack growth

increment, a-a o ' besides that on C*, which is qualitatively comparable with the
dependence predicted byeq. (21.6) (see Fig. 21.3). Finally, the absolute values of
are also predicted well. To show this, insert B 5.6.10-26 MPa-n/s,

n = 8.6, 0e(O) = 0.63, In = 4.63 (from Shih, 1983), a-a o


1.5 mm and Xc = 2jJDl
(assumed, but the result is insensitive to that choice) into eq. (21.6), and
consider

Ef
as an adjustable parameter. The solid line which fits the data in
Fig. 21.7 is obtained if E f = 0.56%. This is in excellent agreement with the
estimate, E f = 0.78%, given in Section 21.1, which was based on creep-constrained cavi-tation in the HRR crack-tip field.

The data can also be fitted by eq. (21.10) which is based on

nucleation
a

is

and

strain-controlled

growth

of

C* is also compatible with the data, and no strong


predicted.

Using

the

stress-controlled

cavities: the linear dependence


temperature

dependence

same numerical values as in the preceding paragraph

together with 01(0) = 2.44 and

= 0.66% and fitting eq. (21.10) with the data


21.7 leads to a nucleation stress of nuc = 450 MPa. According to eq.

in Fig.
(21.7) the

nucleation

rnuc = 680 ~m if C* =

Ef

distance

increases

in

proportion

to

C*

and

is

1 W/m 2 Neither of the numerical values obtained for 0nuc

and rnuc appears unreasonable.

The model based on diffusive growth of a constant number of cavities Ceq


(21.12)] obviously gives an incorrect dependence of a on C*, and a temperature
dependence which is too strong. Absolute values for

a are

too high,

since

the

constraint on growth and continuous cavity nucleation are neglected.


Diffusive growth combined with stress-controlled nucleation leads to

linear

relation
~ C*, and to a weak temperature dependence,
~ 6Db/B Ceq. (21.13)],
which is compatible with the experimental results. However, for eq. (21.13) to
fit the data the nucleation stress must be chosen as 0nuc

150 MPa which, from

eq. (21.7), implies a nucleation distance of rnuc = 15 m if C* = 1 W/m 2 This


means that nucleation should have occurred easily in the whole specimen and eq.

21.5

Comparison with Experiments

285

(21.12) rather than eq. (21.13) should be applied. But eq. (21.12) was

already

found to disagree with the observed behavior.


21.5.3

Conclusions

In conclusion, the observed crack growth rates can be explained consistently if


local

failure

at the crack tip is assumed to be strain controlled. The agree-

ment is still good if local failure is brought about by stress-controlled cavity

nucleation

and

strain-controlled growth. Among the models examined, those

based on diffusion-controlled cavity growth disagree with


crack

growth

behavior.

The

case

of

that

such

model,

if

it

observed

creep

diffusive growth and strain-controlled

nucleation could not be treated because of


possible

the

mathematical

difficulties.

It

is

were available, could also explain the

measured data.
It should be kept in mind that the whole discussion so far has
purely

been

based

on

viscous deformation behavior. The effect of elastic transients on crack

growth rates will be described in Chapter 23. These transient effects are not
negligible in materials which are not too ductile. Further, the subject of
crack growth modeling will be taken up again in Chapter 27 in the framework
continuum damage mechanics.

of

22 Specimen Size Requirements for C* -Testing


Caused by Crack-TIp Blunting and by 3-D Effects

In fracture mechanics, test specimens must generally

satisfy

certain

minimum

size requirements. In linear elastic fracture mechanics, it is the plastic zone


compared to which the specimen must be sufficiently large. This criterion has
led to the well known ASTM-E 399 rule for linear elastic fracture testing.
Under fully plastic conditions, the

crack-tip

opening

displacement

must

be

small compared to the specimen dimensions, which is expressed by the ASTM-E 813
rule. Crack-tip blunting sets an analogous limitation to C* as will
Further,

the

be

shown.

crack growth behavior is often to be measured under plane-strain

conditions, which requires sufficiently thick specimens. This is so in rate-independent fracture mechanics as well as in creep crack growth testing. The discussion of this problem in Section 22.2 is applicable to both cases.

In

later

chapters, further limitations to C*, which arise from the constitutive behavior
of the material, will be described.

22.1 Limitations to C* Set by Blunting


Before a crack starts growing, for example by coalescence with
cavities,

grain

boundary

its tip is blunted by creep flow of the surrounding material. In the

derivation of the HRR-field, on the other hand, the crack is treated


mathematically

as

being

snarp. Therefore the validity of the HRR-field is restricted to

distances from the crack tip which are large compared to the crack-tip

opening

displacement, 6t This problem has been examined in detail for time-independent


plasticity by Rice and Johnson (1970) and by McMeeking (1977). They include
crack-tip blunting in their analyses and find that blurtting disturbs the stress
field over distances 3 to 5 times 6t , whereas outside that zone the analysis
which neglects blunting becomes increasingly accurate. Thus the range of validity of the HRR-field is limited by blunting towards small
crack

tip,

and

by

distances

from

the

the outer specimen geometry towards large distances. This

means that the specimen must be large enough compared to 6t in order

to

allow

22.1

287

Limitations to C* by Blunting

for a finite range of validity of the HRR-field. This is expressed by:


(a, W-a) > 2 MOt.

(22.1 )

The factor M, which specifies how much larger the specimen


crack-tip

opening

displacement,

depends

on

must

be

than

the

the desired accuracy and on the

specimen geometry. Numerical studies, like that of McMeeking and

Parks

(1979)

show that the singular HRR-field has a reasonable range of approximate validity
if M

25 in compact specimens, whereas for

center-cracked

plates

in

plane-

strain tension, M = 200 is required. It is important to realize that the existence of a unique asymptotic field is essential for C*-testing since it
tees

guaran-

a unique behavior of crack tips in differently shaped specimens. If there

is no range of validity for the asymptotic field, the conditions at


tip

the

crack

depend on details of the specimen geometry, and different specimens cannot

be compared on the basis of C*.


In rate-independent plasticity, the crack-tip opening displacement is given
0t

= J/(2a ),

by

where ay is yield stress and J is the J-integral. With this value

of 0t' eq. (22.1) represents the ASTM-E 813 requirement for valid J-testing.

The crack-tip opening displacement in nonlinear viscous materials is calculated


next.

Following

Knauf and Riedel (1980), we define 0t as the distance between

the crack faces, where two lines drawn through the apex of
inclined

by

the

crack

profile

30 0 to the crack plane intersect the crack profile (Fig. 22.1).

Displacement rates associated with an HRR-field have the form


(22.2)
where the dimensionless functions Ui(S) are tabulated by Shih (1983).
as

the

crack

As

long

is stationary and C* is constant, the displacement is u i = uit.


u i at S = ~ and r = us/tan 30 o gives the

Taking twice the S-component of

crack-tip opening displacement of a stationary crack:


(22.3)
The term in brackets is equal to 1.25 and 1.11
for

plane

strain.

Figure

for n

5 and 7,

respectively,

22.1 also shows the evolution of the crack profile

when the crack grows. The profiles were calculated by integrating the displacement rate, eq. (22.2),

for a crack growth

history as described in eqs. (21.2)

288

22.

Blunting and 3-D EffectE

Fig. 22.1. Definition of the crack-tip opening displacement of a stationary

crack, and evolution of the crack profile during crack growth.


Test on CT-specimen of 21/~Cr-1Mo steel (magnification 4x; Detampel, 1986).

to (21.6) taking n = 7 and f = 1%. Since the calculation involved analytic


approximations, the results are not numerically accurate, but they exhibit the
correct features. After the crack has gro.wn for some multiple of the initiation
time, t i , which was given in eq. (21.3), the crack tip re-sharpens, which is
also observed experimentally. It is unknown how the
growing

crack

wedge-type

shape

of

the

affects the validity of C*. Assuming that only blunting matters

rather than growth, the effect of blunting is greatest after a

few

times

ti'

say at t = 3t i From eqs. (22.3) and (21.3), the crack-tip opening displacement
at t = 3ti for n = 7 is found to be

'\

1600

(n+1) In
f
xc

(22.4)

For f = 1%, Xc = 3~m and n = 7, eq. (22.4) yields 0t = 25~m. Typical specimen
dimensions are 'sufficiently large compared to 25~m to satisfy eq. (22.1). In
conclusion, crack-tip blunting generally implies no severe

limitation

to

C*,

but the related effect of wegde-like growth may restrict the use of C* to small
crack growth increments. The latter effect is hard to quantify at present.

22.2 The Third Dimension in Fracture Mechanics and its Practical


Consequences
So far only two-dimensional problems have been considered. Plane-strain conditions

are

approached in very thick speCimens, while plane stress requires very

thin specimens. In reality, specimens have a finite thickness, and it is impor-

22.2

289

The Third Dimension

tant to know under which conditions the idealized two-dimensional limits are
good approximations. In the following, a few facts are listed which the author
found useful in relation to fracture mechanics testing both for rate-independent materials and for creeping materials.
22.2.1

The C*-integral in three dimensions

In nonlinear viscous materials, the surface integral in

three-dimensional

de-

formation fields,
(22.5)
vanishes for all closed surfaces S bounding regions which contain no cracks

or

holes (see Budiansky and Rice, 1973).


In order to utilize eq. (22.5) for the usual fracture mechanics specimens, consider

a surface as shown in Fig. 22.2. The parts of the surface along the side

surfaces and along the crack contribute nothing to the integral

since

n1 = 0,

and aijn j = 0 on the crack due to the traction-free boundary conditions. Since
the whole surface integral must vanish, the integrals over the partial surfaces
Sl and S2 must be of equal magnitude and opposite sign. This means that the 3-D
C*-integral defined on a partial surface Sl is independent of the choice of
that surface. In other words, the thickness average of the 2-D C*-integral
evaluated on a cylindrical surface is path-independent. It is
integral

related

to

the

on the load/displacement rate curve by the usual formula, eq. (20.2).

Thus, irrespective of how complicated the 3-D stress fields are,

the

C*-value

measured at the load pOints gives the average value of C* at the crack front.

Fig. 22.2. Application of the three-dimensional C*-integral.

22.

290

22.2.2

Blunting and 3-D Effects

Asymptotic crack-tip fields in specimens of finite thickness

To start the discussion of the 3-D fields it is first shown that the asymptotic
field

near a crack tip is a plane-strain field (crack-tip bluntirtg being negl-

ected beforehand). In power-law viscous materials, especially, the plane-strain


HRR-field

is

the

asymptotic

field

near a crack tip. This is so because the

strain component parallel to the crack front, e 33 , cannot become singular except directly at the point where the crack front intersects the side surface of
the specimen. If e 33 had a singularity in r over a finite length in x3-direction, the displacement u 3 = fe 33 dx 3 would become infinite, which is inconsistent with the compatibility condition. Therefore, the equations for the leading,
asymptotic

terms

of

the

stress

and strain fields are subject to the plane-

strain condition. Hence, the resulting asymptotic field must be that for

plane

strain. Its range of validity will be discussed in Section 22.2.4.


22.2.3

The singularity at the intersection of the crack front with the surface

Figure 22.3 shows a plan view on the crack plane and on the ligament ahead of
the crack. The crack front intersects the free surface at a right angle. The
intersection point is called a vertex. At the vertex, a new type of singularity
develops

which

is

favorably analyzed in spherical co-ordinates

that the definitions of


for

e and

(p,e,~).

Note

are inverted here compared to the common usage,

compatibility with the definition of

e in this book. In these coordinates,

the vertex field can be factorized according to


(22.6)

un cracked
ligament
specimen
surface

vertex

cracktront

o
crack
plane

Fig. 22.3. Spherical co-ordinates around a vertex. The angle e (not shown)
pOints out of the plane being e = 0 on the ligament and e = n on the crack.

22.2

The Third Dimension

Substituting this into the


materials
function

leads
Fij(6,~)

291

three-dimensional

field

equations

for

power-law

to a two-dimensional eigenvalue problem for the dimensionless


and the eigenvalue, s.

This eigenvalue problem has been solved only for linear elastic (or, by
gy,

and by Bazant and Estenssoro (1977) employing different


They

find

that

numerical

techniques.

0.5, 0.452 and 0.332 if Poisson's ratio is v

0, 0.3 and

0.5, respectively. This means that the vertex singularity is


usual

analo-

linear viscous) material by Benthem (1977, 1980), Benthem and Douma (1980)

weaker

than

the

inverse square root singularity except when v = 0; if v = 0, the surface

plays no particular role. For small

~,

i.e., close

to

the

crack

front,

the

vertex field approaches the plane-strain crack-tip field as one expects:


as

~ ~

0,

(22.7)

where f ij (6) describes the angular stress distribution in linear elastic cracktip fields.
The factor kI in eq. (22.6) is undetermined by the asymptotic analysis of the
vertex field. Its determination would require a full 3-D finite element
analysis of the Whole specimen. However, from the scaling

laws

for

power-law

materials (which include linear elastic materials), it is clear that kI must be


proportional to
(22.8)
where Bt is the total specimen thickness (as distinct from the net thickness in
side-grooved specimens described later).
As a consequence of eqs. (22.6) and (22.7), the coefficient of the
tip

1//r-crack-

Singularity, which we call the local stress intensity factor, goes to zero

at the free surface according to


(22.9)
This behavior is shown in Fig. 22.4
normalized by its thickness average,

for the local J-integral, J


K2I ,
J. In viscous material, J is replaced by

C*, which is also defined locally here as the strength of the singularity.

292

22.

Blunting and 3-D Effects

,
CT-specimen
o~~--~~--~~

o a1 a2 a3

I
surface

a~

x3/ B, - -

as

center

Fig. 22._. J-integral normalized by its average, J, vs. dlstanc~ from


free surface, x 3 Solid lines: Finite element results (P L = Paa~tic
limit load). Dashed line: Asymptotic vertex solution, J ex x 3 "

For comparison, the result of a 3-D finite element calculation by deLorenzi and
Shih

(1983) is also shown in Fig. 22.4. The numerical calculation was done for

a standard ASTM compact specimen. At the free surface, the numerical result

is

compatible with the required asymptotic behavior. Also shown is the numerically
calculated distribution of J for elastic strain-hardening plastic material at a
load

which

is

25%

above the plastic limit load. Obviously, J falls off more

distinctly towards the free surface if the specimen is fully plastic. A remedy
for

the

undesirable

non-uniform

distribution

of

J is Side-grooving of the

specimens as will be shown in Section 22.2.8.


22.2.4

Ranges of validity of singular fields in parallel-sided specimens with


straight crack fronts

Figure 22.5 qualitatively shows the ranges


fields

of

validity

of

various

singular

on the ligament of a cracked, parallel-sided specimen. At the vertices,

Benthem's (1977) solution, or its as yet unknown nonlinear analogue, is


For small angles

~,

valid.

this field converges to the plane-strain HRR-field. Further

away from the crack front the in-plane stress gradients become

smaller.

Since

all stress gradients are related through the equilibrium equation, the throughthe-thickness gradients must also become small. Because of thiS, in combination
with

the

traction-free

boundary

conditions

plane-stress conditions are approached.

on

the side surfaces, 03i

= 0,

22.2

293

The Third Dimension

for field on
ligament

era

Fig. 22.5. Ranges of validity of singular fields on the ligament ahead of


a crack (schematic_for a thin specimen, Bt a); p-o = plane stress,
P-g = plane strain, r s = vertex field; hatched area affected by blunting.

In relatively thin specimens, the range of the

plane-strain

HRR-field

scales

with the specimen thickness, Bt , and the outer limitation to the plane-stress
HRR-field is some fraction of the crack length or ligament width. For thick
specimens,

on

the

other

plane-stress HRR-field,
fraction

of

the

hand,

and

the

there is no finite range of validity for the


plane-strain

HRR-field

is

valid

to

some

crack length or ligament width. Towards small distances from

the crack tip, the validity of the asymptotic fields is cut

off

by

crack-tip

blunting as discussed in Section 22.1.


The fact that different singular fields occupy different regions ahead
crack

front

validity.

This

problem

will

plane-stress

specimen

finite

range

be examined below in greater detail. On the

other hand, the macroscopic response


idealized

the

has important practical consequences. Plane-strain fracture test-

ing, for example, requires that the plane-strain HRR-field has a


of

of

if

of
the

specimen

approaches

that

of

an

plane-stress HRR-field has a finite

range of validity. The macroscopic response will be discussed in Section 22.2.9


together with that of side-grooved specimens.
22.2.5

Conditions for plane strain near the crack tip

It is often desirable to investigate crack growth


strain

near

under

conditions

plane

the crack tip since this represents a well-defined limiting case.

It should be emphasized that the existence of a plane-strain HRR-field


finite

of

over

range is independent of whether the overall specimen response is closer

to plane stress or plane strain. These two questions should not be confused.

294

22.

Blunting and 3-D Effects

For practical purposes the range of the plane-strain HRR-field should be


numerically

rather

than

only

known

in terms of the proportionalities given in the

preceding section. Unfortunately, three-dimensional finite element calculations


have

rarely

been evaluated systematically from this point of view. Therefore,

two analytic estimates for linear elastic material will be described next.
Yang and Freund (1985) perform an approximate analysis based on
assumption

that

the

the

kinematic

displacement field in a thin parallel-sided specimen has

the special form


(22.10)
where the strain components are independent of x3 The Greek index a has the
range 1,2. Of course, the analysis is also valid for linear viscous materials,
if displacement and strain are replaced by their time rates.
Two-dimensional field equations are obtained by substituting eq.
the

original

(22.10)

into

three-dimensional equations and then integrating with respect to

x3 over the specimen thickness. Yang and Freund (1985) solve the

two-dimensio-

nal equations for a large, cracked plate of thickness Bt using a boundary layer
approach. This means that the crack is considered as being of semi-infinite
length

and

the fields far from the crack tip must asymptotically approach the

elastic singular field for plane stress. After lengthy calculations,


Freund

Yang

and

arrive at' the following result for the lateral contraction of the plate

near the crack tip:


(22.11)
The leading term Rear the crack tip is constant,
agreement

with

our

i.e.

independent

of

r,

in

previous conclusion that E33 must not become singular. In

order to assess the relevance

for

the

validity

of

the

plane-strain

field

consider the lateral stress which one obtains from Hooke's law:
(22.12)
have an inverse
While E33 is nonsingular, the stress components all and
square root singularity and therefore dominate at the crack tip. Now we define,
arbitrarily, that the range of approximate validity of the
ends

where

the

nonsingular

term,

EE 33 ,

reaches

plane-strain

field

30% of the singular term,

22.2

The Third Dimension

295

v(011+ 022) Taking E33 from eq. (22.11) and 011 = 022
the outer range of validity of the plane-strain field:
(22.13)

Towards the crack tip, the range of validity of the plane-strain field ends

at

about 5 6t due to blunting. Thus the plane-strain HRR-field has a nonzero range
of validity if the specimen thickness exceeds about 40 6t [taking v = 0.5 in
eq. (22.13)]. For rate-independent plastic materials is 6t = 0.5 J/oy , and
hence the thickness condition becomes: Bt > 20 J/oy This is
ASTM-E 813 rule, which recommends a factor 25 instead of 20.

close

to

the

For creep cracks, 6t was found to be small. Hence, the thickness condition for
plane-strain crack-tip fields is often easily fullfilled as far as blunting is
the limiting factor. Wedge-like crack growth as shown in

Fig.

might

22.1

be

more critical, and so is the occurrence of a process zone as will be shown.


An alternative approach to study the range

of

validity

of

the

plane-strain

asymptotic field is to inspect the stress fields at the vertex given by Benthem
and Douma (1980). Their numerical solutions for the vertex field show that 033
on the ligament has dropped to 60% of the plane-strain value if ~ = 14 0 with ~
as defined in Figs. 22.3
approximate

plane-strain

or

22.5.

conditions

This
are

small

value

of

indicates

that

confined to a small sector near the

crack front. If we require that this sector extends 56 t ahead of the crack
front over at last 70% of the specimen thickness, the specimen thickness must
be Bt > 133 6t This is by a factor 3 more stringent than the criterion derived
from the Yang and Freund solution.
In both cases, the definitions of what is considered a sufficient approximation
to plane strain were somewhat arbitrary. In practice, the acceptable deviations
from plane strain will also depend on the sensitivity of the

fracture

process

to

triaxility. Thus, the factor between Bt and 6t necessary to approximate


plane-strain conditions must eventually be verified experimentally.

22.2; 6

Thumbnail-shaped crack fronts

In Section 22.2.3 it was shown that the local stress intenSity at a

crack

tip

decreases when a free surface is approached. As a consequence, the crack starts


growing

in

the

center

of

the specimen which leads to a so-called thumbnail

296

22.

Blunting and 3-D Effects

c)

aJ

flat precrack

crack front

Fig. 22.6. a) Moderate and strong tunneling of a crack front.


b) Shear lips. c) Side grooves.

shape of the crack front which is also called crack tunneling (Fig. 22.6a). The
degree

of tunneling can he estimated from an analysis by Bazant and Estenssoro

(1977) They mention unpublished work

of

theirs

which

indicates

that.

for

elastic material and v = 0.3. an inverse square root stress singularity


is re-established at the vertex i f the crack front deviates by ~ = 11 0 from the

linear

normal on the surface. Then the degree of the singularity is the same along the
whole crack front. which is a prerequisite for a uniform stress
fact,

3-D

numerical

calculations

of

Smith.

intensity.

In

Towers and Smith (1984) give a

nearly constant value of KI along the crack front if it intersects the surface
at an angle between ~ = 90 and 110. In agreement with these predictions.
fatigue cracks grown under linear elastic conditions in fracture mechanics test
specimens often exhibit surface angles between ~ = 100 and ~ = 15 0 .
For nonlinear materials. vertex fields have not been

investigated.

experiments

for

and

3-D

finite

element

calculations

curved

but

both.

crack fronts

(Kikuchi and Miyamoto, 1984a), indicate that crack tunneling must be much
pronounced

than

in

linear

elastic

value of local J or C* along the crack


numerical

result

shown

in

Fig. 22.4

more

material in order to maintain a constant


front.
that

This
the

is

compatible

with

the

non-uniformity of J is much

greater for nonlinear than for linear material if the crack front is straight.
22.2.7

Shear lips

Even if the front of a growing crack adjusts itself such that constant
constant

J-.

or

C*-. conditions prevail along the crack front. there is a gradient of

stresses. in particular of 033' towards the surface at any finite distance from

22.2

The Third Dimension

297

the crack tip. The fracture process in many materials is sensitive to this loss
of triaxiality. As a consequence, the fracture mode may change
separation

of

the

specimen

halves

in

the

the

formation

of

shear

lips

normal

center of the specimen to shear

fracture along shear lips at the surface (Fig. 22.6b). In


testing,

from

creep

crack

growth

is less predominant according to the

author's experience than it is in ductile materials at room temperature.


22.2.8

Crack-tip fields in side-grooved specimens

An effective means
side-grooving

to

suppress

both,

crack

tunneling

and

shear

lips

is

of the specimens as shown in Fig. 22.6c. Side grooves counteract

the tendency of the stress intensity to falloff towards the free surface. By a
suitable choice of the included side groove angle, 28, the degree of the vertex
singularity, s, can be adjusted to that within the material. Further, the depth
of

the

side

grooves can be varied in order to optimize the smoothness of the

stress intensity along the crack front.


The vertex singularity at sharp side grooves has not yet been analyzed, neither
for nonlinear nor for linear material. However, 3-D finite element calculations
have been performed for side-grooved compact specimens by Shih,
Andrews

(1977)

for

linear

material

and

by

deLorenzi

constant

at 2S

grooves

in

terms

of

kept

out

a to

to

be

a uniform distribution of the local J-integral and in

terms of approximate plane-strain


Figure

was

while the depth of the side grooves was varied from

50% of the half specimen thickness. A relative depth of 25% turned


favorable

and

and Shih (1983) for

elastic/plastic material. The notch angle of the sharp side


= 45 0 ,

deLorenzi

conditions

along

the

whole

crack

front.

22.7 shows that J and the transverse stress component 033 at a distance

about 1% of the crack length ahead of the crack front, normalized by 011 + 22 ,
vary to a much lesser extent in side-grooved than in parallel-sided specimens.
Accordingly, side-grooved specimens exhibit
This

is

practically

no

crack

tunneling.

demonstrated in Fig. 22.8, which shows fractured compact specimens of

Nimonic BOA. The dark part of the fracture surface is due to creep crack growth
at 650 0 C, while the remaining ligament was fractured at room temperature giving
a

bright

materials,

fracture

appearance.

Sometimes,

particularly

in

more

ductile

the crack advances faster at the side grooves which suggests to use

~ slightly less acute groove angle than 45 0

Kikuchi

and

Miyamoto

draw the same conclusion from their 3-D finite element analysis.

(1984a,b)

22.

298

a)

><:'"

"""l

a6

/ / i?
/

elastic

" "-kelastic-plastic

'~
.......
'-'-

00

" I P=l25 If.


I
I
I

01 a2 03 at: 05
.vBt

a5

b) 1

..

P=125 If

1-

--------

elastic-plastic

--. at:

'"
b'

I
03 /~_l
I

:::;:: 02 r
J
0-

b~

Blunting and 3-D Effects

elastic,

v=03

al

aa

01

--

Q2 03 at: a5
x3 /Bt

Fig. 22.7. J-integral (a) and plane-strain constraint (b) vs. distance
from specimen surface for 25% side-grooved (solid lines) and parallel-sided
(dashed lines) CT-specimens. (From deLorenzi and Shih, 1983).

22.2.9

The compliance and C* in parallel-sided and side-grooved specimens

The macroscopic load/displacement rate response of a specimen is interesting in


relation to the calculation of C* from the formulas shown in Section 20.2.1.
For linear elastic materials, Shih, deLorenzi and Andrews

(1977)

compare

3-D

finite element calculations with measured compliances. The agreement is excellent.

Thei~

results for the normalized compliance, E

Bt/P (where E is Young's

Fig. 22.8. Creep crack growth in side-grooved and parallel-sided specimens.

22.2

299

The Third Dimension

modulus and A is load line deflection), of compact specimens with


side

grooves

22.1. Later results by deLorenzi and Shih (1983) are


line.

and

without

are compared with plane-stress and plane-strain results in Table

for

included

in

the

bottom

stress are taken from Saxena and Hudak (1978), while


plane-strain values are obtained by multiplying the latter values by (1-v2 )
with

Values

plane

= 0.3. As the table shows, the compliance of a parallel-sided standard

compact specimen is somewhat closer to plane stress than to plane strain.


grooving

raises

the compliance to slightly above the plane-stress value. (Re-

call that the compliance is normalized by the total specimen


Thus

in

the

Side

elastic

thickness

here).

range, 3-D effects in standard specimens are relatively

small. Compliances are within a few per cent of the 2-D

limits.

The same is
2
true for the J-integral, which also differs by only a factor 1-v between plane

strain and plane stress in linear material.


In nonlinear material, C*
C* = 0net ~ g2'

if

the

can

be

determined

displacement

from

eq.

(20.13),

i.e.

insensitive to the choice of plane-strain or plane-stress values for g2. It


applicable

to

from

rate has been measured. This formula is


is

both, parallel-sided specimens, which fall between plane strain

and plane stress, and to side-grooved speCimens, for which deLorenzi and Shih's
(1983) finite element calculations suggest very good accuracy of eq. (20.13) if
plane-strain values for g2 are used. Kikuchi's and Miyamoto's (1984b)
tions lead to the same conclusion.

calcula-

If, however, the displacement rate is not known (e.g., before the test has been

n+1 gl' must be used, which gives


carr i e d ou t) ,eq. (20 .7), i .e. C* -- a B 0net
C*-values differing by typically a factor 10 between plane strain and plane
stress.

For

parallel-sided specimens, the general trend of 3-D finite element

calculations indicates that the response of typical test specimens is closer to


plane stress than to plane strain. For side-grooved specimens, the author tried
to evaluate results of the 3-D calculations of deLorenzi and Shih (1983) compa-

plane
stress
a/W = 0.5
a/W = 0.6
a/W = 0.6

plane
strain

3-D, no
side grooves

37.0
33.7
63.4
57.7
(from deLorenzi and Shih, 1983)

36.3
60.9
62.1

3-D, 25%
side grooves
38.2
63.6
63.3

Table 22.1. Normalized compliances, EASt/P, of compact specimens with


25% side grooves and without side grooves (standard ASTM dimensions).

22.

300

Blunting and 3-D Effects

red to the 2-D results from the EPRI-Handbook of Kumar et al


sults,

which

are

(1981).

The

re-

not quite conclusive, are shown in Fig. 22.9. DeLorenzi and

Shih give the load/displacement response

of

side-grooved

and

parallel-sided

CT-specimens of elastic-plastic material with a hardening exponent n = 10. Only


the fully plastic limit, i.e. the limit of large loads, is
For

for plane strain (p-E) and plane stress (p-o) and for the
used

of

interest

here.

this case, the load-line displacement was calculated also from eq. (20.11)
material

parameters

by deLorenzi and Shih. The applied load per unit thickness was either re-

ferred to the net specimen thickness between the side grooves, Bnet , or to the
total thickness, Bt , or to an effective thickness defined by Shih, deLorenzi
and Andrews (1977) as
(22.14)
As Fig. 22.9 indicates, the
approximated

3-D

calculation

for

Side-grooved

specimens

best by plane stress using the effective thickness. This 2-D cal-

culation underestimates the displacement by a factor of about 1.5. The


with

is

problem

Fig. 22.9 is that the 3-D result for no side grooves gives a greater dis-

placement than the plane-stress calculation using Bt , which is the only reasonable thickness for parallel-sided specimens. This disagrees with intuitior
and with other results from the literature which indicate that
configuration

is

plane-stress

more compliant than the corresponding specimen having finite

thickness. Further work is needed in this area.

2000

1000

p-I1,Bnet

700

no side grooves

500
400

25% side grooves


(from de Lorenzi
and Shih)

300
200
CT 5-specimen

100 L--L_---'L-_ _--'_ _ _- '


QI

mo

load-line displacement, Lf in mm

Fig. 22.9. Comparison of 2-D (straight lines) and 3-D calculations.

23 Elastic/Nonlinear Viscous Materials.


Applicability of KI and C*

In

the

preceding

approximated

sections,

the

deformation

response

of

materials

was

as nonlinear viscous. The theoretical description is now extended

to include elastic deformation. Elastic strains cannot be

neglected,

and

may

even become dominant, in materials which are relatively creep resistant and
brittle, i.e., which exhibit substantial creep crack growth before the whole
specimen

creeps

extensively.

The

distinction

between predominantly elastic

deformation and extensive creep of a cracked specimen


this

chapter.

This

will

be

quantified

in

question is of some practical importance since the macro-

scopic load parameter to describe creep crack growth is


factor, KI , in predominantly
extensive creep limit.

elastic

situations,

the
while

stress

intenSity

C* applies in the

The discussion is based on an elastic/nonlinear viscous material law:


(23.1 )
with the notation as
law

reduces

to

E=

~efined

alE

in Section 3.2. In uniaxial tension, the

material

Bon. It comprises elastic behavior as the short-time

response to load application, and viscous creep as the long-time limit.

23.1 Stationary Crack under Step Loading


The basic concepts of high-temperature fracture mechanics in elastic/nonlinear
viscous materials will now be developed considering a stationary (i.e. non-propagating) crack. The presentation follows work of Riedel and Rice (1980) and of
Ohji,

Ogura

and

Kubo

(1980).

Some

of

the pertinent ideas were applied to

Mode-III loading earlier (Riedel, 1978). A creep crack growth test


by

is

modeled

letting a constant tensile load be applied to the cracked specimen suddenly

at time t = O. The instantaneous response of an elastic/nonlinear viscous

mat-

23.

302

Elastic-Nonlinear Viscous Material, KI and C*

erial to load application is purely elastic. In particular, the crack-tip field


at time t

0 is the inverse-square-root singular field:

(23.2)

where KI is the Mode-I stress intensity factor and the angular functions fij(e)
are given in Appendix B.
Also known is the stress field which is valid

at

very

long

times.

In

this

limit, the creep strains, which increase continuously, become exceedingly large
compared to the elastic strains, which remain bounded. Then
behaves

as

if

the

material

stress

field

were purely nonlinear viscous. The stress field

becomes time-independent with a spatial distribution as


20.

the

described

in

Chapter

In particular, the crack-tip field is the HRR-field with C* being path-in-

dependent and determining the severity of the HRR-singularity.


Another conclusion which can be drawn
(23.1),

is

directly

from

that very near the crack tip (i.e. if r

the
~

material

law,

eq.

0), elastic strain rates

can be neglected compared to creep rates for any time t > O. The reason is that
the

stress is unbounded near the crack tip, and therefore the creep rate (<< on

with n > 1) has a stronger singularity than does the elastic strain rate (<<
As

consequence, the

~ymptotic

a).

fields for t > 0 are determined by power-law

viscous creep alone, and must therefore be HRR-type fields:


_ ( C(t) )1/(n+1) - ( )
0ij ~
0ij e

(23.3)

Except for the time-dependent strength of the singularity, C(t), the quantities
were

defined

in

Section 3.4.4. The functional form of C(t) generally remains

undetermined by the asymptotic analysis, and more detailed considerations


be

presented

will

in the following sections. But we already know that in the long-

time limit, C(t) must approach its steady-state value, C*:


C(t)
23.1.1

C* for t

(23.4)

CD.

Similarity solutions in the small-scale creep, or short-time, limit

In the initial elastic stress concentration near the crack

tip,

the

material

creeps rapidly, which, in turn alleviates the initial stress concentration. For

23.1

Stationary Crack under Step Loading

303

short times after load application, the deformation field far

from

the

crack

tip is still predominantly elastic. The incipient evolution of creep strains in


a small but growing zone around the crack tip can

then

be

analyzed

using

boundary layer approach. The boundary conditions on the actual specimen surface
are replaced by the much simpler boundary condition that the

crack

is

consi-

dered as being of semi-infinite length and the stress field must asymptotically
approach the elastic singular field, eq. (23.2), at infinity.
specimen

geometry

enters

Then

the

outer

only into the linear elastic problem of determining

KI , which is a standard task in finite


geometries, KI-values are tabulated

element analysiS. For many specimen


(Tada, Paris and Irwin, 1973, Rooke and

Cartwright, 1974).
With the elastic singular field, eq. (23.2), being the
the

remote

initial

ions as Riedel (1978) and Riedel and Rice (1980) pointed out for
Mode-I

condition

and

boundary condition, the governing equations have similarity solutMode-III

and

loading, respectively. The similarity solutions are most easily derived

from the equation for the stress function [eqs. (3.11), (3.15) or 3.19)J

using

dimensional analysis and noting that E,B and t enter into the gove~ning
equation only as the product EBt and can therefore appear in the solution only
in this form. The similarity solutions are then found to be:
(23.5)
wlth the Similarity coordinate
r

R = K2 (EBt)27(n

I)

(23.6)

The dimensionless function L .. may depend on the material parameters n and \}


lJ
besides on the non-dimensionalized polar coordinates Rand 9. In fact, for
plane stress Lij is independent of \} (Riedel and Rice, 1980) . The validity of
eqs. (23.5) and (23.6) can be verified by insertion into the governing
equations. The material law then takes the form
(23.7)
where the dimensionless strain rate is defined as

~ij

Eij (EBt)n/(n-1)/B,

Le

is the dimensionless equivalent tensile stress and the prime in Llj denotes the
deviator. Compatibility and equilibrium are obviously satisfied by eq. (23.5)

23.

304

Elastic-Nonlinear Viscous Material, KI and C*

if they are satisfied by Eij and Lij . Further, insertion of eq. (23.5) into the
remote boundary condition, eq. (23.2), shows that:
for R

(23.8)

-+ "'.

Thus the whole system of equations for Lij and Eij depends on R, but not on r
and t seperately, which proves that eq. (23.5) is a solution of the field equations and boundary conditions. To obtain a complete

solution,

the

system

of

equations for Lij and ~ij must be solved. This has been carried out for Mode
III by Riedel (1978) numerically based on a stress-function formulation. For
Mode I,

finite

element

solutions

by Ehlers and Riedel (1981) start from the

original field equations making no use of the similitude of the solutions. Some
features of the solutions can also be derived analytically. All this is summarized in the following sections.
Once the solution for the stress field is known, elastic strains

are

obtained

by Hooke's law, whereas the creep strains take the form


3

(n-1) (EBt)-1/(n-1)

'" Ln - 1L!.

EIR

1J

IR

dR.

(23.9)

This latter result follows by time-integration of the creep rate with the timedifferential replaced according to eq. (23.6): dt/t
23.1.2

-[(n-1)/2J dR/R.

The crack-tip field in the short-time limit

An HRR-field which is compatible

with

the

similarity

solutions

necessarily

depends on KI , rand EBt in the following form:


2
KI (1-}) /E )1 /(n+1)
(n+1 ) I B r t
n
ct

C\j

0.. (e).

(23.10)

1J

or, in terms of the factor C(t) which was introduced in eq. (23.3):
C(t)

ct

2
KI (1-v2 )/[(n+1)EtJ.

This result is for plane strain; for plane stress,


deleted.

The

numerical factor

ct

(23.11 )
the

factor

1-v2

must

be

depends on n (and additionally on v for plane

strain), but its precise value remains undetermined by the asymptotic analysis.
However,

Riedel

and

Rice (1980) provide arguments that the J-integral, which

23.1

Stationary Crack under Step Loading

305

is, strictly speaking, path-dependent in elastic/nonlinear


should

viscous

materials,

be

approximately path-independent in the present problem. Equating the


far-field value, J = Ki(1-v2 )/E, with the J-value calculated from the near-tip

field leads to the estimate a = 1. This analytic estimate was confirmed numerically by Ohji, Ogura and Kubo (1980), Ehlers and Riedel (1980),

Ehlers

(1981)

and Bassani and McClintock (1981). In the range n = 3 to 10, the finite element
results fall within the range a = 0.9 to 1.24, the precise value

depending

on

the way in which a is evaluated numerically (Ehlers, 1981).


Since the time appears in the denominator, eq. (23.10) describes
crack-tip

stress

that the crack-tip field in the short-time limit is determined


stress

decreasing

field, as one would expect. Further, it is important to note

intensity

factor

as

solely

by

the

far as the loading system and the outer specimen

geometry are concerned. Near-tip fields in differently shaped specimens are the
same

if

KI

is

the same. This is so since in the short-time limit a boundary

layer formulation is possible in which KI alone determines the remote

boundary

condition.
23.1.3

The complete stress field in the short-time limit

In the short-time limit, the near-tip HRR-field is embedded within

the

remote

elastic singular field. Over the whole range of distances the equivalent stress
can be described by the interpolation formula between eqs. (23.2) and (23.10):

ae

(23.12)

which reproduces the limiting cases of small and large r and which agrees

with

the finite element results of Ehlers (1981) in the transition range to within a
few

per

cent.

approximating

Figure
the

23.1

shows

an

example.

An

even

of

the

way

of

stress field, which is also illustrated in Fig. 23.1, is to

calculate the line on which the equivalent stresses of the


and

simpler
near-tip

HRR-field

elastic far field are equal and to take the HRR-field inside, and

the elastic field outside that line. This matching of

singular

fields

yields

stresses which do not deviate by more than 15% from the finite element results.
The coordinates used in Fig. 23.1 are the
eqs.

Similarity

coordinates

defined

in

(23.5) and (23.6). In physical dimensions, the picture is momentarily the


same, but the near-tip field diminishes in proportion to t- 1 /(n+1) and the
transition to the far field moves to larger distances as r ~ t 2 /(n-1).

23.

306

Elastic-Nonlinear Viscous Material, KI and C*

l5

. . C:~~

to \(FE

interpolation
formula

05

01

(elastic

".

Q2 03 04 05
distance R ~

06

Fig. 23.1. Normalized equivalent stress vs. normalized distance from


crack tip along 6 = 90 0 Comparison of interpolation formula, eq. (23.12),
matched asymptotic fields and finite element results; n = 5. v = 0.3.

23.1.4

The creep zone

From the time-dependent HRR-field given in eq. (23.10), the creep

strain

near

the crack tip follows by time integration of the material law to be


e:~~r)

(23.13)

lJ

n- 1
where e: ij
(3/2)
1J.. Obviously, the creep strain has a singularity of
-n/(n+r)
the order r
, whereas the elastic strain is proportional to stress.
(el) r- 1/ (n+1). This confirms the remark made earlier that near the
i.e. , e:
crack

tip

creep

strains dominate, while far away elastic strains dominate in

the short-time limit. The different behavior of creep strain and elastic strain
suggests

defining

a creep zone around the crack tip. Its boundary is defined.

somewhat arbitrarily, as the line (in two-dimensional problems) on which the


equivalent creep strain is equal to the equivalent elastic strain. This
definition is. of course, not confined to the short-time limit. but it will

be

evaluated for this case first.


From the fact that the stress and
manner
its

strain

fields

develop

in

self-similar

in the short-time limit, it is clear that the creep zone also preserves

sha~e

while it grows in this limiting range. The form of the similarity co-

23.1

307

Stationary Crack under Step Loading

ordinate, R

r/t 2/ (n-1), requires that the creep zone must grow according to
r

(23.14)

cr

As all features of the field in the short-time limit, rcr depends on

load

and

specimen geometry only through Kr and expands around the crack tip according to
the time law rcr ~ t 2/ (n-1)
The dimensionless shape function Fcr(S) was
calculated by Riedel and Rice (1980) using an approximate analytic method based
on eq. (23.12), which leads to the dashed curves in Fig. 23.2. Figure 23.2 also
includes

results

of

finite element calculations by Ehlers and Riedel (1981).

These calculations were extended


creep,

limit.

Figure

to

beyond

the

short-time,

or

small-scale

23.3 shows the evolution of the creep zone in a compact

specimen for longer times.

plan.e 1.0
strOIn

.,_

_./'nolytic

\\

c5

O.S

1.0

Fig. 23.2. Shape of the creep zone for


plane strain (upper half) and plane
stress (lower half) for small-scale
creep. n = 5, v = 0.3.
(From Riedel, 1984a).

Fig. 23.3. Evolution of plane-strain


creep zone in a CT-specimen for
tlt1 = 2.8, 4.1, 4.4, 5.9, 9.1, 12.8
[t 1 is defined in eq. (23.15)].
n = 5, v = 0.3.
(From Riedel, 1984a).
0

23.

308

Elastic-Nonlinear Viscous Material, KI and C*

Fig. 23.4. Maximum distance of creepzone boundary from crack tip vs. time.
Plane strain, n = 5, v = 0.3. Dashed
line: small-scale creep approximation.
CCT = center-cracked tension
3PB = three-point bend
CT = compact tension
DECT = double-edge-cracked tension

CT

~B

OL-~'--~2~~3~~'~~
time ( ~'la)(EBt) 2/{n-I} - -

Figure 23.4 shows finite element results of Ehlers (1981) on the maximum extent
of

the creep zone in various specimen geometries all having a/W-ratios of 0.5.

In the short-time limit, all specimens must approach the

universal

short-time

result, eq. (23.14). At longer times, the deviations from the short-time behavior parallel tnose discussed for the plastic zone size in elastic-plastic material

(Larsson and Carlsson, 1973, Rice, 1974). For the creep-zone size the so-

called T-stress term, that is the second term in a series expansion


times,

of

which

eq.

(23.14)

is

the

first

term,

for

short

must be proportional to

K~.anet(EBt)3/(n-1). As Fig. 23.4 shows these deviations from the

small-scale

creep limit may have either sign depending on the specimen geometry.
23.1.5

A characteristic transition time

The transition from the small-scale creep limit to the steady-state creep limit
was

described above by the spread of the creep zone across the ligament of the

specimen. A useful characterization of the duration of


obtained

by

the

the

transient

the

be

following definition. A characteristic time t1 is defined by

equating the relaxing HRR-field, eq. (23.10), which is valid for


with

can

steady-state

short

times,

HRR-field, eq. (20.16), which is valid for long times.

This definition illustrated in Fig. 23_5 leads to


2

a KI (1-v )
(n+1)

C*

(23.15)

for plane strain; for plane stress the factor (1-v2 ) is deleted. It was pointed

23.1

309

Stationary Crack under Step Loading

100

10

01

time t / t , -

Fig. 23.5. HRR-field amplitude, C(t), after a step load. Long-time, and
short-time limits (dashed lines) intersect at t=t 1 rnterpolation formula,
eq. (23.16), and finite element results of Ehlers (1981).

out earlier that the factor a is well approximated by unity.


For a given load and specimen geometry, Kr and C*, and
computed.

Comparing

therefore

t 1,

can

be

the computed value of t1 with the expected or actual test

duration provides an idea of whether creep crack growth occurs under Kr-controlled elastic conditions, or under C*-controlled conditions. The ranges of
validity of Kr and of C* will be displayed on a load parameter map
25.1.4.

The

characteristic

time

in

Section

t 1 ,and other similar characteristic times,

separate these regimes.


23.1.6

Interpolation formulas for the transient regime

The finite element results for the HRR-field amplitude C(t) can be approximated
well

in the whole range from short to long times by the simple formula (Ehlers

and Riedel, 1981, Ehlers, 1981):


C(t)

(1 + t 1 /t)C*,

(23.16)

Which reduces to eq. (23.11) in the short-time limit and to C* for long
As

Fig.

23.5

shows,

this

formula

times.

fits numerical results for most specimen

geometries sufficiently accurately. The possible use of C(t) as a parameter

to

describe creep crack growth will be discussed in a later section.


~he

instantaneous load-line deflection upon load

deflection

~el.

application

is

the

elastic

The next term in a series expansion for small times must be

of

the order ~elrcr/a


in

Elastic-Nonlinear Viscous Material, KI and C*

23.

310

elastic-plastic

t 2/ (n-1), which corresponds to the plastic zone correction


materials (Edmunds and Willis, 1977). Hence, the load line

deflection rate for short times must vary as


~

el

cr

la t-(n-3)/(n-1).

(23.17)
A~

If this is normalized by the steady-state deflection rate,

a B a~et'

and

if t1 is introduced, it follows that for short times


(23.18)
with a factor of proportionality,
results

of

is

obtainable

from

finite

element

Ehlers (1981) as roughly a = 1.3 for CT-specimens and a = 0.77 for

three-point bend and double-edge


shows,

which
cracked

tension

specimens.

As

Fig.

23.6a

the numerical results can be approximated well in the whole time domain

by a Simple additive superpostion of the short-time and long-time limits.

(The

numerical results are indistinguishable from the analytic curve in Fig. 23.6a).
It may appear counter-intuitive that the deflection rate is
creep

zone

is

still

small

than

it

is

when

the

extensively. But apparently the rapid localized straining


highly

stressed

creep

zone

greater

whole
in

when

specimen
the

the

creeps

small,

but

leads to greater load-line deflection rates than

does creep in the whole specimen after stress redistribution. As a consequence,


the

deflection-vs.-time

plot

resembles a uniaxial creep curve with a primary

stage (Fig. 23.6b), although the material law does not include

primary

creep.

The tertiary stage shown in (Fig. 23.6b) arises from crack growth .

100

oj

3PB
DECT

~
.~

.
lil
;;::

,d/,d.=1 +0.77(t, /r/n-3J/(n-IJ


...... ......

..... .....

.......

.............

...
-..:....-------.....

0.1

10

time t/t, -

10

15

20

t/t, - -

Fig. 23.6 a) Load-line deflection rate vs. time for n=5, v=0.3 (log-log plot).
b) Load-line deflection vs. time (schematic for constant load).

23.1

Stationary Crack under Step Loading

311

Finally, it is interesting to note that A(t) and C(t) exhibit a similar


ior

for

behav-

the whole range of times. In the short-time limit, the two quantities

differ slightly, but for large n the different time exponents, -(n-3)/(n-1) vs.
-1,

are

similar.

This

similarity in behavior of the deflection rate and the

crack-tip field suggests to define a quantity


(23.19)
which is easily measurable with g2 as defined in eq. (20.15). The parameter Ct
is meant to be an approximation to C(t). In fact, the two quantities converge
to C* for long times, while for short times they exhibit the slight

difference

in the time dependence mentioned above. Hence Ct can be used as a parameter to


characterize creep crack growth in the transition range in an approximate way.
The

ct-parameter proposed by Saxena and Landes (1985) is practically identical

with the present definition, while Saxena's (1986) definition differs slightly.
23.1.7

Possible generalizations and related work

The analysis

of

stationary

cracks

in

elastic/nonlinear

viscous

materials

described above can readily be generalized in several ways.


1) If Young's modulus and the coefficient of Norton's creep
functions

of

time,

the

law

are

explicit

pertinent solutions are obtained by replacing EBt by

fEBdt in all formulas (Riedel and Rice, 1980).


2) If the applied load increases according to a power-law

in

time,

solutions

are also possible as will be shown in Chapter 28 on cyclic loading.


3) A sharp notch in elastic/nonlinear viscous material can be analyzed just
a

as

sharp crack. As described in Appendix B, the elastic stress field at a sharp

notch is
(23.20)
where NI is a notch stress intensity factor, s is the eigenvalue tabulated in
Table B.2 in Appendix B as a function of the included notch angle, 2a, and the
gij'S are the angular eigenfunctions pertaining to s. With eq. (23.20) as an
initial condition and as a remote boundary condition, similarity solutions are
possible just as in the limiting case of a sharp crack.

312

23.

Elastic-Nonlinear Viscous Material, KI and C*

4) If strain rates are measured over a wide range of


hyperbolic-sine

stresses,

the

following

law is often more suitable than Norton's power law to describe

the observed creep rates:


(23.21)
where Eo' 00 and n are material parameters. Bassani (1983) has carried out a
Mode-III analysis of cracks in such a material. Many of the features found in
power-law materials are also found in hyperbolic-sine law materials. There is a
similar

transient

from initially KI-controlled conditions to extensive creep,

where C* dominates. The near-tip fields, however, have a different form with
logarithmic stress singularity, 0

~n(1/r),

ions

are

given

in

Section

20.2.1

no

and the equations for C*-calculat-

longer

exact,

although

the

form

C* = g20netd is expected to remain approximately correct.

23.2 Stress Fields at Growing Cracks in Elastic/Nonlinear


Viscous Material
In purely power-law viscous materials, the crack-tip singularity was shown to
have the form 0 r- 1/ (n+1) irrespective of whether the crack is stationary or
whether it grows. If, however, elastic strains
crack-tip

fields

remain

HRR-type

are

taken

into

account,

the

fields only if the crack is at rest, while

crack growth brings about a change in the character of the fields (Hui and Riedel, 1981). The stress singularity will be found to have the form 0 r- 1/ (n-1)
for both, Mode-I and Mode-III loading. Since the
simpler,

formulas

for

Mode

are

the arguments which lead to the new singularity will be presented for

Mode III, but the results to be given in Section 23.2.2 will refer to
Sections

23.2.3

to

23.2.5

deal

In

many

practical

Mode

I.

with the entire - as distinct from only the

asymptotic - stress fields for steady-state and non-steady-state growth


tions.

III

condi-

cases, the new field is not very important, except

possibly under KI-controlled conditions. The reader who is primarily interested


in practical C*-testing may proceed to Section 23.3.
23.2.1

Derivation of the singularity at growing cracks for Mode III

The governing equation for the stress function


was

given

Win antiplane shear (Mode III)

in eq. (3.11). If the coordinate system moves with the crack tip in

the positive x 1-direction, the time derivative at a material point is given by

23.2

Stress Fields at Growing Cracks

313

(23.22)
where a/at is the time-derivative in the moving system and a is

the

(possibly

time-dependent) crack growth rate. Then eq. (3.11) becomes

o
with

B = ~+lB.

the

elastic

(23.23)

The first and second terms are linear in wand

strain

rates,

while

the

correspond

to

third term is nonlinear and describes

power-law creep.
For an asymptotic analysis, it is important to note that the first term in

eq.

(23.23) exhibits a stronger singularity near the crack tip than does the second
term, since it contains higher differentials in Xi' Therefore, the asymptotic
problem is always a steady-state problem with the term a/at deleted.
Next, it is shown that the HRR-field cannot persist near a growing
To

show

tip.

this, assume that the third (nonlinear) term in eq. (23.23) dominates

and the elastic terms can be neglected. Then


equation

crack

is

the

HRR-field

which

the

solution

of the remaining
is characterized by W r n/ (n+1). To check

whether the neglect of the first and second terms was really justified,
substitute w r n /(n+1) into eq. (23.23). Then it is found that the first term
is proportional to r-(2n+3)/(n+1), the second varies as r-(n+2)/(n+1) and the
third

as

r-(2n+1)/(n+1). Obviously the first term has the strongest Singular-

ity, which is incompatible with the assumption that the third term dominates.
Next assume that the first term in eq. (23.23) dominates and
are

negligible.

The

solution

of

the

remaining

equation,

well-known inverse square root stress singularity or, in terms

that

the

others

V2w = 0, is the
of

the

stress

function,

w r1/2. Substitution into eq. (23.23) shows that now the first and
-5/2
-(n+2)/2
third terms vary as r a n d r
,respectively. Hence, the assumption
that the first term dominates is correct if n < 3. In this case, the asymptotic
field near a growing crack tip is
possibility

will

indeed

the

elastic

singular

field.

not be pursued here and the interested reader is referred to

work of Hart (1980, 1983). If, however, n is larger than 3, the first
less

singular

This

than

the

third.

Hence

term

is

the elastic singularity cannot be the

correct asymptotic field if n > 3.


Since for n > 3, neither the linear nor the nonlinear term alone can

determine

23.

314

the

asymptotic

behavior,

Elastic-Nonlinear Viscous Material, KI and C*

they must be of equal order of magnitude to balance

each other. If we try solutions which have the factorized form


(23.24)
that

s = (n-2)/(n-1).

the balance of linear and nonlinear terms requires


singularity is therefore G a r- 1 /(n-1).

The

function f( 8)

(23.24)

stress

(neglecting
differential

is

obtained

by

substituting

eq.

dimensionless
into

The

angular

eq.

(23.23)

the second term, i.e. a/at = 0) and solving the resulting ordinary
equation

differentiation

of

numerically.

~.

The

stress

components

follow

by

This procedure was carried out by Hui and Riedel (1981)

whose results for Mode I are shown below.


23.2.2

The growing crack singularity; results for Mode I

The asymptotic stress and strain fields near the tip


elastic/power~law

of

crack

growing

in

viscous material with n > 3 are given by:

)l/(n-1)

EBr

Eij

(Il /E)

EBr

Gij

)l/(n-l)

(23.25)

()

Eij

(8)

(23.26)

Such a field with a r- 1 /(n-1)-singularity will be referred to


(after Hui and Riedel, 1981). Figure 23.7 shows the

as

dim~nsionless

an

HR-field

angular func-

2.0

0~--~30----6~0----9~0~--~12~0~--~15~0~~180

e-

O~---L--~----~--~----~~~

30

60

90

e-

120

150

180

Fig. 23.7. Angular distribution of the asymptotic stress at cracks grow ing in
elastic/nonlinear viscous materials. Plane strain, v=0.5. After Hui and Riedel
(1981) reproduced by permission of Martinus Nijhoff Publishers, Dordrecht.

23.2

Stress Fields at Growing Cracks

ti ons a ij
A

is

equal

315

e , which are normalized such that the maximum of ae = (3 aija ij /2)1/2

()

to

unity.

The

n-dependent

numerical

factor

A,

was computed as

a 4 = 1.042 and a6 = 1.237 for plane strain (with Poisson's ratio v = 0.5), and
as a4 = 0.815 and a6 = 1.064 for plane stress, where the governing equation is
independent of v. For Mode III, the numerical result is represented
0.2%

accuracy

~ = [0.29(n-3)]1/(n-1),

by

replaced by G and

and

E and

to

B.

The asymptotic fields at growing cracks described by eqs. (23.25) and


have

within

B in eq. (23.25) are

remarkable

(23.26),

properties. Stress and strain are entirely independent of the

applied load. Only material parameters and the current crack growth rate, but
not growth rates at prior times, determine the asymptotic fields. (Of course,
this independence of load and prior history cannot hold for the remote fields).
Possible consequences for creep crack growth will be discussed later.
It should be noted that some of the properties of the growing-crack singularity
depend sensitively upon the material model used. In this connection, the reader
is referred to recent work by Yang and Freund (1986) on the interrelationships
between crack-tip fields in rate-dependent and rate-independent materials.
23.2.3

Fields for steady-state crack growth under small-scale creep conditions

Small-scale
behavior

creep

conditions

are

characterized

by

predominantly

elastic

of the specimen. A boundary layer approach is then appropriate, i.e.,

the stress and strain fields near the crack tip are analyzed subjected

to

the

remote boundary condition of asymptotic approach to the elastic singular field,


eq. (23.2). Steady state means

that

the

a/at-term

in

eq.

(23.23)

can

be

neglected compared to the other terms in this equation.


The equations and boundary conditions which

define

the

steady-state,

small-

scale creep problem contain a set of parameters which can be combined to form a
group having the physical dimension of length:
(23.27)
This is the characteristic length for the range of validity of
HR-field.

Further,

the

asymptotic

for reasons of dimensional consistency, the entire steady-

state stress field in the small-scale creep limit must have the form

23.

316

Elastic-Nonlinear Viscous Material, KI and C*

(23.28)
R1 = r/r 1

where

~ R~1/(n-1)aij(S)

The
dimensionless
function
Eij
must
approach
for small R1 in order to recover the near tip HR-field Ceq.

(23.25)]. For large R1 , eq. (23.28) must give the elastic singular field, i.e.,
Eij = f ij (S)/(2wR 1 ) 1/2 The knowledge of Eij at intermediate distances requires
a numerical solution of the field equations. Such a finite element analysis was
carried

out

suggested

by

by

Hui

Riedel

(1983)
and

who

found

that

(1981)

is

Wagner

a simple interpolation formula


reasonably

accurate.

For

the

equivalent stress the interpolation formula is:


(23.29)
where fe is the angular function pertaining to the
linear

elastic

case.

Alternately,

the

field

equivalent
can

also

stress

in

the

be approximated by

matching the near-tip field and the far field together on the line on which the
two terms on the right-hand side of eq. (23.29) are equal. This is analogous to
the procedure adopted in Section 23.1.3.
The steady-state solution is valid as long as the alat-term in eq.

(23.23)

is

negligible. It can be shown that this is equivalent to requiring that the crack
growth rate,

i,

and the stress intensity factor, KI , vary to an only small


extent while the crack tip traverses the zone of size r 1 . Thus the steady-state
requirement is lalla < a/r 1 Inserting r 1 from eq. (23.27) gives:
(23.30)
where
23.2.4

is the acceleration of the crack tip.


Steady-state crack growth during extensive creep of the whole specimen

If the whole specimen creeps extensively, the asymptotic HR-field


in

an

remote

HRR-field
boundary

characterized
condition,

the

by

C*,

general

is

embedded

eq. (20.16). With the HRR-field as a


form

of

the

field

follows

from

dimensional considerations to be (Riedel and Wagner, 1981):


(23.31 )
with

(23.32)

23.2

317

Stress Fields at Growing Cracks

Now, r 2 characterizes the range of validity of the HR-field. The requirements


for the approximate validity of the steady-state solution are given by:
(23.33)
In many practical cases

of

C*-controlled

crack

growth,

the

characteristic

length r 2 turns out to be exceedingly small numerically, often smaller than


atomic dimensions. Therefore the HR-field will henceforth be neglected under
C*-controlled conditions.
23.2.5

The evolution of the asymptotic field under non-steady-state conditions

Let a constant load be applied at time t = 0 and let the crack grow at a

time-

dependent rate a ~ t B For B = 0, this includes the case that the crack starts
growing at a constant rate at time t = O. The growth rate is artificially
prescribed here rather than calculated from a crack growth criterion.
The evolution of
procedure

the

stress

field

can

be

visualized

to growing cracks by Bassani and McClintock (1981). There


fields

using

the

matching

described in Section 23.1.3 for stationary cracks, which was applied


are

three

singular

which are matched together as illustrated in Fig. 23.8a, namely the HR-

field, eq. (23.25), which is valid very near the crack tip, the elastic
lar

field, eq. (23.?), which is valid far from the tip if t

{ield, eq. (23.3). Each of the fields

is

assumed

to

singu-

< t 1 , and the HRR-

dominate

in

region

bounded by the lines 'on which the equivalent stress equals that of the adjacent

HR

a)

elastic field

....

GJ
lJ

c::
.E
.!!!

c::

.E

.!!!

"tJ

----

a:rcr 'i

'fiR
log r

....

GJ
lJ

\:;)GJ

c) ts= 10 t,

b) t, ts

"tJ

ts
time t - -

tl
time t - -

Fig. 23.8. a) Ranges of validity of HR-field, HRR-field and elastic field.


b,c) Evolution of characteristic lengths after crack growth initiation.

318

23.

Elastic-Nonlinear Viscous Material, KI and C*

singular field. Figure 23.8 is drawn

a=

for

constant

crack

growth

rate

(i.e.

0) and for constant stress intensity factor. Then, among the three singular

fields, only the HRR-field is time dependent. It is apparent how the ranges
validity

grow

and

shrink

if

of

the line representing the HRR-field is shifted

downwards for increasing time. The range of validity of the HRR-field

has

the

order

of the creep-zone size and is denoted by r cr ' while that of the HR-field
is r HR The results are shown schematically in Figs. 23.8b and c. At short
times, the range of validity of the HR-field expands according to

(23.34)

Here, all numerical factors such as (n+1) were omitted for simplicity. After

time of the order


(23.35.)
the range of the HRR-field has shrunk to zero

and

the

HR-field

is

directly

embedded

in

the

elastic field with r HR approaching r 1 from eq. (23.27). This


steady-state growth situation is depicted in Fig. 23.8b. If, however, the
specimen

approaches the steady-state creep limit before steady-state growth is

attained, i.e., if t1 < t s ' then the HRR-field displaces the elastic field and
the HR-field is eventually embedded in the steady-state HRR-field characterized
by C*. This Situation, in which r HR must approach r 2 defined in eq. (23.32) for
long times, is shown in Fig. 23.8c.
Now, by Simple matching, a rather complete picture of the transient fields
been

obtained.

The

can be shown by the formulation of boundary layer


obtained

has

question is how accurate are the results. Analytically it


problems

that

the

results

by matching are reasonable in an asymptotic sense (Hui and Wu, 1986).

Numerically, Hawk and Bassani (1986) have performed finite element calculations
for the transient, growth problem (for Mode III). They found that the fields obtained by matching agree with their finite element results to within 15%.
If the prescribed crack growth rate

varies

as

oc

t-(n-3)/(n-1),

the

crack

growth increment, a-ao ' and the creep zone expand at a fixed proportion as long
as small-scale creep conditions prevail. In this case, the full governing eq.
(23.23),

including

the

a/at-term, has simi-Iarity solutions as Hui (1986) has

23.2

319

Stress Fields at Growing Cracks

pointed out. The character of these similarity solutions depends on

a/r cr

For

slow

the

ratio

crack growth, the HR-field is embedded within the HRR-field,

which itself is contained within the elastic singular

field.

For

fast

crack

growth, the HR-field is directly surrounded by the elastic singular field.

23.3 Crack Growth in Elastic/Nonlinear Viscous Material Subject


to a Critical-Strain Criterion
In the preceding section, stress fields were analyzed for
growth

rate.

Now

prescribed

fields calculated above with a critical strain criterion: the crack


at

such

rate

crack

the crack growth rate is determined by combining the stress


that

must

grow

at a distance Xc ahead of its tip the equivalent creep

strain reaches the critical value for local failure,

Ef

Depending on the relative magnitude of the characteristic lengths r cr ' r HR , Xc


and a-a o ' there are several limiting ranges. In the following, the two most
important ones will be analyzed. In the first example, r HR is assumed to be so
small that the HR-field can be totally neglected. This is an important case for
ductile materials which fail in the C*-controlled limit, but
non-negligible

elastic

transient

at

the

beginning

which

exhibit

of the test. The second

example refers to failure under small-scale creep conditions in which case

the

HR-field may play an important role.


23.3.1

Analysis of the case r HR

< Xc and a-ao < rcr

We start with the analysis of the case where the


crack

propagation

since

its

range

HR-field

plays

no

in

of validity is smaller than the relevant

length for the fracture process, xc. At the same time, it is assumed
speCimen

role
that

the

reaches the limit of extensive creep before it fails, but the initial

stages of crack growth

may

be

influenced

by

the

elastic

transient.

This

corresponds to the situation shown in Fig. 23.8c for the limit r 2 ~ O. Hence,
the theory to be developed here is an extension of the crack growth theory for
viscous

materials

to

include

the effect of the initial elastic transient on

crack growth behavior.


As in Chapter 21, the stress field around the crack is approximated by the HRRfield,

but now with a time-dependent amplitude, C(t) from eq. (23.16), instead

of C*. A time-dependence of C(t) does not invalidate the methods used in

Sect-

23.

320

ion

21.1

Elastic-Nonlinear Viscous Material, KI and C*

since, in the integral equation (21.2), the combination of variables

[C(t)J n/ (n+1)/a(a') can be regarded as the dependent variable. This is possible


because

C(t)

is

an

implicit function of a' since a' and t are single-valued

functions of one another during growth. Therefore, the results for a


eqs.

(21.4)

and (21.6) remain valid if C* is replaced by C(t)

For short times, the term t1/t

i~

C(t),

which

represents

given

by

(1 + t 1 /t)C*.

elastic

transient

effects, enhances the crack growth rate substantially.


Figure 23.9a schematically shows the predicted crack growth rate in a constantload

test. During the transient, the growth rate decreases, until the trend is

reverted by an increase in C* due to crack elongation. The left branch of the


[
curve in Fig.9a is obtained if a is plotted against C*
aBo n+1
net g 1 according to
eq. (20.7)J. Plotting a against C(t) leads to the right branch. A similar curve
is expected if Ct = g20net~' which was introduced in eq. (23.19), is used on
the abscissa. After the elastic transient is over, C(t) and Ct approach their
steady-state value, C*, and all branches merge into a single curve. When integrated, the cPack growth law yields the crack length as a function of
shown

in

Fig.

time

as

23.9b. Initially the crack extends rapidly due to the enhanced

strain rates during the elastic transient. It should be noted that

this

rapid

crack extension may reduce the lifetime of a cracked specimen considerably. The
magnitude of the effect increases with the ratio t 1 /t f . It
materials

which

are

is

pronounced

for

not very ductile. An experimental example on an aluminum

alloy will be given in Chapter 27, and another one on 2 ' /.Cr-1Mo steel will
shown in Section 25.3, after primary-creep effects have been considered.

time

---

Fig. 23.9. a) Crack growth rate in the transition range,


plotted against C*, Ct and C(t). b) Crack length vs. time (schematic).

be

23.3

Crack Growth Subject to Critical-Strain Criterion

321

23.3.2 Crack growth subject to a critical-strain criterion in small-scale creep


In

this

section,

(calculated)

the

creep

case

zone

is

analyzed

boundary

where

relatively

the

early

crack
in

overtakes

the

the lifetime of the

t s ). The initial transient stage


evolution of the HR-field is neglected and the stress field is assumed

specimen as illustrated in Fig. 23.8b (t 1


in

the

to be the steady-state field at a growing crack tip, given by eqs. (23.28)


(23.29).
in

a and

The steady-state assumption is valid for t > ts provided that changes


KI obey the conditions for steady-state growth stated in eq. (23.30).

Now the stress field is combined with the critical-strain criterion


growth.

and

for

crack

By time-integration of the n'th power of 0e one obtains the equivalent

creep strain. In analogy to Section 21.1, one is led to

an

integral

equation

for the growth rate, which now takes the form:


A

(23.36)

Under the integral sign, the normalized growth rate A is taken at the point A'.
Further, A = (a-ao)/x c and

A ae: f
BXc

2Tfx c

(23.31)

KIfe

KI fe (Cln Oe )n-1
/ 2TfX c E

(23.38)

e: f

The strain accumulated while the crack is stationary, e: o ' is approximated by


the creep strain which would be developed in an elastic singular field, i.e.

e:o/e: f
field

(1+A)-n/2. The angular function

(x c /r)n/2
ahead

of

the

elastic

singular

the crack is fe = (1-2v) in plane strain, and fe = 1 in plane

stress, while 0e at

e = 0 can be taken from Fig. 23.7.

For large A, the solution


steady-state

of

growth

rate

Unfortunately the resulting

of

eq.

(23.36)

approaches

is obtained by setting
integral

cannot

be

A=

steady

state.

The

const and letting A ~

evaluated

in

closed

00.

form.

Therefore the result obtained numerically by Hui (1983) is shown in Fig. 23.10.
For a given KI , the growth rate has two branches. The lower branch is unstable
in the sense that the strain ahead of the crack increases if the growth rate

23.

322

Elastic-Nonlinear Viscous Material, Kr and C*

10

K.l/K.'I,mm
. ~
Fig. 23.10. Steady-state growth rate vs. Kr after Hui (1983).

increases. The upper branch is stable. For large a, the range of the asymptotic
HR-field, r 1 , becomes negligible so that the elastic field alone determines the
evolution of creep strain. rn this limiting case the steady-state growth rate
is found to be

A=

2/(n-2) or, in physical dimensions:

fe Kr

121TXc

The minimum stress intensity factor, below

which

(23.39)

no

steady-state

growth

is

possible, is proportional to that derived below for nonsteady-state growth Ceq.


(23.41)]. Numerical values for Kr,min and

amin were

given by Hui (1983).

Nonsteady-state solutions of eq. (23.36) were obtained


(1981)

by

Riedel

Wagner

numerically. They point out that, in analogy to the two branches in the

steady state, there are either two solutions to eq. (23.36), one
stable,

and

or

of

which

is

there is none. A series expansion of eq. (23.36) for small A shows

that a solution exists only if initially kr is large enough to satisfy


kr > 0.5 n(2n-1)/(n-1)/(n_1),

(23.40)

or in physical dimensions:
n(2n-1)/(n-1)
(23.41)

------------~,--- 121TXc EE f

2(n-1)(an e)n-'f e

Results of Riedel and Wagner (1981) showing the evolution of the

crack

growth

23.3

Crack Growth Subject to Critical-Strain Criterion

rate

are

given

in

323

Fig. 23.11. The picture includes two cases in which KI is

assumed to decrease as a function of crack length in order

to

show

that

eq.

(23.36) ceases to have solutions at small kI Beyond the points marked by


crosses, only the trivial solution of eq. (23.36),
= 0, remains pOSSible,

i.e.,

the

crack

must

stop. Of course, the present theory, which is based on

steady-state stress fields, becomes incorrect at the discontinuities of


at

a,

both

A = 0 and at the point where the crack stops. In reality, a creep zone will

spread around the decelerating or arrested crack and crack growth will continue
or

will

be

re-initiated.

Wu,

Bassani

and Vitek (1986) have carried out an

improved analysis by including the HRR-field, which was neglected by Riedel and
Wagner

(1981).

Wu

et

al used the nonsteady-state stress fields described in

Section 23.2.5. They find that crack growth is possible also at small values of
KI , but the behavior
defined in eq. (23.41).

becomes

very

irregular

in the range below KI,min as

The reason for the instabilities in crack growth predicted


theory

due

by

the

simplified

to Riedel and Wagner (1981) is the fact that the strain associated
.1/(n-1)

with the asymptotiC field varies with crack growth rate as g ~ a


according to eq. (23.26). Thus if the growth rate decreases, the strain also
decreases until the critical strain is no
counter-intuitive,
here (which are
criterion.

reached.

This

behavior

is

but it is a necessary consequence of the stress fields used

admittedly

Whether

longer

or

approximate)

not

these

combined

effects

with

the

critical-strain

will be observable in creep crack

growth tests has not yet been clarified. The experiments

described

not carried out to stvdy this subtle problem.

"'
oi

e.0.5
:S

o ~--~--~--~--~--~

2
crack extension

{!:go Xc

Fig. 23.11. Crack growth rate as a function of crack length.

next

were

Elastic-Nonlinear Viscous Material, KI and C*

23.

324

23.4 Application to Experiments


23.4.1

The appropriate load parameter

As the analysis of the deformation fields in elastic/nonlinear viscous material


has

shown, the crack-tip fields are determined by the stress intensity factor,

KI , in the short-time limit and by the C*-integral in the long-time limit.


Therefore these parameters should be used in their respective ranges of validity to correlate crack growth rates. As an indication of which limit applies, it
is

recommended

to

calculate

the

characteristic

time

t1 from eq. (23.15).

Alternately, the applicability of KI can be proven by showing that the measured


load-line deflection is primarily caused by the elastic compliance of the specimen. Elastic compliances are available from a number of sources

(e.g.

Saxena

and Hudak, 1978). On the other hand, the use of C* is justified if the deflection rate has reached a steady state. In power-law viscous

materials,

the

de-

flection rate is related to the load and to the crack length by eq. (20.11).
In the following two subsections, two extreme cases are described in

order

to

illustrate the use of KI and C*. The transient regime, which may be important
in other materials, will be discussed in a later section, after instantaneous
plasticity and primary creep have been taken into account.
23.4.2

A 1Cr-1/2Mo steel

The heavy deformation of the 1Cr-1/2Mo steel specimen shown in Fig. 23.12
tainly

precludes

the

use

partly

described in Section 21.5.1.

cer-

of

KI , which is based on linear elastic behavior.


Experiments on this steel were carried out by Riedel and Wagner (1985) and were

ICr-l/2 Mo steel

Under testing

conditions under which the

Nimonic BOA

Fig. 23.12. CT-specimens after creep crack growth at 535 0 C (steel)


and 650 0 (Nimonic 80A).

23.4

Application to Experiments

325

lifetime of the specimens is several weeks, t1 was found to be a few

hours

or

less. Correspondingly, elasticity effects can be neglected and KI gives no correlation among the measured crack growth rates, whereas C* does (Fig. 23.13).
It is interesting to note that the crack growth rates in the as-processed steel
shown in Fig. 23.13 are very similar to those measured on the same steel having
a prior service history (Fig. 21.7Y. This is so despite a considerable loss
creep

in

strength due to carbide coarsening during service. It should be remarked

that the similarity of crack growth rates obtains only

if

the

comparison

is

based on C*. If the materials were compared for equal loads, the more creep-resistant material would exhibit slower crack growth. The
compatible

observed

behavior

is

with the prediction of eq. (21.6), where the creep properties enter

only weakly into the crack growth rate, for constant C*.
23.4.3

Nimonic 80A

In the nickel-basesuperalloy Nimonic 80A, creep crack growth at 650 0 C is

not

accompanied by any visible permanent deformation of the specimen (Fig. 23.12b).


Indeed, the characteristic times t1 obtained by Riedel and
these

Wagner

(1985)

for

tests were many years, while the specimens failed at much shorter times.

Therefore the crack growth rates are plotted as a function of KI in Fig. 23.14.
The

data show that the crack growth rate is temperature dependent when plotted

against KI , and that the gaseous environment has a marked effect. In an Ar/H 2 mixture, the crack growth rate increases by about a factor 10 as compared to

..'"

III[] C

goo
o
~

-L.

co
C

CC

Fig. 23.13. Crack growth rates in as-processed 1Cr-1/2Mo steel at 535 0 C,


plotted against KI and C*. Arrows indicate data from early stages of the tests.

23.

326

Elastic-Nonlinear Viscous Material, KI and C*

tests in air. No quantitative model exists to explain this


ect.

environmental

eff-

Tentativelyeq. (23.39), which is based on a critical-strain crack growth

criterion, was fitted to the data in air.


Riedel

The material parameters given by


44
Wagner (1985) are B = 3.310 (MPa)-n/ s and n = 13. The solid line

and

with slope 13 in Fig. 23.14 represents eq. (23.39) fitted to


pertinent

to

650 0 C in

air

at

low

the

data

points

stresses. The absolute value of the fit

requires that XcE~/(n-2) = 16 ~m; if Ef = 10%, this means Xc = 24 ~m. In the


low-stress range, the solid line is compatible with the data, whereas at higher
stresses the data were fitted with a line having slope 4 without a
reason

for

this

choice.

theoretical

The dashed line in Fig. 23.14 was drawn by a factor

7.56 below the solid line. This is the predicted effect of the reduction of
temperature from 650.oC to 600 oc, i f the temperature dependence .of B in eq.
(23.39) is identified
Qv

= 271 kJ/mol.

This

with

that

line

is

of
to

the

diffusion

coefficient

of

nickel,

be compared with the few data measured at

600 o C.
Although the data can be explained partly by eq. (23.39), it should be

ned

that

continuum

damage

mentio-

model to be described in Section 27.4.1, which

should theoretically be preferable, predicts an a Ki relationship, which is


very different from eq. (23.39). In conclusion, a complete understanding of
creep crack growth under small-scale creep conditions is lacking.

The

problem

is aggravated by the fact that neither of the models takes corrossive effects
into account, which obviously playa role in Nimonic 80A.

.0

10

's?

Fig. 23.14. Crack growth rates


in Nimonic BOA. The line with
slope 13 represents eq. (23.39).
650C ArlH2
650Cair
x 600C ArIH2
" 6(J(J'C air

+
.0

50

100

. AAN -3/2
K1 In""
m

24 Instantaneous Plasticity

In this chapter, the instantaneous plastic response of materials is taken


account.

In

into

order to illustrate the relevance of instantaneous plasticity for

creep crack growth we assume that the total strain rate can be written as a sum
of elastic, plastic and creep strain rates:
(el) + g(pl)

E: ij

ij

(cr)

E: ij

(24.1)

The elastic strain rate is given by Hooke's law, eq. (3.5), the creep
assumed

to

exhibit

power-law

viscous

behavior,

rate

is

eq. (3.6), and the plastic

strain rate is described by the incremental, power-law hardening relation

~~~l)
IJ

(3/2)

(B IN)
0

1/N-2'
ae
ae

(24.2)

aij'

Here, N is the hardening exponent, which lies in the

range

0.05

to

0.3

for

typical structural materials, and the coefficient Bo is more commonly expressed


by the yield stress, B a 1 - 1/N /E, for plastic loading (0 > 0), and is Bo
o
y
e ( 1)
1 IN
for unloading (~ < 0>. For uniaxial tension, one obtains E: p
= B a
e
0
Equation (24.1) is not meant to be an
instantaneous

plasticity

and

accurate

creep,

but

constitutive

model

combining

it allows to sort out and describe

E~~l) = 0, eq. (24.1) describes an elastic/nonlinear viscous material, which was already discussed in Chapter 23. For Ei~r) = 0,

several limiting ranges. If

eq. (24.1) describes an elastic-plastic material, which


time-independent,

elastic-plastic

fracture

plastic limit. This area has been summarized, for


Paris

(1979).

At

for

the

the

including

exam~e,

by

basis
the

for
fully

Hutchinson

and

the same time, the elastic-plastic deformation fields which

are established instantaneously upon load application,


condition

forms

mechanics

evolution

represent

the

initial

of creep strains in a material described by eq.

(24.1). These initial fields are briefly described next.

24.

328

Instantaneous Plasticity

24.1 Deformation Fields in Elastic/Plastic Material


If a cracked body of elastic-plastic material is loaded from zero load, plastic
straining

takes

place

in a plastic zone around the crack tip. As long as the

plastic zone is small enough compared to crack length and ligament


small-scale

yielding

approximation

is

valid.

In

this

width,

the

low-load limit, the

overall specimen response is elastic and the plastic zone size is given by
(24.3)
where the numerical factor is approximate.
For increasing load, the plastic zone spreads across the whole ligament and the
fully plastic limit is approached. In this case, elastic strains can eventually
be neglected compared to plastic strains. The analysis

of

the

fully

plastic

limit is relatively simple because of the fact that, under conditions specified
below, the

str~s

field in an incrementally plastic material

is

the

same

as

that in the corresponding nonlinear elastic material characterized by


(24.4)
The equivalence of incremental plasticity and nonlinear
sometimes

called

deformation

plasticity)

is

valid

elasticity
only

(which

is

for proportional,

monotonically increasing loadings and for stationary cracks. In this case,


stress

fields in incrementally plastic materials are proportional fields. This

prerequisite for the equivalence with deformation plasticity is


the

violated

near

tip of a growing crack, a fact which bounds the validity of the J-integral

in slow stable cracking in elastic/incrementally plastic materials. The


who

the

reader

is interestea in this subject is referred to work of Rice, Drugan and Sham

(1980), Gao and Hwang (1980) or of Hutchinson and Paris (1979).


The reader is also reminded of the elastic-viscous analogy which
all

results

described

implies

that

for nonlinear viscous materials in Chapter 20 are also

applicable to the fully plastic limit of elastic-plastic material.


The field near the tip of a stationary crack
determined

in

elastic-plastic

material

is

by plastic straining alone (linear elastic strains can be neglected

asymptotically). For power-law hardening, the crack-tip field is the HRR-field:

24.1

329

Fields in Elastic-Plastic Materials

(24.5)

where I1/N and Gij are taken for 1/N instead of n as in the
case, and J is the J-integral. For small-scale yielding is

nonlinear

(24.6)

while J-values in the fully plastic limit are given by Kumar et


their

Plastic

Fracture

viscous

al

(1981)

in

Handbook as a function of load and specimen geometry.

They also give a useful interpolation formula between small-scale yielding

and

the fully plastic limit.

24.2 Growth of a Creep Zone in an Initially Fully-Plastic Body


Under creep conditions, the elastic-plastic deformation

fields

represent

the

initial condition for time t = O. For strong strain hardening (i.e., if


> 1/n), the evolution of a creep zone and stress relaxation near the crack

tip

occur

in a similar way as in elastic/nonlinear viscous material. The HRR-

field, eq. (24.5), replaces the elastic singular field in its role as the initial condition and the remote boundary condition in the short-time limit.

>

Toe condition,

tip

finite time t > 0 and that a creep zone growing around the crack

for

any

1/n,

guarantees that creep strain dominates near

tip can be defined. Using similarity arguments paralleling those

the

crack

described

in

Section 23.1.1, Riedel (1981b) shows that the creep zone must grow according to
r

cr

a:

(J/Bo) (BtiB ) (N+1)/(nN-1)


0

'

(24.7)

as long as it is small and grows within a plastically deformed environment.

If

the instantaneous plastic zone is small, the creep zone grows beyond the
initial elastic-plastic boundary after some time. Subsequently, it grows in an
elastic

field

as described in Chapter 23. Similarity arguments also show that

in the short-time limit, the near-tip HRR-field relaxes in time according to


a J

(n+1) In

r t

) 1/(n+1) -G

(
)
ij B,n ,

(24.8)

24.

330

irrespective of
prevail

whether

initially.

small-scale

yielding

or

Instantaneous Plasticity

fully

plastic

conditions

Equation (24.8) above reduces to eq. (23.10) if the linear

elastic value for J is inserted. The dimensionless factor a may depend on v,

and N. The assumption that J is approximately path-independent leads to a = 1.


After long times, creep strains dominate everywhere in the

specimen,

and

the

C*-integral determines the crack-tip fields. Equating the short-time field, eq.
(24.8) and the long-time field, eq. (20.16), gives the characteristic time

for

the transition from the initial elastic-plastic behavior to steady-state creep:


J

(24.9)

(n+1)C*
This is a generalization of eq. (23.15) to
and

it

obviously

reduces

include

instantaneous

plasticity,

to that equation if the linear elastic value for J

from eq. (24.6) is inserted. If creep crack growth occurs at times smaller than
t 1,

the

crack

tip

fields

are still determined by the J-integral, while for

times larger than t 1 , C* starts to dominate. An experimental


J-controlled creep crack growth is described in Section 24.4.

example

for

If the strdin-hardening exponent and the stress exponent of Norton's creep

law

24.3 The Special Case N = lIn

happen

to

be related by N = 1/n, a simple description of the transient fields

is possible for step loading. We assume

that plastic loading at any

material

point occurs only immediately at load application (thus neglecting any additioional plastic straining by
strain

later

stress

redistribution).

Then

the

plastic

rate can be combined with the creep rate by replacing the coefficient B

in the creep rate by B(t) = B + Boo(t), where oCt) is the Delta

function.

Now

the problem is reduced to an elastic/nonlinear viscous problem with time-dependent B(t). According to Section 23.1.7, such a material law can be treated like
one with a constant coefficient by replacing time by f[1+B o o(t)/B]dt = t+Bo/B.
By this substitutlon, the crack-tip field can be obtained from eq. (23.10).
However,

in order to satisfy the initial condition at time t = 0, which is now

given by eq. (24.5), the factor a cannot be set equal to unity. If a


determined

is

again

from the requirement that the J-integral be path-independent, which

leads to a differential equation for a as a function


stress field becomes:

of

Bt/Bo'

the

near-tip

The Special Case N = lin

24.3

331

I Br [(n+l)t + BIB]

)l/(n+l)

0iJ'(S).

(24.10)

This result exhibits the correct behavior for t

0 and for Bo

O.

24.4 An Experimental Example forJ-Controlled Creep Crack Growth


Saxena, Ernst and Landes (1983) performed creep crack growth tests on AISI Type
stainless

316

steel

at

594 0 C using single-edge notched specimens in tension

(SENT). Their tests were relatively short-term tests done at high load
so

that

the

specimens

became

fully plastic directly upon load application.

Whether or not complete stress redistribution by creep had taken


crack

growth

can

be

decided

levels,

by

comparing

their

test

place

duration

before
with the

characteristic time tl given in eq. (24.9).


As an example, tl is calculated for one of their tests in which the crack
length was a = 18.8 mm, the specimen width W = 50.8 mm, the specimen thickness
Bt = 25.4 mm, and the load P = 102.3 kN (i.e. Gnet = 126 MPa). Saxena, Ernst
and Landes (1983) report the material parameters as: B = 3.4.10- 27 , n = 8,
Bo
1.5.10-6 and N = 0.53 in units of Megapascals and seconds. From eq. (20.7)
(or its nonlinear elastic analogue), J and C* follow as J = 110 kJ/m 2 and
C* = 1.5.10-3 W/m 2 , and from eq. (24.9) there results tl = 2,300 h. Typical
lifetimes

of the specimens were less than, or at most of the same order as the

characteristic time so that the specimens spent most of their lifetimes in


J-controlled

the

short-time limit. Therefore it is not surprising that creep crack

growth rates correlated well with J but did not correlate with C*.
The dominance of J in these
experimental

proof

that

experiments
the

load

was
line

additionally
deflection

supported
was

mainly

instantaneous plasticity while creep contributed only a fraction of


of ten per cent.

the

by

the

due

to

order

2S Primary-Creep Effects

In materials which exhibit a pronounced primary-creep stage it may be necessary


to

take this into account in the analysis of creep crack growth. Two different

constitutive models will be considered,


model

(Riedel,

1981b)

already

power-law

strain-hardening

and a creep recovery and hardening model (Kubo, 1983).

Time-hardening of the form


was

namely

= B(t)~

can easily be reduced to viscous creep as

pOinted out in Section 23.1.7. Time-hardening will not be pursued

further since it usually does

not

provide

an

accurate

description

of

the

can

be

deformation behavior of metals.

25.1 Strain-Hardening Model for Primary Creep


The decreasing strain rate in the primary

part

of

the

creep

curve

described by the following power-law strain-hardening constitutive relation:


~~~r)
lJ

(312)

(25.1)

B1

where B1 , m and p are material parameters, Ee = (2EijElj/3)1/2 is the equivalent strain, and the superscript (pr) denotes primary creep. For constant
stress in uniaxial tension, eq. (25.1) can easily be integrated in time to give
(25.2)
Here, strain is a power function of time.
creep law, E ~ t 1 / 3 , is obtained if p

In

particular,

Andrade's

primary-

2. A comparison with a hardening/re-

covery model is shown later (Fig. 25.2).


In Section 25.1.1, crack configurations
creep

alone

are

analyzed

assuming

that

primary

determines the deformation fields. In Sections 25.1.2 and 25.1.3,

primary creep is combined

with

elastic-plastic

strains

and

with

secondary

25.1

creep,
a

333

Strain-Hardening Model for Primary Creep

respectively. In Section 25.3 the results are applied to experiments on

21/~Cr-1Mo

steel. In Section 27.5, eventually, primary creep is included in a

damage mechanics description of creep crack growth.


25.1.1

Primary creep of the whole specimen

The stress field in a power-law hardening material described by eq. (25.1) is a


proportional

field

if the applied loading is proportional, i.e., if the loads

vary only in magnitude but not in direction (Hult, 1962, Riedel,

1981b).

Then

the stress field has a form like eq. (20.4), and the strain field calculated by
time-integration of eq. (25.1) is

related

to

the

stress

field

as

if

the

material were nonlinear elastic according to:


(25.3)
The effective coefficient

B1 (t)

is independent of

the

spatial

variables

and

depends on load ana time as

B1

[B 1 (1+p) fP(t,)m(l+p)dt,]l/(l+P)/p(t)m,

(t)

(25.4)

where pet) is the time-dependent load


pm = Ipl m- 1 p (P may be negative).

and

powers

of

P are

understood

as

Similarly, by time-differentiation of eq. (25.3), one finds that the stress and
strain-rate fields are related as if the material were power-law viscous:
t
-p/(l+p)
(3/2) B1 /(1+p) [(l+p) fp m(l+ p )dt']
pmp om-1 0i'Jo.
1

(25.5)

Since the stress/strain and the stress/strain-rate relations are

found

to

be

independent of the spatial variables, it can be concluded that in two-dimensional crack problems

th~

J-integral and the C*-integral are both path-independent

if a material characterized by eq. (25.1) is subjected to proportional loading.


Because of the time-dependent coefficients in the
(25.5),

relations

eqs.

(25.4)

even for constant load (note that the time dependences are given by J
Clf

OE).

and

and C* generally depend on the prior history and are time-dependent


In

(1981b) defined

order

to

remove

the

dependence

OE

and

on the prior history, Riedel

334

25.

Primary-Creep Effects

(25.6;

where the subscript 'h' indicates the suitability of


Obviously,

C~

C~

for

hardening

creep.

is path-independent since it differs from J or C* only by time-

dependent factors. It reduces to C* if p=O and it has the desired


depending on the current load only. In
related to the current net section stress by

property

the notation of Chapter 20,

01

C~

iE

(25.7)
Since under proportional loading
behaves

as

if

it

were

the

nonlinear

material
viscous,

characterized
by

coefficient in eq. (25.5) and n replaced by m. In terms of

[B 1 (1 +p) ] 7 (1 +p) 1m r

25.1.2

eq.

(25.1)

the crack-tip field must be the

HRR-field, eq. (20.16), with the coefficient B replaced

----,rrrCT~:;:;::,r-_
1

by

the

correspondine

C~:

) 1 / (m+1) ;;. j (a, m)

(25.8)

Growth of a primary-creep zone in an elastic field

Now we consider step loading of a cracked body which consists of material whicn
can deform elastically and by primary creep
+ (pr)
Eij
,

where the primary creep rate is given by eq.


m > 1.

The

instantaneous

response

of

(25.9)
(25.1)

such

with

stress

exponent

a material is elastic. In close

analogy to elastic/nonlinear viscous materials, stress redistribution occurs by


the growth of a primary-creep zone, which now expands according to
(25.10)
for short times. The near-tip stress
(25.8), with C~ replaced by

field

is

an

HRR-type

field

like

eq.

25.1

335

Strain-Hardening Model for Primary Creep

K2 (1_,,2) IE

h [(~+1)t]11(1+P)

(25.11)

in the short-time limit and approaching eq. (25.8) in the long-time limit.
characteristic

time

for

stress

redistribution

is

The

obtained by equating the

values in the two limiting cases:

1 +p

(25.12)

This characteristic time together with other characteristic times will be

dis-

played on a load parameter map in Section 25.1.4. Equation (25.12) also describes the transition from an elastic-plastic state to primary creep provided that
Ki(1-,,2)/E is replaced by the J-integral (Riedel, 1981b).
25.1.3

Growth of a secondary-creep zone in a primary-creep field

Now we consider the case that the whole specimen creeps in


that

elastic

primary

creep,

so

strain can be neglected, but near the crack tip the material has

already reached the secondary stage of

creep.

To

model

this

situation

the

material law is assumed to be


(3/2) B e: -p m(1+p)-1 I + (3/2) B n-1 I
1
e D"e
D"ij
D"e
D"ij'
Upon step loading, the primary-creep term first dominates
field

like

eq.

(25.8).

again

has

crack-tip

secondary-creep

zone.

This

boundary-layer

similarity solutions. From their form it is clear that the

secondary-creep zone
r cr

For short times, this represents the remote boundary

condition for the evolution of


problem

giving

(25.13)

m~st

grow according to

C* [B- 1/ (m+1) B1+p t P j(m+1)/[(n-m)(1+ p )].


h
1

(25.14 )

Since the near-tip stress field must be an HRR-type field compatible

with

the

similarity solution, it must have the form


a(n,m,p) Ch*
1/(
1)
)
n+
(p+1)BI tP/(l+p)r
n

The factor a cannot be

determined

from

a..(a,n).

(25.15)

lJ

similarity

arguments

alone.

As

approximation, we assume that the J-Integral is path-independent, which gives

an

25.

336

a(n,m,p)

Primary-Creep Effects

(n+p+1)/(n+1).

(25.16)

The assumpt10n that C* be path-1ndependent 1s equally justified as an


mation;
and n

this

leads

to a(n,m,p)

9 and we proceed using a

approxi

1. The two estimates differ by 20% if p

1.

For long times, the secondary-creep zone spreads across the whole

specimen

so

that nonlinear viscous behavior 1s approached with a crack-tip field character


ized by C*. As in previous similar cases, a characteristic time is found to be
t2 = (

C*
h
)(p+1)/P.
(1+p)C*

(25.17)

An interpolation formula between the primary-creep regime and the steady


is

state

obtained if in the HRR-f1eld, eq. (20.16), C* is multiplied by the express-

ion 1+(t 2 /t)P/(1+ P ). This reproduces eq. (25.15) in the short-time limit. If,
additionally, the elastic-plastic response is to be taken into account, the
following interpolation formula for the HRR-field amplitude is suggested:
C(t)

25.1.4

(25.18)

Summary and introduction of a load parameter map

Having included elastic strain, plastic strain,


creep,

primary

creep

and

secondary

it appears necessary to summarize the results. This can be done conven-

iently on a load parameter map, which is a diagram with time

on

the

vertical

axis and 0net on the horizontal axis. In this plane, the areas are indicated in
which each of the four deformation mechanisms considered so far determines
specimen

behavior.

the

Associated with each deformation mechanism is a load para-

meter which describes creep crack

growth

macroscopically

in

the

respective

regime. The regimes are separated by the characteristic times t1 and t 2


Figure 25.1a schematically shows an

example.

In

the

low

stress/short

time

regime, the elastic response dominates and therefore KI is applicable. At high


stresses (markedly beyond the ASTM E-399 criterion), the specimen becomes fully
plastic, which requires the use of the J-integral, until, at very high stresses
crack-tip blunting bounds the regime of J-controlled crack growth

(cf. Section)

25.1

Strain-Hardening Model for Primary Creep

337

/oganetFig. 25.1. Load parameter maps


for two hypothetical sets of material parameters.

22.1 and the ASTM E-813 rule). The primary-creep regime, which is labeled by
-(m+1)(1+p)
C~,
is separated from the elastic regime by t1 anet
Ceq. (25.12)J,
while after the time t 2 a~:~n)(p+1)/P Ceq. (25.17)J, steady-state creep
starts

to

dominate. At higher stresses, where no primary-creep regime eXists,

the secondary-creep zone catches up with the primary-creep zone while both
still

small,

and

the

further

evolution

effectively in an elastic surrounding. The

are

of the secondary-creep zone ensues


transition

to

steady-state

creep

then occurs with the characteristic time t1 a~~~-1) Ceq. (23.15)J. At even
higher stresses the transition from instantaneous plasticity to secondary creep
1/N-n Ceq. (24.9)J. At long times, crack-tip blunting
occurs at around ..u1 anet
bounds the range of validity of C* unless fracture of the specimen intervenes.
n+1
Combining eqs. (22.1) and (22.3) and recalling that C* anet shows that the
limitation set by blunting varies as tb a~~t.
The load parameter map is a deformation map of cracked bodies, but
any

sense

related

is

not

in

to the fracture-mechanism maps described in Chapter 2. The

precise position of the lines representing the characteristic times depends

on

the material parameters, E, B, Bo ' B1 , and on the geometrical functions


g1(a/W,n), but not on the absolute speCimen size. For example, the regime
dominated by C~ will be larger in a material with pronounced primary creep,
i.e., with a large B1 -value. Figure 25.1b shows such a situation.

25.

338

Primary-Creep Effects

25.2 Hardening/Recovery Model for Primary Creep


25.2.1

The constitutive equations

An alternative description of
superior

to

the

previous

primary
one

dates back to Bailey (1926) and

creep

effects,

which

appears

to

be

under loading conditions like cyclic loading,


Orowan

(1946).

They

introduce

an

internal

variable, called internal stress ~ .. , which counteracts the applied stress and
IJ
which increases by strain hardening and decreases by recovery. Following Kubo
(1983),

whose

formulation

is

based

on work of Pugh and Robinson (1978) and

Robinson (1978), we describe these processes by


E .

( 3/2)

n-1
C1 'e
'ij

..

(2/3)

C2

IJ

IJ

IJ

(25.19)

-Ii - C n-Ii-1
3 (1e

~e

..

(25.20)

IJ

(25.21 )
C1 , C2 , C3 , n and Ii are material parameters. Table 25.1 shows a few examples,
which are taken from Kubo (1983) who refers to Mitra and McLean (1966). Delph
(1980) uses a slightly more general formulation than eqs.

(25.19)

to

(25.21)

and reports the material parameters for 2'!.Cr-1Mo steel at 566 o C.


From eqs. (25.19) to (25.21) it is clear that (1 .. and , .. are deviators with no
IJ
IJ
hydrostatic component. The first term on the right-hand side of eq. (25.20) represents strain hardening, since the internal

stress

while

of the internal stress which was

the

second

term

describes

recovery

increases

with

strain,

previously built up.


Equations (25.19) and (25.20) cannot be integrated in closed form even for uni-

MateriB.l:
n

Ni

Al

Zn

4.7

4.6

1.7

1.1

1.5

Table 25.1. Material parameters n and Ii after Kubo (1983).

25.2

Hardening/Recovery Model for Primary Creep

339

axial tension under constant stress. Figure 25.2b schematically shows the shape
of

the

creep curve predicted from the following arguments. At time t = 0, the

internal stress is zero and the strain rate is E = C1 0 The


from a series expansion solution, increases initially as

internal

stress,

(25.22)
For long times,

e saturates to
(25.23)

and the strain rate decreases to the steady-state value


(25.24)
where
(25.25)
A characteristic time for the duration of the
equating the expressions
eqs. (25.22) and (25.23).

+5

~r

....e

can

be

obtained
Q

.s:
....e

11)

11)

o
c: 1

norm. time BrY't/e, ___

by

e from

'"w
.~

transient

for the short-time, and long-time values of

time ---

Fig. 25.2. Creep curves.(~a))Strair-hardening model, eq. (25.13); E1 is an


abbreviation for [B om +p /Bon] /p. (b) Hardening/recovery model, eqs.
(25.19) and (25.20); schematic. Dashed lines: primary creep alone.

340

25.2.2

25.

Primary-Creep Effects

Solutions for crack geometries

As the creep curve in Fig. 25.2 shows, the initial response of a


covery

material

can

be

approximated by the viscous law ~

hardening/re-

c,on. Therefore,

following Kubo ('983), we note that the initial stress field must be that of

nonlinear viscous material. In particular, the crack-tip field must be the HRRfield, eq. (20.'6), with B replaced by C,. Depending on the stress experienced,
different

portions of a cracked specimen approach the long-time, steady-state,

creep stage within different periods of time. If n-' > B (which is usually
case

for

metals)

the

the

transition is accomplished first at the crack tip. For

short times, the transition can be

described

by

similarity

solutions.

From

their form it follows that the steady-state creep zone must grow according to
r

ss

(C*/C,) (C C t)(n+')/(n- p -')


, 2

(25.26)

with a factor of proportionality which depends on the precise definition chosen


for rss and on n, Sand H" which was defined in eq. (25.25). The crack-tip
field well inside this zone is again the HRR-field, but now with B replaced by
C,/H, and with an unspecified factor which cannot be determined from similarity
arguments alone. However, if J or C* are assumed to be path independent, the
factor is found to be unity in either case.
For long

times,

steady-state

creep

prevails

in

the

whole

specimen.

The

crack-tip field is still given by eq. (20.,6) with B replaced by C,/H,.


Hence, in the hardening/recovery model, the crack tip field has the

same

form

in the short-time limit and in the long-time limit, viz.,

(25.27)
However, the C*-integral becomes time-dependent even

for

constant

load.

The

short- and long-time values of C* are given by eq. (20.7) with B replaced by C,
and C,/H"

respectively:
C*(t=O)

n+'
a C,Onet g,(a/W,n),
(25.28)

C*( t="')

25.2

Hardening/Recovery Model for Primary Creep

341

At intermediate times. C* is not a well-defined parameter. But as an approximation one can continue to assume its path-independence and to measure it according to C* = g20netA. This will probably provide an estimate for the strength of
the HRR-field. C(t). Theoretically. Kubo (1983) concludes that C(t) must
decrease for short times according to
1 - y[~-1-B C C t]1/(B+1)
net
1 2

C(t)/C*(t=O)

(25.29)

where Y is an unknown numerical factor. Equating this with the long-time value.
eq. (25.28). gives a characteristic time which varies with stress as
(25.30)
This characteristic time describes the transition from primary to

steady-state

creep of the whole specimen. It corresponds to the time t2 given in eq. (25.17)
which was. however. derived from a different model of primary creep.
25.2.3

Elasticity effects and load parameter map

If elastic strains are added to the material law. eq. (25.19). the sequence
events

after

step load is as follows: At short times. a primary-creep zone

grows within the elastic singular field. The primary-creep zone


relaxing

of

size

and

the

HRR-field are the same as in elastic/nonlinear viscous materials with

B replaced by C1 in all formulas of Section 23.1. Within the


zone. a secondary. steady-state creep zone grows according to

primary

creep

(25.31 )
Since the steady-state zone increases with a higher power of time than does the
2l (n-1)
primary zone. rpr t
two different cases are possible: first. the
steady-state zone may catch up with the primary zone while both are still small
compared with the ligament width. Primary creep then never dominates the
specimen behavior and the transient occurs effectively from elastic behavior to
steady-state

creep.

Second.

the

primary

zone

may

spread across the whole

specimen while the steady-state zone is still small.


The corresponding load parameter map which describes the evolution
ing/recovery

creep

in

cracked

specimen

is

shown

stresses. primary creep never dominates and the transition

in

of

harden-

Fig. 25.3. At low


time

from

elastic

25.

342

response

Primary-Creep Effects

to steady-state creep is t1 as given eq. (23.15). At higher stresses,

the transition from elastic behavior to primary creep occurs also at t1


calculated with a C*-value which is greater by the factor H1 At t = t 2 ,
(25.30), the whole specimen approaches steady-state creep.

but
eq.

Comparison of the load parameter maps in Figs. 25.1 and

the

25.3

shows

that

shape of the region occupied by primary creep depends on the constitutive model
chosen.

25.3 Analysis of an Experiment in the Thmsition Range


Between J, C~and C*
<In this section, results of Detampel (1986) on 21/~Cr-1Mo steel tested at 540 0 c
are

reported.

In

order

to obtain measurable creep crack growth within a few

months in this material, like in other similar steels, it is necessary to

load

the specimens beyond the elasticity limit so that instantaneous plasticity must
be taken into account. Further, it turned out that primary creep played an even
greater

role

during

the

initial transient. In the analysis described below,

primary creep was modeled by the strain hardening law,

eq.

(25.1), with the


29
following estimate for the material parameters: B1 = 4.5'10- , m n/3, p = 2,
while the steady-state creep parameters were B = 1.3'10- 24 , n = 7.3 (all
dimensions in MPa and s).
As an example, a test on a CT1-specimen having 20% side grooves and an initial

toga

net

___

Fig. 25.3. Load parameter map based onhardening/recovery model.

25.2

Analysis of an Experiment

343

crack length a o = 25 mm is described. The specimen was loaded to P = 20.12 kN


which resulted in a lifetime of t f = 596 h. The instantaneous displacement at a
distance z

12.5 mm behind the

load

line

was

A = 492 pm,

while

purely

elastic response would have led to Ael = 269 ~m. From the Merkle-Gorten (1974)
formula, the J-integral immediately after load application was obtained as
J = 18.7 kJ/m 2 , while its linear elastic (or small-scale yielding) value would

7 kJ/m 2 From eq. (20.7) the G*-integral at the beginning of


7.2.10- 2 W/m 2 Equation (24.9) then gives the
was calculated as G*

have been J el

the test
characteristic
secondary

time

creep

for

as tl

the

transition

from

elastic-plastic

creep range intervenes between the J- and G*-controlled


sponds

to

the

as t2

to

regimes.

This

corre-

case shown by the load parameter map in Fig. 25.1b. The trans-

ition time from primary to secondary creep is obtained from


(25.7)

behavior

8.7 h. However, in the example considered, a primary-

eqs.

(25.17)

and

83 h. This is indeed not negligible compared to the lifetime of

the specimen, 596 h. In fact, the measured displacement rate exhibited a transient

behavior

like that predicted in Fig. 23.6. A steady state was approached

after about 50 h Which is close to the calculated value of t 2 . (Of course,


displacement rate increased again later in the test due to crack growth).

the

The observed crack growth behavior is shown in Fig. 25.4. During the transient,
the

crack

grows rapidly and decelerates subsequently until after some 100 h a

minimum growth rate is reached.

The

figure

alsa

shows

several

curves, which are derived from eq. (21.6) (taking the limit Xc
ing the Gamma functions). Equation (21.6) is based on a
erion

and

on

theoretical

0, i.e. delet-

critical-strain

crit-

purelw nonlinear viscous deformation of the specimen. It can be

generalized to include elastic-plastic and primary-creep transients by

replac-

ing G* with G(t) as was already pOinted out in Section 23.3.1; G(t) is given in
eq. (25.18). If tl and t2 are approximated by constant values pertaining to the
beginning

of the test, and if G* is inserted from eq. (20.7), the crack growth

law given in eq. (21.6) can be integrated by separation of the variables, a and
t.

The

resulting integrals must be evaluated numerically. In a later section,

the procedure is

describ~d

in greater detail for slightly more general

condit-

ions [eqs. (27.19) to.(27.21)].


This integration was
different

carried

assumptions.

which means that G(t)

First,

out

for

only

plane-strain
secondary

conditions

under

three

creep was taken into account,

G*. This calculation leads to the dashed line

in

Fig.

25.4. The critical strain, which appears in eq. (21.6) was chosen as f = 1.82%
in order to adjust the calculated lifetime to the observed one. In a second

25.

344

i
E
E

.!:
c

33
32
31
30
29
28
27
26
25
24
0

Primary-Creep Effects

i/4 Cr-1Mo steel


540C
CT1-specimen
P = 20.12 kN

."..

":;",:;:,:;,,,::::,:::,:::,:::,';:,';:'::;

100 200 300 400 500 600


tin h---+

Fig. 25.4. Crack length vs. time. Circles: measured data; dashed line:
calculated with secondary creep only; dotted line: elastic-plastic strain
included; solid line: primary creep and elastic-plastic strain included.

step, elastic-plastic material response (but no primary creep) was

taken

into

account, which means that C(t) is taken from eq. (23.16) with t1 = 8.7 h. Numerical integration of eq. (21.6) then leads to the dotted line in Fig. 25.4.
Here,

Ef =

2.05% was chosen to adjust the theoretical to the measured lifetime.

The dotted line still underestimates the magnitude of


Finally,

primary

creep

the

initial

transient.

was taken into account in addition to elastic-plastic

strains and secondary creep, i.e. C(t) was taken

from

eq.

(25.18).

If

this

choice for C(t) is inserted for C* in eq. (21.6), one obtains the solid line in
Fig. 25.4, where

~f

= 3.28% was taken to fit the lifetime. Now the theoretical

result comes much closer to the experimental data.


The same set of data is shown in Fig. 25.5 after
crack

length.

If

although C* is

no~

the

time-differentiation

of the
aBon+1
net g1 ,
obtains Fig.

measured growth rate is plotted against C*

a valid parameter during the

tranSient,

one

25.5a. If Ct = ~ 0net g2 defined in eq. (23.19) is plotted on the abscissa with


Aas measured, the two branches of the curve for decreasing and increasing
growth

rates fall closely together as shown in Fig. 25.5b. A similar result is

obtained if

a is

plotted against C(t) calculated as in

eq.

(25.18).

At

long

times, all definitions used in Fig. 25.5 merge into the steady-state value, C*.

25.2

Analysis of an Experiment

345

b)

a)

10-'

'/
V

c)

...lL

n+l

-8

11\

"E

10

.!;

oCJ

2~Cr-l""o

10-9

540C
CTl
P=20.12kN

10-10
0.D1

al

10

100

C=aBa::;:g, -

,/

al

10

al

100

10

100

crt) -

C1=.tiq.,., g2 - inW/m 2

FIg. 25.5. Measured crack growth rates plotted against different parameters

all of which approach C* at long times. Dashed curves: theoretical results.

The dashed lines in-Fig. 25.5 are theoretical curves derived in Chapter 21.
Fig.

25.5b,

only

In

the slope n/(n+1) is shown which is typical for strain-con-

trolled creep crack growth. The dashed line in Fig. 25.5c represents

the

com-

plete solution of eq. (21.6), which was already shown as the solid line in Fig.
25.4. It includes primary-creep effects by calculating C(t) from
Apparently, after fitting the critical strain as f

eq.

(25.18).

3.28%, the model predicts

the measured growth rate well apart from a slight deviation in the shape of the
cur-ve.
for Xc

Possibly,

the agreement could be improved if the calculation

~ere

done

> o.

As was already mentioned, the parameter Ct is very similar to the parameter


which was introduced by Saxena and Landes (1985) and by Saxena (1986) and which
was also denoted by Ct. The justification for its use is that it is
approximate

the

for the special case of elastic/nonlinear viscous material. Its


vantage

to

C(t)

contribute

itself

to

the

is

practical

ad-

Ct can easily be measured during a test,


whereas C(t) must be calculated taking into account all deformation mechanisms
which

over

likely

crack-tip parameter C(t) as was pOinted out in Section 23.1.6


that

deformation of the specimen. Equation (25.18) is an

example for how C(t) can be calculated, provided that the


entering into the equations for t1 and t2 are available.

material

parameters

26 Diffusion Creep

26.1 Constitutive Law


Diffusion creep predominates at very low stresses in metals, while in
it

is

the

common

high-temperature

ceramics

deformation mode. In the diffusion-creep

range, the deformation behavior is usually approximated as linear viscous.


besides

diffusion

creep,

If,

elastic deformation and dislocation creep are taken

into account, the material law becomes:


(26.1 )

The material parameter Bd is an abbreviation for (cf. Section 1.3)


(26.2)

where Nabarro-Hering creep is described by the first term in parentheses, while


Coble

creep

is

represented by the second term. The additive superposition of

the two terms is understood as an approximation.

26.2 The Effect of Diffusion Creep on the Deformation Fields


in Cracked Bodies
If linear viscous creep alone predominates, the stress field in a cracked

body

is identical with a
analogy. The crack-tip

linear elastic stress field due to the viscous/elastic


field can be expressed equivalently by eq. (23.2),
1/2
; KI and C* become equivalent
0ij a KI/Ir, and byeq. (20.16), 0ij a (C*/r)
in linear viscous materials and are then related by
C*

This is the viscous analogue to the formula for

(26.3)

the

J-integral

under

linear

26.1

347

Deformation Fields

K2I (l-v 2 )/E, with v = 1/2 and l/E replaced by Bd

elastic conditions, J

If power-law creep is admitted in addition to linear viscous creep, the


ial

law

in uniaxial tension is ~ = Bdo

mater-

Ban. Due to the elastic/viscous ana-

logy, this is entirely analogous to an elastic/plastic problem if plasticity is


modeled

as

deformation

plasticity. At small loads, the response of a cracked

body is determined by the linear term except in a

small

power-law-creep

zone

near the crack tip whose size is of the order


(26.4)
where the subscript is to denote nonlinear creep. This estimate is obtained
equating the strain rates Bdo and Ban with a

by

KI/Ir. For increasing loads, the

power-law-creep zone grows until, when rnc = a, power-law creep determines


whole specimen response. In terms of net section stress this occurs when

the

(26.5)
This equation is represented by the vertical line on

the

load

p'arameter

shown in Fig. 26.1.

c or K]

t....

(linear
(nonlinear
viscous)
viscous)

log (JnetFig. 26.1. Load parameter map showing the regime of diffusion creep
(upper left regime).

map

Diffusion

26.

348

Cree~

The transition time from the initial elastic behavior of a cracked specimen
nonlinear

viscous

creep

at

tc

high loads is t1 as given in eq. (23.15). At low

stresses, the transition from elastic behavior to linear viscous creep involves
no stress redistribution, apart from the 033-component which depends on
Possion's ration in plane strain. In the elastic regime is typically v = 0.3
while

diffusion

creep

is

incompressible, i.e. v = 0.5. A transition time is

reasonably defined as the time at which creep strain becomes equal

to

elastic

strain,
(26.6)
which coincides with the limit of eq. (23.15) for n = 1, apart

from

numerical

factors. Equation (26.6) is represented by the horizontal line in Fig. 26.1.

26.3 Crack Growth Rates Assuming a Critical-Strain Criterion


If the crack grows in a linear viscous material

such

that,

at

distance ahead of its tip, the strain attains its critical value,

Ef

structural
, the crack

growth rate is obtained from eq. (21.6) by setting n = 1:

(26.7)

Here, numerical factors were omitted and the limit a-a o Xc was taken. As eq.
(26.7) shows, the crack growth rate increases in linear proportion to Kr in the
present case.
If the material develops a
field,

the

power-law-creep

zone

within

the

linear

ViSCOUE

crack growth rate is given by eq. (21.6) with n > 1 as long as the

crack grows within the nonlinear zone. With

C*

from

eq.

(26.3),

the

crack

growth rate becomes


a

K2 )n/(n+l)/
[ B(a-a )]1/(n+l) (
o
Bd I
Ef

Here, the crack growth rate increases in proportion to

(26.8)

Ki n /(n+l).

27 A Damage Mechanics Approach to


Creep Crack Growth

27.1 Introduction
In the preceding chapters, the deformation behavior of materials was
by

elastic,

described

plastic and creep strains. Whenever damage was taken into account

in order to model creep crack growth, it was assumed that the stress distribution remained unaffected by damage. This assumption is now dropped.
27.1.1

The constitutive model

The constitutive model to be employed in the


model

sequel

is

the

phenomenological

of Kachanov and Rabotnov, which was already introduced in Section 2.3.4.

It describes the stress/strain-rate response using an internal variable


the

parameter w. This model is to be applied to crack geometries here.

damage

Therefore, the generalization of the equations to multiaxial states


is

needed.

(1~84).

both

Here

we

employ

the

maximum

pr.incipal

of

stress

formulation proposed by Hayhurst and Leckie

The kinetic law for the evolution of the damage

the

called

stress,

aI'

and

parameter

depends

on

on the equivalent stress, ae ,

according to
(27.1)

The material parameters D,X and


the

determination

of

the

are the same as in the uniaxial

material

parameter

case,

while

requires creep rupture tests

under at least two different stress states such as tension and torsion.
The damage parameter itself is not meant to be measurable. Only its
the stress/strain-rate relation is detectable:

effect

on

27.

350

A Damage Mechanics Approach to Creep Crack Growth

(27.2)

IJ

that means the creep rate increases sharply, or the stress drops to zero, if
approaches

under

stress-,

or

strain-controlled conditions, respectively.

Elastic strains and primary creep will be included in later sections.


Table 27.1 shows the material parameters pertinent to eq.
The

data

for

copper

(27.1)

and

(27.2).

and for the aluminum alloy were taken from work of Hay-

hurst, Brown and Morrison (1984), while the data on ferritic

steels

were

ob-

tained from creep curves provided by Bendick and Weber (1984). Since the latter
authors carried out uniaxial tension tests only,

could not be determined.

Recently, attempts were undertaken to develop constitutive models for


ing

materials

cavitat-

based on the physical laws for cavity growth (Hutchinson, 1983,

Tvergaard, 1984). Models of that type were described in Sections 12.5 to


However,

these

12.7.

model-based descriptions are not yet in a shape to replace the

empirical description proposed by Kachanov.


27.1.2

The relation between fracture mechanics and damage mechanics

There are several interesting pOints which should be noted before the
of

creep

crack

growth

is

described

analysis

in detail. First, the solutions of the

damage mechanics equations automatically contain crack growth. Whereever w = 1,


the material has failed and transmits no tractions. A pre-existing crack starts
growing immediately upon load application since the failure

criterion,

is reached instantaneously directly at the crack tip.

E/[GPaJ
Al
Cu
Stl
St2

60
66
150
150

B/[MPa

-n

Is]

3.2'10- 21
3.6'10- 10
1.3'10- 24
2.7'10- 31

D/[MPa-X/s]

6.9
3.0
7.3
10

5.0'10- 18
1 .7'10- 7
7.5'10- 21
2.0.10- 23

6.5
1.2
6.2
7.0

<P

9.5
3.8
5.5
12.3

Table 27.1. Material parameters for an aluminum alloy at 210 o C,


for copper at 250 o C, for 2'/ 4 Cr-1Mo steel (St 1 ) at 540 0 C and for
1 12Cr-1/2Mo-l 14V steel having a service history (St2) at 540 o C.

0
0.7

w = 1,

27.1

Introduction

351

Further, it is important to note that a 'small-scale damage' limit can

reason-

ably be defined. Damage is developed rapidly near the crack tip in a zone which
may be called the 'process zone' and which will be
later.

As

long

defined

more

specifically

as the growing process zone is small enough, a boundary layer

problem can be formulated by prescribing the nonlinear viscous HRR-field (without

damage)

as

the

remote boundary condition for the solution of the damage

mechanics equations. The existence of a small-scale damage limit has


practical

consequences.

Only

in

this

important

limiting case, the fracture mechanics

approach based on C* remains valid. Outside the small-scale damage

limit,

the

process zone becomes so large to perturb the singular fields which validate the
use of

the

respective

description

load

parameters.

Therefore,

no

fracture

mechanics

of creep crack growth based on macroscopic load parameters appears

to be possible outside the small-scale damage limit.


Now the questions arise whether the small-scale damage limit
validity

which

is

sufficiently

what can be done outside its range


question

was

given

by

has

range

of

large to use C* in practical situations, and


of

Hayhurst,

validity.

Brown

The

answer

to

the

second

and Morrison (1984). They carry out

finite element analyses of cracked specimens based on the full damage mechanics
equations.

However,

this

procedure is not only computationally difficult and

time-consuming since every configuration requires

separate

finite

element

analysis, but it may also lead to serious error. The continuum damage equations
do not include any specific corrosive effects which may
creep

crack

growth

enhance

the

element solution of the continuum damage equations would drastically


mate

the

rate

of

in several structural materials. In such a case, a finite


overesti-

lifetime of a cracked component. The fracture mechanics approach, on

the other hand, includes the corrosive effects if the laboratory tests are done
in

the same environment in which the component operates. Crack growth data are

merely transferred from the laboratory test specimen to the component using the
appropriate

load parameter. Considering these advantages of the fracture mech-

anics approach it remains to explore the range of validity of

the

small-scale

damage approximation. This is one of the main purposes of this chapter.


The following analysis is done for two-dimensional problems (plane
plane

strain

and

stress). If specific numerical results are given, they are understood to

be for plane strain. The geometry and the coordinate systems are the same as in
earlier
(1985b).

sections

(cf.

Fig.

20.1).

The

presentation follows work of Riedel

27.

352

A Damage Mechanics Approach to Creep Crack Growth

27.2 Small-Scale Damage in Extensively Creeping Specimens


27.2.1

Similarity solutions

The instantaneous response of a


(27.2)

material

characterized

by

eqs.

(27.1)

application is concentrated in the vicinity of the crack tip and can be


zed

and

is nonlinear viscous. The evolution of damage at short times after load


analy-

under the remote boundary condition that the stress field at infinity must

approach the HRR-field given in eq. (20.16). If the


initial

HRR-field

represents

the

condition at t = 0 and the remote boundary condition, dimensional ana-

lysis shows that the damage mechanics equations must have similarity

solutions

of the form
(Dt)-l IX I: .. (R, e)

(27.3:

w(R,e)

(27 .4;

(Br/C*) (Dt)-(n+1) / x

(27.5:

ij

IJ

with

The dimensionless functions I: .. and ware


validity

of

these

IJ

solutions

as

yet

unknown.

Of

course,

the

can be verified by insertion into the governing

equations and boundary conditions.


27.2.2

Crack growth rates

Due to the form of the similarity coordinate R in eq. (27.5), the

contours

01

constant stress or constant damage expand around the crack tip according to
(n+1)/X In particular, the crack tip, which is characterized by w = 1,
r t
moves according to
6a

a(n,X,.,K) (C*/B)(Dt)(n+1)/x,

(27.6)

where 6a = a-a o is the crack growth increment and a


initial

and a o are current and


crack length, respectively. The dimensionless factor a(n,x,.,K) cannot

be determined from similarity arguments. Its magnitude


Section

27.2.3.

will

time by the crack growth increment through eq. (27.6) gives


rate:

be

estimated

in

Differentiating eq. (27.6) with respect to time and replacing


the

crack

growth

27.2

Small-Scale Damage

353

n+1

(27.7)

It is interesting to compare the so calculated crack growth rate with that

ob-

tained in Chapter 21. There, the stress field had been assumed to be the undisturbed HRR-field moving with the crack tip, an assumption which is not
in

adopted

the damage mechanics description. Further, the conditions for local failure

at the crack tip were derived from cavity growth laws in Chapter 21, while
the

now

formal damage parameter description is employed. In spite of these differ-

ences, the results are remarkably similar.


damage

mechanics

To

realize

this,

structural length xc. Howeyer, this difference to Chapter 21


limit

of

large

note

~a/xc.

the

vanishes

criterion.

In

in

the

Further the theories can only be comparable if an equi-

valent failure criterion is used. Equation (21.6) is derived from


strain

that

equations contain no characteristic length comparable to the

the

critical-

Kachanov model, the failure criterion reduces to a

critical-strain criterion if X = n. In this case, eqs. (27.7) and


dict the same dependence of a on C* and on

~a.

(21.6)

pre-

Further, if D is replaced by the

strain to failure using eq. (2.12) with X = n, it becomes evident that the

two

models also predict the same dependence of


on E f and on B. The only possible
difference is a numerical factor. It will be seen shortly that the factor given
by

the

older

theory corresponds to the upper bound estimate for the factor a

derived below if, at the same time, the strain to failure is

replaced

by

the

which

is,

Monkman-Grant product in the older theory.


27.2.3

Approximate and numerical methods in small-scale damage

A very simple approximate way to determine the


however,

compatible

with

the

assume that the stress field is

required
the

factor

a(n,x,~,K),

self-similarity of the fields, is to

undisturbed

HRR-field

centered

at

the

initial crack tip position. For such a time-independent stress, the kinetic law
for ~ is easy to integrate. It is found that the crack tip moves in

accordance

with the form of the similarity coordinate, and the factor a is then given by
(27.8)
Since the HRR-field was assumed to remain centered at

the

original

crack-tip

position, eq. (27.8) is expected to be a lower bound.


An upper bound is obtained if the HRR-field is attached to the moving crack tip

27.

354

A Damage Mechanics Approach to Creep Crack Growth

but is otherwise undisturbed. If one lets the crack tip move in accord with eq.
(27.6), a is obtained as:
(27.9)
with s

x/(n+1). Incidentally, this gives the same growth rate as that derived

in eq. (21.6) if X = n,

Ef

= BID and Aa xc.

An independent estimate for a can be obtained from the finite element


of

the

small-scale

damage

limit

analysis

carried out by Riedel (1985b). Usually the

contour w = 1, which describes the crack, does not spread along the line
directly ahead of the crack. For the special case n

X = 5,

= ,

a =0

= 1, for ex-

ample, the calculation predicts a plane-strain crack to branch along planes inclined

by

24 to the symmetry plane. This does not necessarily imply that

a real crack will actually branch. The crack may also follow the direction +24 0
along

certain

scopically

a=

along

portions

meters, the

the

crack

front,

and -24 0 along others. Macro-

average, the crack surface may remain flat. If Aa is measured

on~he

24 0 ,

of

the

upper-

= 37.

factor a is found to be a
and

lower-bound

estimates

For the same set of para-

give

217

and

29.8,

respectively.
27.2.4

The process zone

Somewhat arbitrarily, the process zone is defined as the zone within which
equivalent

strain

rate

is

at

the

least doubled by damage compared to undamaged

viscous material. An estimate for the size of the so defined

process

zone

is

obtained by calculating the damage from eq. (27.1) and the strain rate from eq.
(27.2) using the undisturbed HRR-field centered at the original crack tip. Then
the process zone size normalized by the growth increment is found to be
(27.10)
For the material parameters of copper and of the aluminum alloy shown in
27.1,

this

ratio

is 2.6 and 0.69, respectively. For n

= ,

Table

X = 5, the ratio

becomes 0.99.
Figure 27.1 illustrates that this definition of the process zone is reasonable.
The

figure

shows

finite

element

results for the stress ahead of the crack,

which has grown from 0 to 1 in the normalized

units used in Fig. 27.1.

Hence,

27.2

355

Small-Scale Damage

t
~

'"

b'"

:;::..

11)
11)

05

11)

g
c::

,,

/",HRR

" '1"'--__
I
I
I
I
I
I

CI.I

I
I

I
I

1p-!

2
r/Lla -

Fig. 27.1. Stress distribution in the process zone ahead of a growing crack.
Finite element results for n = ~ = X = 5, K = 1, e = 24 0 , after Riedel (1985b).

the stress is relaxed to zero there.


according

Also

shown

is

the

process

zone

size

to eq. (27.10). Obviously, rp is a measure for the zone within which

the stress deviates markedly from the undisturbed HRR-field.

27.3 The Range of Validity of the Small-Scale Damage Approximation


in Extensively Creeping Specimens
In analogy to the limitation caused
damage

by

crack-tip

blunting,

the

small-scale

approximation is valid only as long as the process zone is smaller than

a certain fraction, denoted by 2.5/M, of the crack length and ligament width
rp < 2.5 aiM.

(27.11)

Otherwise the HRR-field has no finite range of approximate validity between the
process zone and the outer specimen surface. Then C* cannot characterize cracktip fields uniquely. The factor M must have values of around
geometries

including

the

compact specimen and M

= 25

for

bend

200 for the center-cracked

plate in plane-strain tension. The condition for the small-scale damage approximation

to

be

valid,

eq.

(27.11),

can be expressed as a condition for the

relative crack growth increment using eq. (27.10):


(27.12)

27.

356

A Damage Mechanics Approach to Creep Crack Growth

With M = 25 and n = $ = X = 5, eq (27.12) gives the allowable amount


growth

of

crack

as 10% of the crack length. At first sight, this might look like a very

restrictive condition. But the condition for the allowable crack growth

incre-

ment can alternately be expressed as the requirement that the time must not exceed a certain fraction of the lifetime, t f , of the cracked specimen, and it is
found experimentally and theoretically (Section 27.6), that 10% crack extension
correspond to a substantial fraction of the total lifetime. On a load parameter
map, the regime of large process zones Is represented by the hatched band shown
in Fig. 27.2. Although C* should no longer be valid in this regime, the experiments

reported

in Chapter 21 show a good correlation of the crack growth rate

with the C*-integral up to greater fractions of crack extension than 10% of the
ligament. This agreement might be fortuitous, but it may also indicate that the
process zone Is not quite as large as predicted by the Kachanov equations.

27.4 The Evolution of Damage and Crack Growth


for Small-Scale Creep
If elastic strain rates are added to the creep rate in eq. (27.2), two limiting
ranges

of

behavior

can

be

distinguished

which

are

described

subsections below. In brittle materials creep strains can be


where

w ~ 1.

In

other

zone between the process

in the two

neglected

except

words, in this first limiting case, there is no creep


zone and

the remote

elastic field.

In more ductile

....

-8'
J
log (/netFig. 27.2. Load parameter map showing the limitation
to C* by too large process zones (hatched area).

21.4

Damage and Crack Growth in Small-Scale Creep

351

materials, on the other hand, the creep zone grows faster than the process zone
so

that final failure occurs under extensive creep conditions, but the initial

stages of crack growth may be influenced by the elastic transient.


21.4.1

Crack grows faster than creep zone

In brittle materials, a crack may grow so fast that no appreciably sized


zone develops before the specimen fractures. This situation,
growth, i.e. a = const, was already analyzed
critical-strain

criterion,

in

Section

for

creep

steady-state

23.2.3

based

on

rather than on the damage mechanics equations. The

r/r 1 , and stress, Lij , defined in eqs. (23.27) and


(23.28) can again be used to non-dimensionalize the stress/strain-rate equation

dimensionless length, R1

and the remote boundary condition, which is the elastic singular field here. In
terms of these variables, the condition w

[(D/EB) (Ki E RLa)(n-1-x)/(n-3)]

J [KE I

1 at the crack tip takes the form

(l-KlLeJx dR 1

(27.13)

1.

o
This equation is obtained by integrating eq. (27.1) over the time, replacing dt
by -dx/a and inserting the dimensionless variables. The integral is taken along
the line ahead of the crack tip,

a=

O. The integral has some

as

yet

unknown

value which may depend only on the dimensionless material parameters and on the
bracketed dimensionless combination of variables, which appears in front of the
integral, but not on R1 , nor on a, B, E or D separately. Thus eq. (27.13) above
is an implicit relation for the bracketed combination of variables. If resolved
for the crack growth rate, the result must have the form:
(27.14)
where a is an unknown dimensionless factor. This result for
rate

is

an undisturbed elastic singular field and a


fullfilled

the

remarkably different from eq. (23.39) which predicted


at

some

d~stance

critical-strain

crack
a

growth

K~ assuming

criterion

to

be

Xc ahead of the crack tip. The comparison of eqs.

(27.14) and (23.39) illustrates that seemingly minor changes in the

model

can

sometimes lead to drastic disparities in the results. Although the damage mechanics model appears to be theoretically preferable, experiments tend to support
the stronger dependence, a

K~, as was illustrated by the tests on Nimonic 80A

described in Section 23.4.3. In conclusion, creep

crack

growth

under

scale creep conditions is not as well understood as C*-controlled growth.

small-

27.

358

27.4.2

A Damage Mechanics Approach to Creep Crack Growth

Creep zone grows faster than process zone

If the process zone grows well inside the creep zone (which itself is initially
small compared to the crack length), the evolution of damage can be analyzed as
the following boundary layer problem: far from the crack tip, the stress
must

asymptotically

elastic transient and is given by eq.


elastic

strains

can

is

(23.10).

Inside

this

boundary

layer,

be neglected so that the problem is governed by the same

equations as in the extensive creep


condition

field

approach the relaxing HRR-field which is valid during the

limit

except

that

the

remote

boundary

now time-dependent. Again, there are similarity solutions having

the form of eqs. (27.3) and (27.4). The similarity coordinate is now
(27.15)
Therefore, the crack growth increment and the crack growth rate are
n+1
(n+1)/x
Dt)
n+1-x

~a

K2(1_v 2 )

In general. the dimensionless factor


in

n+1 (a
I
)X/ n+1
X
(n+1) ESt

)
D~

a(n,x,~,K)

(n+1-x)/(n+1)

(27.16)

(27.17)

may be different from that used

the extensive-creep case, but in the following, the two factors are assumed

to be equal. An interpolation formula between the crack

growth

rates

in

the

short-time limit, eq. (27.17), and in the extensive-creep limit, eq. (27.7), is
obtained by replacing C* in eq. (27.7) by
C*

(1 + t 1 /t) C*,

(27.18)

The characteristic time, t 1 , was defined in eq. (23.15) and is meant here to be
calculated for the initial crack length.
A comparison with Section 23.3.1 shows that, in the case of slow
inside

the

crack

growth

creep zone, the continuum damage model leads to basically the same

crack growth rates as the previous model based on the undisturbed HRR-field and
on

a critical-strain criterion. The results differ only by a numerical factor,

if corresponding cases are compared, i.e. X = n and Xc =

o.

21.5

Primary-Creep Effects

359

27.5 Primary-Creep Effects


Primary creep is described here by the

strain-hardening

model

introduced

in

Section 25.1. Damage is taken into account by using eq. (21.1) for wand
(21.19)
Two limiting cases are distinguished in the following analysis.
21.5.1

Small-scale damage in a specimen which creeps in the primary stage

If only the primary-creep term in eq. (21.19) plays a role, the remote boundary
condition

for

the

evolution of the process zone is given by eq. (25.8). Thus

the structure of the mathematical problem is almost identical with that


in

solved

Section 21.2. The final result for the crack growth rate, eq. (21.1), needs

to be modified only by replacing n by m and

C*/B

by

C~/B1 1/(1+p),

and

the

factor a has a different numerical value, which is unknown.


21.5.2

The transient from elasticity over primary to secondary creep

Now creep crack growth is considered for


creep

and

secondary

the

case

that

elasticity,

primary

creep consecutively determine the specimen response, and

the process zorle is small and well contained wi thin the

secondary-creep

zone.

In analogy to previous cases, this case can be dealt with approximately by taking the crack growth nate from eq. (21.1), but with C* replaced by
(21.20).
The same interpolation was suggested in eq. (25.18).

27.6 The Evolution of the Crack Length and the Lifetime


Lifetimes of pre-cracked specimens are estimated now by integrating
growth

laws

the

crack

derived in the preceding sections. In doing so the limitations to

the small-scale damage approximation are ignored. Whether or not this leads
serious

error

to

will be discussed in Section 21.1. There is yet another problem

associated with the integration of the crack growth laws. It is the rule rather
than

the

exception

that

crack growth is predicted to occur along directions

A Damage Mechanics Approach to Creep Crack Growth

27.

360

which are inclined to the original crack plane. This implies two
First

difficulties.

it is not clear whether a crack in a specimen of finite thickness really

branches or kinks macroscopically. Experience in this laboratory

with

various

steels indicates that the crack surface remains macroscopically flat and normal
to the tensile direction. Side-grooving of the specimen supports this tendency.
Second,

if

the

crack

did

actually

branch

macroscopically,

branched cracks would not be available and the crack-tip fields


Mode-II

For

component.

these

failure is controlled by the


equivalent

stress,

i.e. if

C*-values for
would

principal

is

done

for

tensile

stress

rather

than

reasons

that

branches

the

calculat-

(K

< 0.5) and

if

there

are

no

the crack should remain planar, crack growth under

6 = 60 0 is considered. The C*-integral is estimated, then, by


crack

by

in-plane crack extension. If, on the other hand, failure is

controlled primarily by the equivalent stress


experimental

> 0.5, or if the crack is known from experiments

to remain flat, for example, because deep side-grooves are used, the
ion

have

reasons the following procedure is adopted: If

projecting

the

onto the symmetry plane and to consider a planar crack of that

length. Among the examples shown below, this latter procedure is

applied

only

which

crack

to the aluminum alloy.


The following calculation is performed for ductile materials,
growth

occurs

primarily

in

in the C*-controlled limit, but the transient due to

the elastic/plastic response and to primary creep is also taken


Hence,

the

calculation

into

account.

of the crack length as a function of time starts from

the crack growth rate, eq. (27.7), with the correction for the transients,

eq.

(27.20), and the e,xpression for C*, eq. (20.7), inserted. Then the crack growth
law can be integrated by separation of the variables, a and

t.

There

results

the following implicit relationship between a/Wand time:


F(a/W) = [(n+1)/x] (n COS6)x/(n+1) DO~ t G(t/t 1 )

with

F(a/W)

ar

(0 /0

ne

t)X da

J (ag )x/(n+1)( a-a )'-xln+l)


a
o
1

(27.21)

(27.22)

tlt1

G( tlt1 )

(t 1 /t)

f [1 +1 /-r+('2/'r)P/(P+1 )]x/(n+1 )d,.

(27.23)

o
Here, ,

t 2 /t 1 ,

00

and a o are net section stress and crack

length,

27.6

361

Evolution of Crack Length and Lifetime

respectively, at the beginning of the test. The load is assumed to be constant,


and the geometrical function g1 defined in eq.
strain.

The

(20.10)

was

be

for

plane

factor cose accounts for the projection of the local crack growth

direction onto the symmetry plane. The dimensionless functions


generally

taken

F and

G must

evaluated by numerical integration; G accounts for the transient

correction and approaches 1

for

long

times

when

transient

effects

become

negligible.

The group DoXt represents the time in units of the failure time of
o
an unnotched specimen subjected to the tensile stress 0 0 If failure of the
cracked specimen occurs at longer times than 1/(D oX), the material is called
o

notch strengthening; otherwise it is called notch weakening.


Equation (27.21) was evaluated numerically for various specimen geometries
material

parameters.

Results

for

21/~Cr-1Mo

and

steel were already shown in Fig.

25.4. Figure 27.3 shows further results. They were calculated

for

double-edge

cracked tension specimens (DECT) and center-cracked tension specimens (CCT) for
plane-strain. Material parameters for copper and aluminum were taken from Table
27.1. The factor a was computed from the upper-bound estimate, eq. (27.9). The

D(j; t

(forAIIDECi) 10
15

lO~----~--~-T~~~~--~

a)DECT

09

,
,

t)

~ 08

~
tJ Q7

/,~

~-----

Q6

0.5

1.0

./

~~----- AI
1.0

I.

b)CC'"

,,
AI./

.... /

1.5

Fig. 27.3. Normalized crack length vs. normalized time calculated from
eq. (27.21). Note separate scale for DECT-specimens of AI. Dashed line:
without elasticity correction [G(t/t 1 ) = 1J. (After Riedel, 1985b).

27.

362

A Damage Mechanics Approach to Creep Crack Growth

dashed lines represent results which neglect transient effects, i.e. which
the elastic transient. Primary creep effects are neglected in
copper,

are

by setting G = 1, while the solid lines include the correction due to

obtained

the

Fig.

27.3.

For

elasticity correction is apparently small. In fact, the corrected

curve for the CCT-specimen was omitted since it deviates by only 2% in lifetime
from

the

curve

calculated

with

G = 1. For the aluminum alloy, however, the

elasticity correction may be substantial. It enhances


due

to

the

crack

growth

initially

high stresses at the crack tip during the elastic transient. This

initial crack growth increment reduces

the

lifetime.

The

magnitude

of

the

correction

depends primarily on the ratio t f /t 1 Numerical values for this


ratio will be given in the last row of Table 27.2 (which will be shown shortly). The exponent 1-x/(n+1) in eq. (27.22) also plays a role for the importance
of elasticity effects. Values of X near n favor a pronounced

transient,

while

for X < n/2 the transient is often hardly visible in plots like Fig. 27.3.

27.7 Discussion
The general features of the evolution of the crack length,
initial

in

particular

cies in detail may arise primarily from four sources. First, the damage
nics

the

transient, are described qualitatively correctly by theory. Discrepan-

equations

axial stress field at the crack tip. (Recall that the parameter
determined

mecha-

may be inappropriate to model the situation in the highly triK

is

usually

from tension and torsions tests both of which exhibit far less tri-

axiality than a crack tip field). Second, the continuum damage approach becomes
definitely

inappropriate

if

corrosion plays a major role in crack growth as,

for example, in nickel base superalloys. Thirdly, the plane-strain

calculation

based on eq. (20.7) usually under-estimates the C*-integral. Crack growth rates
are underestimated correspondingly. A difference by a factor 10 in crack growth
rates

appears

to

be

possible

according

to the discussion in Section 22.2.

Finally, the small-scale damage approximation, which underlies the crack growth
law,

eq.

(27.7), is expected to become inaccurate for increasing crack growth

increments.
The latter problem caused by the use of the
can

be

resolved

by

small-scale

damage

approximation

comparison with finite element calculations of Hayhurst,

Brown and Morrison (1984). These calculations were performed for the same material

parameters

and

specimen

geometries and are based on the same equations

from which eq. (27.21) and Fig. 27.3 were derived except that a finite

element

27.7

363

Discussion

solution

does not require to make the small-scale damage approximation. Incid-

entally, Hayhurst et al report close agreement of

their

calculated

lifetimes

with experiments. However, conSidering the uncertainties in the constitutive


model and the expected inaccuracy of a plane-strain calculation, the agreement
may

be

fortuitous. But here we are only interested in a comparison of the two

calculations, one of which relies on the small-scale damage approximation while


the other does not. Table 27.2 shows the comparison of the normalized lifetimes
Do~tf' In this normalized notation, the net section stress

0 0 enters only
into
1/0~-1). Since Hayhurst et al do
not report the values which they use for 00 , plausible values were assumed.
Further, since no accurate values for the factor a were available, the upper
and lower-bound estimates, eqs. (27.8) and (27.9), were used. As the table
shows, the bounds for the lifetime are close together and they agree well with
the result of Hayhurst et al as far as copper is concerned. In the case of
aluminum, however, the bounds span a range of up to a factor 12, and, for the
DECT-specimen, do not even encompass the finite element result.

the

elasticity

correction

(recall that t1

The failure of eq. (27.21) to reproduce the finite element result for aluminum
may indicate that the small-scale damage approximation is not accurate enough
in this case. This far-reaching conclusion, however, should not

yet

be

drawn

definitively Since there are other possible sources for error. In particular,
eq. (27.21) cannot deal reasonably with the problem of crack branching or
kinking which is, however, important in aluminum.

Al

D oX t f
0
FE
Eq.(27.21)
for

t f /t 1

Cu

CCT

DECT

CCT

DECT

1.07

0.98

1.17

1.03-6.7

0.82-6.5

8.4-73

5.7-68

0.83-0.97

1.04-1.22

50 MPa

100 MPa

50 MPa

100 MPa

30 MPa

30 MPa

12-75

6-48

4.6-41

2.3-26

41-48

20-24

Table 21.2. Normalized lifetimes from finite element calculations of Hayhurst


et al (1984) and from eq. (27.21). Hyphenated entries represent results for
upper and lower-bound estimate for a, respectively.

28 Creep-Fatigue Crack Growth

In this chapter, the technologically important subject of fatigue at high temperatures

is taken up again. In Chapter 18, fatigue failure due to more or less

homogeneous cavitation of grain boundaries in the whole specimen was discussed,


whereas

now

failure

by

the growth of cracks under cyclic-loading conditions

will be investigated.
Figure 28.1 shows the general trends observed in fatigue crack growth

testing.

As

a function of the stress intensity factor range, 8K 1 , the crack growth rate
per cycle, da/dN, increases starting at a threshold value, 8K th . The power law
observed in the intermediate range is often called Paris' law, and the exponent
typically lies in the range 2 to 4. Further, the crack

growth

rate

increases

with increasing temperature and loading-cycle time and is sensitive to environment, particularly at high temperatures. There
creep-fatigue

crack

growth,

which

exists

vast

literature

on

the reader may find access to through the

following references: Skelton (1978a,b);

Yamaguchi and Kanazawa (1979); Taira,

increasing
aggressiveness of
environment,
hold time,

Fig. 28.1. Fatigue crack growth rate as a function of several variables.

28.1

Micromechanisms

365

Ohtani and Komatsu (1979); Runkle and Pelloux (1979); James (1979);
James

Mills

and

(1980); Michel and Smith (1980); Floreen and Kane (1980); Saxena, Willi-

ams and Shih (1981); Tomkins (1981); Shahinian


(1983a);

Wareing

and

Sadananda

(1982);

Riedel

(1983); Lloyd (1983); Sadananda and Shahinian (1980b, 1981b,

1984); Saxena and Bassani (1984); Floreen and Raj (1985); and Pineau (1985).
This chapter tries to provide a basis for the understanding of
and

the

mechanics

of

creep-fatigue

the

mechanisms

crack growth. In Section 28.1, the most

important micromechanisms (i.e. alternating slip at the crack tip, grain boundary cavitation and corrosion) will be described qualitatively. A combination of
the micromechanisms with the stress and strain-rate fields in

different

types

of materials (viscous, elastic-plastic and elastic-viscoplastic) leads to theoretical predictions of fatigue crack growth rates (Sections 28.2 to 28.5).
analysis

applies

The

both to macroscopic cracks of several millimeters length and

to microcracks in the submillimeter

range.

Microcracks

are

often

initiated

early in fatigue life preferentially at the specimen surface (see, for example,
Heitmann, Vehoff and Neumann, 1984), and their growth

determines

the

fatigue

lifetime (Section 28.5.3). The chapter is concluded by a summary.

28.1 Micromechanisms of Fatigue Crack Growth


28.1.1

The alternating slip model (also called the crack-tip blunting model)

The most common mode pf fatigue crack growth up to moderately high temperatures
and

moderate hold times is a direct consequence of the cyclic blunting and re-

sharpening of the crack tip: the new surface created during the opening part of
the

cycle is not removed reversibly during the subsequent closure of the crack

tip. This irreversibility can be caused, for example,


quickly

by

oxygen

atoms

which

adhere to a freshly exposed metal surface and prevent rewelding of the

crack faces. Therefore, as shown in Fig. 28.2a,

the

crack

advances

in

each

cycle by an increment which, ideally, should be on the order of the cyclic


crack-tip opening displacement. Often the opening and closing event leaves a
microscopically

visible

striation

marking

on the fracture surface, but only

seldom does the total number of striations on a fracture surface correspond

to

the number of loading cycles.


While the global picture shown in Fig. 28.2a was proposed by
(1962),

Neumann (1974) and

Laird

Neumann, Fuhlrott and Vehoff (1979)

and

detailed

Smith
the

Creep-Fatigue Crack Growth

28.

366

crack opening

U
I
------tl

after closing

b)

Fig. 28.2. Crack growth by opening and closing of a crack tip,


(a) by continuum plasticity, (b) by alternating slip on two slip systems.
Currently active slip planes are represented by solid lines.

understanding of fatigue crack growth by showing that crack opening


crystals

occurs
is

single

by alternating slip on two slip systems which intersect along

the crack front (Fig. 28.2b). It is likely


polycrystals

in

quite

analogous.

that

Therefore,

the
the

mechanism

operating

mechanism

is

in

called the

'alternating slip mechanism'.


If the cyclic opening and closing of the crack tip as shown in Fig.
the

whole

truth

of

28.2

were

fatigue crack growth, growth rates could be estimated by

calculating the cyclic crack-tip opening displacement. This

will

actually

be

done in the following sections. However, there are several complicating factors
not all of which are well understood. One of these factors
called

is

the

phenomenon

crack closure (Elber, 1970). As the crack grows it leaves behind a wake

of plastically stretched material, which leads to crack closure well behind the
crack

tip

during

unloading

even when the applied load is still tensile. The

mechanical contact of the crack faces reduces the cyclic stress intensity experienced by the crack tip region. This, and the concommitant reduction in crack
growth rate, was modeled by Budiansky and Hutchinson (1978) and by Newman
(1981).

A further complication arises when a corrosion product is accumulated

between the crack faces, which enhances the closure effect (Section 28.1.3). An
empirical

description

of

analysis of fatigue crack


(Section 28.6).

the
growth

crack

closure

rates,

will

effect, which is useful in the


be

given

in

the

Discussion

28.1

367

Micromechanisms

For completeness, the distinction between Stage-I and


growth

should

Stage-II

fatigue

be mentioned. In its early stages, a small fatigue crack, which

is still contained within a single grain, often does not yet grow by
ing

slip

crack

on

two

slip

systems,

alternat-

but rather grows along a single active slip

plane, which is usually inclined to the tensile axis (Stage I).

The

mechanism

by which a Stage-I crack grows consists in the exposure of fresh surface at the
crack tip by slip in the tensile part of the cycle. Due to oxidation

or

other

irreversible processes, the new surface cannot be rewelded upon reversal of the
slip direction. When the crack grows longer, it usually assumes an
which

orientation

is macroscopically normal to the applied stress and grows by alternating

slip. This is called Stage II.


28.1.2

Fatigue crack growth by grain boundary cavitation

The mechanisms of grain boundary cavitation were described in Part II


book.

At

high

temperatures

of

this

and low loading frequencies or long tensile hold

times, fatigue cracks may propagate by coalescence with grain

boundary

ies.

mechanism act in

The

alternating

slip

mechanism

and

the

parallel, and the one which gives the higher

cavitation

growth

rate

cavit-

predominates.

This

comparison will be made in the Summary (Section 28.7).


A special way in which fatigue crack growth may interact with cavitation damage
was described by Min and Raj (1978). Under the testing conditions used by these
authors, fatigue cracks in stainless
alternating

slip.

the crack path


accompanied

in

by

I~,

steel

usually

grow

transgranularly

by

however, a tensile hold time precedes the fatigue test,

fatigue

changes

to

intergranular,

and

this

change

is

a large increase in crack growth rate. The explanation is that

the hold time causes grain boundary cavitation ahead of

the

main

crack.

The

subsequent fatigue crack growth can occur by fracturing the pre-cavitated grain
boundaries by the plastic hole growth and coalescence mechanism. This
preferred
da/dN

>

only

15~m.

at

mode

is

very high crack growth increments per cycle, typically, at

Otherwise the trans granular mode of crack growth

remains

active

despite the presence of grain boundary damage.


This example illustrates that special types of creep-fatigue
occur

interactions

may

under special conditions. In the tests of Min and Raj (1978) the specim-

ens were subjected to creep straining for some fraction of their lifetimes

and

to fatigue for the rest. More often each loading cycle involves both, creep and
fatigue components, and this is the subject of the following sections.

28.

368

28.1.3

Creep-Fatigue Crack Growth

Corrosive effects in creep-fatigue crack growth

In many of the papers on creep-fatigue crack growth, it


fatigue

crack

growth

has

been

noted

that

rates, just as creep crack growth rates, are often much

larger in atmospheres containing oxygen or sulfur than in vacuum or in an inert


gas. An oxygen partial pressure of less than 100Pa
such an environmental effect. In steels and
factor

(=

Ni-base

alloys,

the

enhancement

for the growth rate is typically 10 in the sensitive temperature range,

as the review of Ericsson (1979) shows, but smaller


been

1 torr) suffices to cause

observed

as

well.

little effect of environment


steels,

and

larger

factors

have

Sadananda and Shahinian (1980b), for example, report


on

fatigue

crack

growth

rates

in

austenitic

whereas Floreen and Kane (1980) obtain environmental enhancement fact-

ors of 50 to 100 in aNi-base superalloy at 650 0 C. This variability is not surprising considering the multitude of possible mechanisms and their interactions
with other mechanisms of fatigue crack growth.
A list of mechanisms which possibly explain the environmental effects on

fati-

gue crack growth has been given by Floreen and Raj (1985). Here only the most
important of these mechanisms will be discussed. In the minority of cases, the
presence of oxygen can be beneficial for the high-temperature fatigue lifetime.
One of these strengthening mechanisms is the formation of a

corrosion

product

between the crack faces. This enhances the crack-closure effect, which is known
to retard fatigue crack growth. Suresh, Parks

and

Ritchie

(1982)

show

that

already at room temperature in ferritic steels, fretting corrosion of the crack


faces can lead to oxide layers of some hundred
the

crack-closure

nanometers

applied stress intensity amplitudes. This may be one


threshold

observed

thickness.

Through

effect, these layers can stop fatigue crack growth at small


commonly

for

fatigue

crack

of

the

causes

for

the

growth at low AK r . Ericsson

(1979) explains oxidation strengthening, which is sometimes observed in

super-

alloys at high temperatures, by oxide-induced crack closure, too.


More often, however, an oxygen-containing atmosphere is detrimental for the
high-temperature fatigue lifetime. This may have several causes. First, surface
cracks in an initially smooth specimen are nucleated easier in the presence
oxygen;

of

but this is not the subject of the present chapter. Second, in the ab-

sence of oxygen, fatigue crack growth may be impeded by full or partial rewelding

of

the crack faces during unloading (Neumann, Fuhlrott and Vehoff, 1979),

whereas a layer of oxygen, which quickly adheres to the freshly


surface,

exposed

metal

or a layer of oxide prevent rewelding. By such a mechanism, oxidation

28.1

Micromechanisms

369

effects enhance transgranular crack growth by alternating slip. Third, since no


stable protective oxide layer can form at the continually and heavily deforming
crack tip, oxygen can penetrate into the material. Usually, this
grain

boundaries.

In

the

material,

oxygen

occurs

alloy, for example with carbides. As a consequence, the carbides


from

the

matrix,

and

to

decohere

presence

of

oxygen

intergranular crack growth, and Floreen and Raj (1985) consider such

subsurface corrosion damage to be primarily responsible for


effect

may

pressurized carbon-oxide bubbles and brittle or weakly

bonded metal oxides can be formed. By this mechanism, the


leads

along

reacts with constituents of the

of

the

environmental

oxygen in superalloys. Besides internal oxidation, also sulfidation

(Floreen and Kane, 1982) and, to a lesser extent,

carburization

(Floreen

and

White, 1981) playa role in the appropriate environments.


A global model for fatigue crack growth by subsurface damage formation in

cor-

rosive environments was proposed by Saxena (1983). The corrosive atoms entering
the material through the crack tip spread over a diffusion length which increases with time as (Dt)1/2, where D is their diffusivity in the host material. If
these atoms embrittle the material, for example by oxide formation on grain
boundaries, the material fractures over some, possibly stress-dependent, fraction of the diffusion zone during the tension-going part of the
Then

loading

cycle.

diffusion starts again around the new crack-tip position. This model must

lead to a fatigue crack growth rate which increases with hold time, or decreawith loading frequency, according to da/dN
1/v1 / 2 with an activation

ses

energy half that of the diffusivity D (because of the square-root dependence of


the

diffusion

length

on D) and with a not well defined dependence on loading

amplitude. In fact, Saxena (1983) and Saxena and Bassani (1984)


inverse

square-root

dependence

on

show

that

an

loading frequency is observed at moderate

temperatures, i.e. at 427C in a CrMoV steel, and at 650 0 C in

the

nickel-base

alloys astroloy and Inconel 718.


Since, in accord with the above model, the activation energy of

oxidation-con-

trolled fatigue crack growth rate is often found to be small, corrosion tends
to predominate at int~rmediate temperatures, while at high temperature, the
effects

of creep deformation and grain boundary cavitation determine the time-

dependence of the crack growth rate. From tests by Pelloux and Huang (1980)
the

nickel-base

alloy

astroloy in air and in vacuum, Saxena (1983) concludes

that at 650 0 C oxidation is still the relevant process,


effects

on

while

at

760C

creep

predominate. For a CrMoV rotor steel the transition seems to lie some-

where between 427C and 538c (Saxena and Bassani, 1984).

370

28.

Creep-Fatigue Crack Growth

28.2 Fatigue Cracks in Viscous Materials


The stress analysis of
particularly

cracked

bodies

under

cyclic

loading

conditions

is

simple for viscous materials. The stress field is always a funct-

ion of the current load only but not of the prior history. The crack-tip stress
fields

are

uniquely characterized by the current value of the C*-integral, so

that C* is the appropriate load parameter. It


that

the

range

should

be

mentioned,

however,

of validity of a purely viscous description is rather limited

under cyclic loading conditions, as will be shown later.


28.2.1

Growth rates by the alternating slip mechanism

It was pointed out in Section 28.1.1 that the crack growth increment per cycle,
da/dN, must be of the order of the cyclic crack-tip opening displacement

~O.

In

purely viscous materials, the displacement rate at any point is proportional to


C*n/(n+1)

Ceq.

(22.2)].

Integrating

over

the tensile part of the cycle and

using the definition of the crack-tip opening displacement given in

eq. (22.3)

leads to the crack growth rate in purely nonlinear viscous material


da/dN

(28.1 )

The integral extends over the tensile part of the cycle and a numerical
contained

factor

in eq. (22.3) is approximated by unity. If the tensile strain is not

completely reversed in each loading cycle, the crack blunts

progressively.

In

real materials, this will favor intergranular crack growth by cavitation.


28.2.2

Growth by cavitation in viscous materials

In Chapter 21, creep crack growth rates were calculated based on cavity
laws.

Since

in

that

growth

analysis it was not necessary to assume that C* is con-

stant, it can be applied to cyclic loading as well. Integration

of

the

crack

growth rate given in eq. (21.6) over a cycle gives da/dN. For strain-controlled
failure at the crack tip there results
da/dN

(28.2)

Xc was taken, and the numerical factor 8n combines constants from eq. (21.6), which are of no interest here. If cavitation is revers-

Here, the limit a-a o

ible during compression, the time-integral in eq. (28.2) extends over the whole

28.2

371

Fatigue Cracks in Viscous Materials

loading

cycle

with

the

integrand

being

negative

cavitation damage is not recoverable, the integral

during

extends

compression.
over

the

If

tensile

part of the cycle only. In both cases, the principal dependencies of eq. (28.2)
are given by da/dN ~ c*n/(n+1)/V~' where v~ is the loading frequency.

28.3 Fatigue Cracks in Elastic-Plastic Materials


In this section the material response is described as time-independent elasticplastic.

This

behavior predominates at moderate temperatures and high loading

frequencies. Thus it represents the limiting case opposite

to

purely

viscous

behavior, which predominates at high -temperatures and slow load variations. The
elastic-plastic response of the material to cyclic
usually,

by

total

loading

is

described,

as

strain rate which is a sum of elastic and incrementally

plastic strain rates, but the plastic strain has a special property. The stress
and

strain

variations, 00 and OE, measured from the point of load reversal in

the hysteresis loops must be related by a unique law, independent of the amplitude of the stress, and this relationship is assumed to have the power-law form
B

(00) 1 IN,

(28.3)

where N is the hardening exponent and the factor Bo is related


yield

stress,

to

the

cyclic

0Cy; if, for example, the cyclic yield stress is defined as the

O.2%-offset stress, then it follows from eq. (28.3) that Bo = O.002/(OCy)1/N.


Real materials often exhibit such a unique relationship between 00 and OE to
within a good approximation.
28.3.1

Elastic-plastic deformation fields

Deformation fields in elastic-plastic materials under cyclic loading conditions


were

analyzed

for

Mode-III loading by Hult and McClintock (1957), McClintock

and Irwin (1965) and by Rice (1967b). An important feature of the solutions
that

is

in elastic-perfectly plastic materials the Mode-III deformation fields in

the plastic zone are proportional fields, i.e., material elements are
plastically

power-law material like that described by


neglected

strained

along principal directions which remain fixed during loading. In a


or,

in

other

eq.

(28.3)

(with

elastic

strains

words, in the fully plastic limit), the fields under

Mode-I loading are also proportional fields for the same reason that

power-law

materials develop proportional fields under monotonic loading. Thus the propor-

28.

372

tionality of the fields is guaranteed in

Creep-Fatigue Crack Growth

important

limiting

cases,

and

its

approximate validity between these limits is assumed. Crack closure, i.e. mechanical contact of the crack faces during unloading and compression, renders the
fields

near

the

contact

zone non-proportional. Thus the occurrence of crack

closure over a considerable distance behind the crack tip and over

substantial

fractions of the loading cycle limits the validity of the following arguments.
28.3.2

The cyclic J-integral, Z

If the deformation fields

are

proportional

fields,

as

pOinted

out

above,

incremental plasticity and nonlinear elasticity become equivalent. This has the
important consequence that the cyclic deformation fields at crack tips
characterized

by

path-independent

can

earlier been denoted by 8J (Dowling and Begley, 1976). It is analogous


J-integral

with

be

integral, Z (Wuthrich, 1982), which had


to

the

stress, strain and displacement replaced by their variations,

60, 6E, 6u, measured from the point of load reversal. The crack-tip fields have

the HRR-type form of eq. (24.5) with Z substituted for J and 60 i , for 0, ,.
J
lJ
Since the Z-integral characterizes the crack-tip fields in elastic-plastic materials under a wide range of conditions, at least approximately, it is expected
to be a useful load parameter to characterize fatigue crack growth. The Z-integral

can

be

calculated using the analogy to the J-integral. According to the

Plastic Fracture Handbook by Kumar et al (1981), Z can be approximated

by

the

sum of an elastic (or small-scale yielding) and a plastic contribution as


Z

with

(28.4)

The plastic part, Z l' is given by formulas analogous to those given for C*
Section

20.2.1

and

For small loading amplitudes, Zpl can be neglected and Z and 8K I


valent.
28.3.3

in

in Appendix C with B replaced by Bo' 8 by 8 and n by liN.


become

equi-

Z-controlled crack grOwth rates by alternating Slip

If fatigue crack growth occurs by the

alternating

slip

mechanism

the

crack

growth rate should be of the order of the cyclic crack opening displacement. In
elastic-plastic material, displacements are

given

by

an

equation

like

eq.

28.3

Fatigue Cracks in Elastic-Plastic Materials

373

(22.2) with u i ' B, C* and n replaced by u i ' Bo ' Z and l/N. The crack-tip
opening displacement is analogously obtained from eq. (22.3), and this gives
(28.5)

da/dN
where
the

0
is the cyclic yield stress measured from the point of load reversal,
cy
numerical factor in the square brackets in eq. (22.3) was approximated by

1, and in the second form of eq. (28.5), N was taken as 0.1.


Experimentally, the dependence of da/dN on Z is usually found

to

be

stronger

than the linear dependence suggested by eq. (28.5). A dependen.ce


da/dN ~ Z(1.2 to 1.5) is typical. The deviation may arise from any of the

like
com-

plicating factors mentioned in Section 28.1.2. Of course, the crack growth rate
in rate-independent material is independent of the loading frequency.
The cavitation micromechanism does generally not predominate under

time-indep-

endent elastic-plastic conditions, and will therefore not be considered here.

28.4 Fatigue Cracks in Elastic/Nonlinear Viscous Materials


An elastic/nonlinear viscous material is characterized by the material law given

in

eq.

(?3.1), which reduces to

o/E

Bon in uniaxial tension. Such a

material behaves purely viscously if all load variations occur slowly, with the
characteristic

time

tl

defined in eq. (23.15) being the relevant time scale.

The elastic response of elastic/nonlinear viscous material becomes important if


an otherwise slow loading cycle contains a rapid load variation or if the whole
cycle time is smaller, or not much greater, than t 1 .
28.4.1

Stress fields in elastic/nonlinear viscous material after a load step

As a first example in which the elastic response of the material plays a


a sudden load step,

~P,

is considered. Prior to the load step, the specimen has

experienced a stressing and straining history and therefore


stresses.

This

distinguishes

contains

internal

a load step from the step load considered prev-

iously in Chapter 23. The stress


-

role,

fields

immediately

before

and

immediately

after the load step, 0ij and 0ij' are related by


(28.6)

28.

374

Creep-Fatigue Crack Growth

where 60 ij is the elastic field associated with 6P (Riedel,

1983a).

Near

the

crack

tip, the elastic field, 60 .. , dominates asymptotically since the elastic


IJ
1/!r-singularity is stronger than the 1/r 1 /(n+1)-HRR-type singularity of O~j.

Hence,

the

initial condition for the crack-tip field after a load step is the

same as that after a step load, and the short-time

solutions

after

the

load

be taken from Section 23.1 with 6K r substituted for Kr . For example,


the duration of the elastic transient after a load step is

step

can

(28.7)

(n+1)EC*
where C* is taken as the steady-state value after the load step.
Figure 28.3 schematically shows the
loading,

which

crack-tip

stress

field

for

square-wave

consists of a sequence of load steps. The loading frequency is

1/t 1 . The analytic short-, and long-time behavior of


field indicated in the figure is taken from Chapter 23. Obviously,

assumed to be small, v t
the

stress

the stressing history is not characterizable by a single load parameter;


6K r

and

C*

both,

determine the crack-tip fields during parts of the loading cycle.

Further, one observes that each load step is followed by a stress peak near the
crack tip.
28.4.2

Gradual load variations in elastic/nonlinear viscous material

As a generalization of step loading, a gradual load variation is now considered


which obeys the pOMer law

I~(t)~ LlK,'I1-v')

15

t to

(n+l)Et

~(t)+~

.!::
11)05

load

r-

I\,

If

/'
time--

Fig. 28.3. Stress field at crack tip for slow, square-wave loading.
After Riedel (1983a). Adapted, with permission, from STP 803, copyright
American Society for Testing and Materials

28.4

375

Fatigue Cracks in Elastic/Nonlinear Viscous Materials

ex

til,

(28.8)

with the arbitrary exponent Il. The specimen is now assumed to have no
stresses

due

to

internal

a possible prior loading history. At short times, the fields

developing in response to the increasing load can be described using similarity


solutions

just

as

in

the case of the step load. There is a creep zone which

grows according to eq. (23.14), but now with a time-dependent Kr

ex

til and

with

a shape which depends slightly on the exponent Il (Riedel, 1983a). The crack-tip
field

is

given

K~
replaced
by
For long times, nonlinear viscous behavior is approached. The
for

short

times

by

eq.

(23.10)

with

(1+2Iln)[K r (t)]2.
crack-tip field is then the HRR-field, eq. (20.16), with C* increasing in time
according to C* ex t ll (n+1). The characteristic time of the elastic transient
under continually increasing load is given by:
2

1+21ln Kr (l-v )
n+1
EC*
Figure 28.4 shows the crack-tip fields

for

(28.9)
a

lin~ar

load

increase

(Il

1)

followed by a hold time. The stress is normalized by the value which it assumes
after a long hold time. The behavior depends on whether the load-rise time, t r ,
is smaller or larger than t 1 . For rapid loading, tr < t 1 , the stress at the end
of the rise time is much greater than its long-time value. For slow loading, no
such

stress

peak occurs and the crack-tip field follows the applied load, and

can be described by C*(t), except for the small hump at short times.

f 2.0

::J

1.5

11)

1.0

....~
~

c::

aJ slow loading

0.5
0

3 0

normalized time t/t1

Fig. 28.4. Crack-tip stress in response to a linear load increase followed by


a hold time. The stress is normalized by its long-time value. Dashed line:
short-time solution; n s 4. After Riedel (1983a). Adapted, with permission,
from STP 803, copyright American Society for Testing and Materials.

376

28.

28.4.3

Creep-Fatigue Crack Growth

Stress fields for rapid cyclic loading

If the whole loading cycle is fast, v~ > llt l ,


complete strain reversal must be distinguished.

wave-forms with and without


For completely strain-reversed

cycles, the overall response of the cracked specimen remains elastic. This case
of

rapid, fully reversed cyclic loading can therefore be treated in the spirit

of small-scale creep by prescribing a time-dependent elastic singular field

at

a large distance from the crack tip (Riedel, 1983a). Dimensional considerations
analogous to those used in connection with the short-time solution after

step-

loading show that the near-tip HRR-field must then have the form

(28.10)
The dimensionless function of time,
of

dimensional

consistency,

S(v~t),

cannot be determined

by

arguments

nor by assuming approximate path-independence of

the J-integral, as was possible in Section 23.1.2. Finite element calculations,


carried out by Riedel (1983a), gave the results shown in Fig. 28.5.

A compari-

son of the first with the following cycles indicates that a steady cyclic state
is

reached

very quickly. In other words, the internal stresses created during

cyclic loading do not affect the later evolution of the deformation fields
a long time.

This means that, for example,

for

square-wave loading can be treated

11t~

normalized time
Fig. 28.5. Crack-tip stress for rapid cycling. Solid lines: first loading
cycle; dashed lines: all following cycles; light lines: load. Finite element
calculations for n=5, v=0.3. After Riedel (1983a). Adapted, with permission,
from STP 803, copyright American Society for Testing and Materials.

28.4

Fatigue Cracks in Elastic/Nonlinear Viscous Materials

approximately as a sequence of independent load steps,


shape

377

or the triangular wave-

shown in Fig. 28.5 can be regarded as a sequence of gradual load variat-

ions with a linear dependence of load on time.


The case of rapid cycling without full strain reversal has not yet been
satisfactorily.

The

probably identical to that obtained


superimposed

is

for

strain-reversed

rapid

(Riedel,

slip

1983a).

If

fatigue

crack

growth

amount

of

But

occurs

by

the

mechanism, the rapid cyclic component will play the dominant

role, but the creep component may play at least a secondary


the

cycling.

a slowly varying creep component whose behavior cannot be de-

termined easily
alternating

solved

rapidly varying component of the near-tip stress field is

crack

closure

and

thus

role

by

reducing

accelerating crack growth. If grain

boundary cavitation is the relevant micromechanism, the creep component will be


of primary importance.
28.4.4

Crack growth rates by the alternating slip mechanism

The crack-tip opening displacement, and hence the crack growth rate, in elastic/power-law viscous material is given by eq. (28.1), but with the C*-integral
replaced by the time-dependent amplitude of the crack-tip field, C(t).
slow-loading

limit

is C(t)

In

the

C*. For rapid cyclic loading, on the other hand,


as C(t) = Sn+1
(1-v2 ) v~/E. The crack

8Ki

C(t) is obtained from eq. (28.10)


growth rate results to be
da
dN

(28.11)

In eq. (28.11), the integral extends over the part


which

S(v~t)

0.67

the

loading

cycle

in

is positive. The value of the integral is estimated from Fig. 28.5

to be 0.47 for the sinusoidal wave-shape, 0.44 for


and

of

for

the
will

triangular

wave-shape

trapezoidal wave-shape. Thus in this case, the trapezoidal

wave-form gives the highest crack growth rate.


da/dN ~ 1/V~/n,

the

be

shown

together with other results.

graphically

The
in

dependence
the

Summary

on

frequency,

(Section 28.7)

28.

378

28.4.5

Creep-Fatigue Crack Growth

Fatigue crack growth by cavitatIon ahead of the crack

Intergranular fatigue crack growth by grain boundary cavitation


main

ahead

of

the

crack starts to dominate at high temperatures and long hold times. Incid-

entally, this transition between different micromechanisms has

nothing

to

do

with the transition from KI - to C*-control. The latter transition is related to


the deformation behavior rather than to the cracking behavior of the material.
It appears to be impossible to predict the nucleation and

growth

behavior

of

cavities in response to the complicated crack-tip fields in elastlc-viscoplastic materials. Hence, we model cavitation damage at the crack by
Rabotnov

equations,

the

Kachanov-

and assume that damage that has developed in the interior

of the material during tension is removed reversibly during compression.

Thus,

in a balanced cycle, the crack starts growing in effectively virgin material at


the beginning of each tensile cycle. Then the crack growth increment at the end
of

cycle

can be taken from eq. (27.6) for the purely viscous limit or from

eq. (27.16) during the elastic transient. Equation (27.18) provides


polation

formula

an

inter-

between the limiting cases. Thus one obtains the growth rate

by cavitation under square-wave loading as

da/dN

tt
(aC*/B) [0 f (l+t It)x/(n+1)dt](n+1)/x

(28.12)

Here, tt is the tensile hold time, t1 is the characteristic time defined in eq.
(28.7), a is the dimensionless factor introduced in eq. (27.6), and 0 and X are
material parameters which appear in
damage,

eq.

the

kinetic

law

for

the

evolution

of

(27.1). For slow cycling, the short-time response of the material

can be neglected,
da/dN 1/vin+1)/~.

i.e.
If

dependence
and
the
frequency
is
t1 It
0,
the short-time response dominates, eq. (28.12) reduces
2

to

da/dN

i.e., da/dN
will

be

a (6K I )

2
(l-v )/E

(n+1) Btt

Ott
)(n+1 )/X
l-x / (n+1)

(6KI)2/vin+1-x)/x. The whole frequency dependence of eq.

compared

(Section 28.7).

with

(28.13)

(28.12)

that of the alternating slip mechanism in the Summary

28.5

379

Combined Effects

28.5 The Combined Effects of Elastic, Plastic and Creep Deformation


on Fatigue Crack Growth Rates
An elastic-viscoplastic solid is modeled by a material law which,

in

uniaxial

tension, has the form:


e:

(28.14)

alE + (B IN) (da) 1 IN-l a + Ban.

The notation is the same as in Chapter 24, except that for cyclic

loading

the

plastic response is characterized by the stress difference, da, with respect to


the point of load reversal in the hysteresis loop. Special cases of this material

law

were considered in the preceding sections. Now the limiting cases are

put together to form a more complete picture of creep-fatigue crack growth.


28.5.1

An approximate general expression for the crack growth rate by alternating Slip

The crack growth rate by alternating slip has a time-independent elastic-plastic

component

given by eq. (28.5) and a contribution by creep, eq. (28.1). For

short to moderate hold times, also the transient


cases

between

these

two

limiting

plays a role. The transient is treated in the following approximate way.

The analysis is confined to the case n


without

much

liN, which simplifies the

calculation

loss in generality. Only square-wave loading is considered init-

ially, but the results can be written in a form that does not
speCialization

depend

on

strongly. The analysis of the transient starts from small-scale

yielding conditions. Then, for short times after a load step, the stress
at

the

crack

this

tip

is

given by eq. (24.10) with

Kr

replaced by

~Kr.

field

From the

stress, the crack-tip opening displacement, which defines the crack growth rate
for growth by alternating slip, can be calculated with the result:
daldN

BN Z 1 [1 + (n+l) Btt/B ]l/n.


o e
0

(28.15)

The elastic part of tbe Z-integral, Zel' was defined in eq. (28.4)
the tensile hold time.

and

eefore generalizing eq. (28.15) to large-scale yielding and creep,

we

in

passing
and

is

mention

that similar formulas have been proposed and compared with experi-

mental data by Saxena (1980b), Saxena, Williams and Shih


Shih

tt

Saxena

(1981),

Swaminathan,

(1982), and Saxena and Bassani (1984). They add an empirical

380

28.

Creep-Fatigue Crack Growth

exponent to Zel' allow for effects of the load-rise time and of the decay time,
include

threshold

stress

intensity

factor

and

BN Z
o pI

to

use a slightly different

functional form than eq. (28.15).


We now generalize eq. (28.15) by
yielding,

and

adding

account

for

large-scale

B1/nC*t~n+1)/n [from eq. (28.1)] to include steady-state creep.

This gives the general result for the creep-fatigue

crack

growth

rate

under

square-wave loading:

(28.16 )

da/dN

Of course, the additive superposition of the growth rates from various limiting
cases

can

only

be an approximate interpolation between these limits. In con-

structing eq. (28.16), a transient was included only in the elastic part of
since

the

stresS

field in a material with n

Z,

1/N exhibits no transient when

loaded directly into the fully plastic state.


For practical applications, it is convenient to combine
Zpl

the

terms

containing

and C*. This will be done in the following two subsections for two special

cases.

28.5.2

Creep-fatigue crack growth rates in fracture mechanics specimens

Since in fracture mechanics specimens, the load-line

deflection

rate,

6,

is

generally measured, C* can be expressed through eq. (20.14) as C* ~ 6(n+1)/n in


such cases. Then the last term of eq. (28.16), which will be denoted by
for later convenience, takes the form

g3

b.(n+1)/n /a 1/n
cr

BN Z
o cr

(28.17)

where the dimensionless function of specimen geometry, g3' was defined

in

eq.

(20.15) and b. cr is the load-line deflection due to creep.


Now we recall the theoretical result reported in Section 23.1.6 that the

load-

line displacement rate of CT-specimens and of similar test specimens approximately reflects the transient at the crack tip. Therefore, if b. cr is interpreted
as the time-dependent part of the deflection irrespective of whether it arises
from steady-state creep or from the transient, the transient is

already

into account and the bracketed term multiplying Zel in eq. (28.16) can

taken
be

de-

28.5

381

Combined Effects

leted. Hence, the crack growth rate becomes


da/dN

(28.18)

BNo (Z el + Zne)'

The nonelastic contribution is defined as Zne = Zpl + Zcr' This is very similar
to an approach proposed by Okazaki and Koizumi (1983), who suggested to correlate creep-fatigue crack growth rates by a sum of an elastic-plastic Z-integral
and

a contribution by creep. The derivation of eq. (28.18) shows why and under

which circumstances such a superposition is possible with no regard

to

trans-

ient effects.
Next, a more convenient expression for the determination of Zne is derived.

To

achieve this, Zpl is expressed by ~l using the nonlinear analogue of eq.


1+N + 81+N (8
+ 8 )1+N is made,
(28.17). If, further, the approximation 8pI
cr = pI
cr
the deflections by plastiCity and creep can be combined as 8ne = ~l + 8cr ' and
Zne becomes
(28.19)
The second form is generally more convenient to apply; g2 and g3

were

defined

in eq. (20.15), and 80net is the cyclic range of net section stress. Now Zne
and da/dN are determined by quantities which can be measured unambiguously
during the credp-fatigue test even for wave-shapes other than square-wave loading. except for 8pl and N. (Only for square-wave loading is it possible to
plastiC displacement from creep displacement). However. the dependence

discern

of Zne on these quantities is weak, so that small errors have no great effect.
For practical applications of eqs. (28.18) and

(28.19),

it

is

important

to

consider a few further pOints described in the Discussion (Section 28.6).


28.5.3

Fatigue lifetimes of initially smooth specimens by microcrack growth

As another

applicatio~

of the preceding analysis we consider fatigue failure of

an initially smooth specimen by the growth of surface microcracks. This failure


mode predominates over homogeneous cavitation up to moderately
ures

and

high

temperat-

moderate hold times. To calculate the number of cycles to failure we

consider the growth of a semi-circular surface crack. which corresponds to


crack-front

shape

usually

the

observed. The wave-shape of the applied stress and

the resulting hysteresis loop are shown in Fig. 28.6. It should be noted that

382

28.

Creep-Fatigue Crack Growth

~ ~~

/1

~ ~

~a

1/

I~

--

Fig. 28.6. Hysteresis loop for strain-reversed square-wave loading (schematic).

here the stress is prescribed, whereas experimentally it is easier


the

strain.

In

to

control

terms of strain, the wave-form shown in Fig. 28.6 corresponds

approximately to a 'slow-fast' test, which is characterized by a slow


going strain rate and a fast compression-going strain rate.

tension-

The crack growth rate for arbitrary specimen geometry was given in eq. (28.16).
It

can

be

specialized to a semi-circular surface crack having the depth a by

inserting the following expressions for the load parameters:


(28.20)
(28.21)

C~

= 2.4

Ecr (1+3/n)-1/2

= 1.9

(28.22)

Ecr

where dO is the applied stress range, .dEPl is the instantaneous plastic


range,

is

the

peak

strain

stress and Ecr the creep strain rate (Fig. 28.6). The

second forms for Zpl and C* are valid for n = 1/N = 5. The expressions above
are taken from the solutions of He and Hutchinson (1981) for the penny-shaped
crack, but modified by a factor 1.25 in order to account for the

fact

that

surface crack is considered (Heitmann, Vehoff and Neumann, 1984). Such a factor
occurs between a crack in an infinite body and a surface crack in
elasticity.

Of

course,

this

can

only

be

plane-strain

an approximation for the present

three-dimensional configuration in power-law material.

28.5

383

Combined Effects

As in the preceding section, it is convenient to combine ZpI and C* to

form

nonelastic term, which leads to


(28.23)

da/dN

(28.24)

with

where the nonelastic strain range is ~Ene = ~EPI + ~Ecr = ~Etot - ~Eel In
contrast to the preceding subsection, the transient term multiplying Zel must
now be retained, since strains are now measured far away

from

the

microcrack

and cannot be expected to reflect the transient behavior of the stress field at
the crack tip. Using the material law, the transient term in eq. (28.23) can be
written in a more convenient form by substituting one of the forms
500

~E

(28.25)

cr

where acy is the cyclic yield stress at 0.2% plastic strain. To calculate the
fatigue lifetime, the growth law, eq. (28.23), is integrated by separation of
the variables, crack length, a,

and

number

of

cycles,

N.

The

integration

from N = 0 to the number of cycles to failure, Nf , and from some small


initial crack length, ai' to a critical length, a f . Cracks of length a i are

extends

assumed

to

be nucleated within the first few cycles, which is consistent with

observations en several steels (Heitmann, Vehoff and Neumann, 1984; Ebi, Riedel
and

Neumann,

1986).

The values of a i and a f are of little importance for the


resultIng fatigue lifetime, which is found to be:
9.n(a f /a i )
N

8 0 DCF

(28.26)

acy

Here the quantity


DCF = 1.45

)2

~E

~a

n 1/n

~ [1 + (n+1) ~ (-;;-) J
E

~~PI

+ 1.9 M ~Ene

(28.27)

was introduced, which will be called the damage parameter for creep-fatigue. It
is determined solely by quantities which can be taken from the hysteresis loop,
and it is a measure of the damaging effect of a given loop, provided that failure

occurs

by microcrack growth. For time-independent crack growth, i.e. when

~Ene = ~EpI' DCF degenerates into the damage parameter ZD = Z/a


introduced by
Heitmann, Vehoff and Neumann (1984) for fatigue testing at room temperature.

28.

384

Creep-Fatigue Crack Growth

28.6 Discussion
The following list reviews the assumptions underlying the above results:
(1) The material was treated as an elastic-viscoplastic continuum. This may
a

limitation

be

to the theory when applied to microcracks. Microcracks some-

times obviously interact with grain

boundaries

or

other

microstructural

features (Miller, 1985).


(2) Specific assumptions were made on the crack-growth
slip

and

mechanism.

Alternating

grain boundary cavitation were considered, but corrosive effects

were ignored. The alternating slip mechanism invariably leads to

linear

dependence of da/dN on Z and C* (or on ~Ki for small-scale yielding, or


da/dN a for small cracks). Observed dependences are generally stronger,
although

the

predicted behavior is sometimes approached, in particular in

small-crack experiments (Skelton, 1978b; Yamaguchi et


1981;

Ermi

and

aI,

1978;

Tomkins,

Moteff, 1983; Wareing, 1983). To rationalize the stronger

than linear dependence, it seems reasonable to assume that the alternating


slip mechanism rarely prevails in its pure form, and it is more realistic
to assume that da/dN is equal to some power, q, of
displacement.
(i.e. da/dN

the

crack-tip

opening

Then da/dN is a homogeneous function of degree q of Z and C*


6Kiq or da/dN

zq or da/dN

C*q in the appropriate limiting

cases). Equation (28.18), for example, becomes


da/dN

(Z I
e

Z )q.
ne

(28.28)

The parameter DCF retains its meaning as a damage parameter, which


fies

the

quanti-

damaging effect of loading cycles of various amplitudes and hold

times, provideti that damage consists in microcrack

growth.

In

fatigue

test with constant amplitude, the lifetime would then be predicted as


(28.29)
(3) Crack closure is one of the major effects neglected so far. For

rate-inde-

pendent material, Heitmann, Vehoff and Neumann (1984) find that the closure
effect can approximately be incorporated if 6K I , which enters into Zel'
modified according to

is

(28.30)

28.6

Discussion

385

where R = Pmax/Pmin; ~o/E in the elastic part of DCF should be modified


accordingly. The terms arising from plasticity and creep are assumed to be
unaffected by crack closure.
(4) Pr.1mary-creep effects were ignored. As
primary

creep

will

lead

to

Kubo

(1981 a, b)

has

demonstrated,

enhanced crack growth rates compared to eq.

(28.16). However, eq. (28.16) was subsequently re-written in forms containing

measured

displacements or strains rather than the material parameters

of secondary creep. It is expected that by measuring displacement or strain


one approximately takes into account primary-creep effects on crack growth.
(5) Since the analysis was carried out for square-wave loading originally,
(28.16)

is

not directly applicable to other waveforms. However, by intro-

ducing displacements or strains instead of hold


meters,

the

eq.

time

and

material

para-

equations following eq. (28.16) should be approximately valid

for other waveforms as well.


(6) The assumption n = 1/N is justified approximately by the behavior of real
materials. If it were seriously violated, the transient term multiplying
Zel in eq. (28.16) would have to be modified.

28.7 Summary
The deformation fields in cracked, elastic-viscoplastic bodies have been analyzed and the ranges in which different load parameters dominate have been identified. For rapid strain-reversed cycling, i.e. for
v~

> 1/t1

(n+1)C*/Z,

the

Z-integral

is

instantaneous plasticity is confined to a small


into

its

small-scale

yielding value, Z

high

loading

frequency

the appropriate load parameter. If


plastiC

zone,

Z degenerates

~K2I and the stress intenSity factor

range, ~KI' can be used. For very slow, continuous cycling, v~ 1/t 1 , the
C*-integral is the appropriate load parameter. In the transition range between
slow and fast cyclic loading, or if an otherwise slow

loading

cycle

contains

rapid load variations, the crack-tip fields are not characterizable by a single
parameter, but both Z (or ~KI) and C* are needed. In Section 28.5, mixed parameters were derived, which interpolate between the limiting cases. In fracture
mechanics speCimens, for example, the parameter Zel
(28.18) and (28.19) should be applicable.

Zne as

defined

in

eqs.

Creep-Fatigue Crack Growth

28.

386

Assuming that crack growth occurs by alternating slip leads to linear dependencies

of

the

crack

dependence is

only

growth
rarely

rate on the load parameters. However, this linear


observed

suggesting

that

the

alternating

slip

mechanism does not usually operate in an undisturbed form.


Figure 28.7 compares the frequency, or
growth

rate

hold-time,

dependences

of

the

predicted by various crack growth mechanisms. It is expected that

the mechanism giving the highest rate dominates. At high frequencies, or


hold
~KI

times,

alternating

as

[1+(n+1)Bt IB ]1/n
according to
to'
limit, crack extension by coalescence with
should

be

reiterated

eq.

increases

occur

and Saxena
corrosive

at

with

hold

(28.15). In the slow-cycling

cavities

that the two transitions from

usually
~KI

dominates.

It

to C* and from alter-

nating slip to cavitation have completely different origins and will


not

small

slip is the fastest mechanism. In the range in which

dominates, the crack growth rate by alternating slip

time

crack

therefore

the same loading frequency. If the conclusions of Saxena (1983)

and

Bassani

effects

may

(1984)
control

mentioned
crack

in

Section

28.1.3

are

correct,

growth at intermediate frequencies and

temperatures giving a slope 1/2 in Fig. 28.7.


If the fatigue lifetime of an initially smooth specimen is controlled by microcrack

growth

(i.e.

if crack nucleation occurs early in the lifetime), DCF as

defined in eq. (28.27) is an appropriate damage parameter to


damaging

effect

of

given

loading

cycle

with

given

characterize
stress

and strain

amplitudes. Fatigue lifetimes should be determined by DCF

{og(1l"Vt )
Fig. 28.7. Fatigue crack growth rates vs. inverse loading frequency

for various micromechanisms. Schematic for n = 6, X = 5.

the

Appendices

Appendix A: Material Parameters


Table A.l shows a selection of material parameters for a few pure metals

taken

from the sources indicated. The shear modulus, G, varies approximately linearly
with the temperature, so that the two values given for G can be used to
polate
E

to

other

temperatures.

Young's

modulus,

E,

can

be

calculated as

= 2G (l+v), with Poisson's ratio v having values of typically 0.3.

efficient

of

Norton's

extraThe

co-

creep law, B, can be calculated from the data given in

Table A.l using eq. (1.2). Diffusion coefficients are represented in the

usual

way, 0 = 0oexp(-Q/RT). Surface and grain boundary energies, Ys and Yb , are


moderately temperature dependent. The data in Table A.l refer to low temperatures. High-temperature values would be slightly smaller. The effect of impurity
segregation on the interfacial energies, which can

amount

to

some

30%,

was

examined in Section 8.2.3. The grain boundary diffusion coefficient, oOb' and
the surface diffusion coefficient, oDs, are also sensitive to the segregation
of

trace

impurities.

This

is

illustrated in Table A.l by the two different

values for the graIn boundary diffusion coefficient of a-iron

taken

different sources. They differ by a factor 2.5 for oOb at 973 K.

Material

a-iron
[1,2,3]

Y-iron
[1,2]

nickel
[1,'2]

copper
[1,2]

o in 10- 29 m3
b in 10- 10 m
G(300K) in GPa
G( 800K) in GPa
A* def. eq.(1.2)
n (stress exp.)
ovo in m2 /s
Q in kJ/mole
o~bO in m3/s
from [3]:
Qb in kJ/mole
from [3]:
00
in m3 /s
so
Qs in kJ/mole
Yb in J/m 2 [4]
Ys in J/m 2

1.18
2.48
64
48
7.10 13

1.21
2.58

1.09
2.49
79
63
3.0.10 6
4.6
1.9.10- 4
284
3.5.10- 15

1.18
2.56
42
33
7.4.10 5
4.8
2.0.10-5

6.9
2.10- 4
251
1.1.10-12
10- 15
174
105
2.5.10- 9
232
0.85
2.1

4.3.10 5
4.5
1.8.10- 5
270
7.5.10- 14
159
1.1.10-10
220
2.0

115
4.4.10- 12
199
0.7
2.0

197
5.0.10- 15
104
6.10- 10
205
0.65
1.7

Sources: [1] Frost and Ashby (1982), [2] Swinkels and Ashby
(1981), [3] Stratmann et al (1983), [4] Chuang et al (1979).
Table A.l. Material parameters relatett to creep fracture.

from

two

Appendix A: Material Parameters

390

The greater value was measured by Stratmann et al (1983) on especially purified


iron

(see

Section

11.1.5). Similarly, nickel base superalloys have a smaller

6Db than pure nickel due to the addition of boron and zirconium. A value
Nimonic 80A as reported by Dyson (1979) was quoted in Section 12.4.1.

for

Some of the material parameters of the pure metals can be applied approximately
to

the commercial materials derived from the metals. Among the only moderately

sensitive or insensitive parameters are the atomic volume,

n,

the

elasticity

moduli, the interface energies and, to a certain extent, the diffusion coefficients, with the proviso that impurity segregation affects oD b markedly.
On the other hand, the creep resistance of commercial materials is usually much
greater

than

that

of pure metals. A few examples for the material parameters

appearing in Norton's power law are

shown

in

Table

A.2.

Extrapolations

to

neighboring temperatures are possible based on eq. (1.2) and using the activation energies of the diffusion coeffiCient, Qv' and (less important) the temperature dependence of the shear modulus listed in Table A.l. The creep resistance
of ferritic steels depends sensitively on prior heat
material

treatments.

As-processed

has a substantially greater creep resistance than has the same mater-

ial after it has been in service at temperatures around

530C

for

more

than

100,000 h ('used' materials).

B in MPa- n s- 1
a-iron
at 540C [1,2]
lCr-Mo-V steel at 540C [1,2]
lCr-1/2Mo steel (used) at 535C [5]
21/~Cr-1Mo steel (new) at 540C [6,7]
21/~Cr-1Mo steel (used) at 540C [6,7]
1/2Cr-l/2M6-1/~V (new) at 540 0 C [6,7]
1/2Cr-l/2Mo-l/~V (used) at 540C [7]
Y-iron
at 650C [1,2]
304 stainless steel at 650C [1,2]
316 stainless steel at 650C [1,2]
347 stainless steel at 650C [8]
Nimonic 80A at 650C [5]
at 750C [9,10]

n
6.9
6.0
8.6
6.0
10.8
9.5
14
4.5
7.0
7.9
8.1
13
4.2

Sources: [1,2] see Table A.l, [5] Riedel and Wagner (1985), [6] Bendick and
Weber (1984, private communication), [7] own measurements, [8] Needham
and Gladman (1980), [9] Dyson and Rodgers (1977), [10] Dyson (1979).
Table A.2. Parameters of Norton's law,

E = Bon,

for steels and Nimonic 80A.

Appendix B: Elastic Stress Fields at Notches, Cracks


and Grain Boundary Triple Points
Plane-strain and plane-stress problems in linear

elasticity

described

~,

by

means of the Airy stress function,

suggests

to

conveniently

the governing equation for

which is V4~ = O. The problems to be discussed are shown


symmetry

are

in

Fig.

B.l.

Their

use polar co-ordinates, rand e. In these co-ordinates,

the equation for the stress function takes the form


(B.l )
The differential operator in square brackets is the Laplace operator

in

polar

co-ordinates. Solutions are sought now in the factorized form ~ = K~ r 2 - s fee),


where the exponent s is called the eigenvalue of the problem, which

is

to

be

determined; K~ is a factor of proportionality which will remain undetermined by


the asymptotic analysis. If the unknown function fee) is taken to be dimensionless, K~ has the physical dimension (stresslength s ). Inserting the factorized
form of

into eq. (B.l) leads to the ordinary differential equation for fee):
f""

2 (s2 - 2s

2) f"

s2(2 - s)2 f

0,

(B.2)

where a prime denotes differentiation with respect to e. Such a linear equation


with constant coefficients can be

solved by trigonometric functions.

The sym-

392

Appendix B: Notches, Cracks and Triple Points

metry of Mode-I loading permits only cosines, while Mode


sines.

Insertion

f(a) = cos(ka) or f(a)

of

eq. (B.2) shows that k = sand


coefficients

adjoining

the

k = 2-s

cosines

give

(or

II

is

described

solutions

sines)

of

eq.

(B.2).

The

will be determined from the

boundary conditions of the problem considered. Before we proceed to do so,


components

by

sin(ka) with unspecified k into

the

of stress are given in terms of derivatives of the stress function.

In polar co-ordinates, eq. (3.14) takes the form:

1 04>
r or

-- +

(B.3)

o
ra

=-

The second form of these equations is valid

for

the

Obviously, the stress has a singularity of the form

factorized

form

of

4>.

~ r- s

B.1 Stress Fields at Sharp Notches and Cracks


B.l.l

The eigenvalue equation for sharp notches

From the foregoing it follows that the function f(a) at

sharp

notch

under

Mode-I tensile loading must have the form:


f(a) = a 1 cos(sa)

a 2 cos[(2-s)9J.

(B.4)

For Mode II replace the cosines by sines. The constants of integration, a 1


a2,

and

are

determined next. On the traction-free flanks of the notch, the stress


components 0 99 and 0r9 must be zero, i.e. f( 'TT-a) = f' ('TT-a) = 0. These two
boundary

conditions

Non-zero solutions

lead
exist

to two linear, homogeneous equations for a, and a 2


if the determinant of the system of equations

vanishes. This leads to the eigenvalue equation for s:


sin[2(1-s)('TT-a)J (1-s) sin[2('TT-a)J

0,

where the plus and minus signs apply for Mode.! and Mode II, respectively.

(B.5)

393

Appendix B: Notches, Cracks and Triple Points

0.5

0.500

0.488

0.456

0.384

0.181

Table B.l. The exponent of the stress singularity,


a r- s , at a sharp notch with an included angle 2a.

Table B.1 shows the smallest positive eigenvalue obtained numerically from
(B.5)

eq.

for

Mode I for various included notch angles, 2a. This eigenvalue gives
the stress singularity, r- s , at the notch tip. The boundary conditions on the
notch flanks demand that the constants of integration obey the relation
a 1 /a 2

= -

cos[(2-s)(w-a)]/cos[s(w-a)]

(B.6)

for Mode I. The absolute values of a 1 and a 2 remain undetermined by the

asymp-

totic analysis.
B.1.2

Crack-tip fields

The special case of a crack is characterized by a


term

in

the

O. In this case, the second

eigenvalue eq. (B.5) vanishes. Therefore Mode I and Mode II have

the same eigenvalues, 1/2, 0, -1/2, -1, -3/2 etc. The first of these gives

the

well-known inverse-sQuare-root singularity at crack tips. The unknown factor in


the asymptotic solution is conventionally called the stress

intensity

factor,

KI or KII , in crack problems. The stress function is then found to be:


KI (2w)-1/2 r 3/2 [cos(8/2)

~
=

(4/3) K (2w)-1/2 r 3/2

(1/3)cos(38/2)]

} for Mode I

(B.7)

} for Mode II

(B.8)

COS 3 (8/2)

~ = - KII (2w)-1/2

r 3/2 [sin(8/2) + sin(38/2)]

~ - KII (2w)-1/2 r 3/2 4 sin(8/2) cos 2 (8/2)


The stress components follow from

by eq. (B.3) in the form of eqs. (3.26)

or

(7.1). Their angular parts, f ij = a ij 12wr/K I , are listed in Table B.2. The
Cartesian components are derived from the polar components by the usual tensor
transformation

relations.

The

zero in plane stress and is a33

component
=

v(a 11

perpendicular to the plane, a33 , is


a22 ) in plane strain. Von Mises equi-

394

Appendix B: Notches, Cracks and Triple Points

lj co~

36

36
cOT

= lj" co~ + lj

. 6

lj Sl~ + lj
0 11

sin'26

lj co~

- lj

3 sl.n36
2

+ lj

3 sin! _ 3 sin36

lj

i 36
s n'2

~ co~

cos~ (1 - sin~ sin~6)

sin~ (2

*cos~6

lj

t::

co~ co~)

cos~ (1

sin~ sin~6)

6
6
36
sin'2 cos'2 cos2

6
cos'2

36
cOT

co~ (1 - sin~ sin~6)

Si~

cos~ [(1_2\1)2
co~ [1

3 (sin~)2J1/2

3 (sin~)2J1/2

. 6

sm2

\I

[(1-2\1)2(sin~)2
[3 - 8 (sin~)2

3 -

+
+

sin 2 6J 1/2

9 (sin~)4J1/2

. 6 - [1 - lj
3 sln
. 26 J 1 12
- sln'2

Table B.2. The angular functions, f .. (6) = 0 .. 12wr/K I , of elastic crack-tip


lJ
lJ
fields; 01 is the maximum principal tensll'e stress, 0e is the von Mises
equivalent tensile stress for plane strain (p-E) and plane stress (p-o).

valent stresses and the maximum principal tensile


Alternative

expressions

stresses

are

listed

also.

for the trigonometric functions, as well as displace-

ments are given by Rice (1968b), by Tada, Paris and Irwin (1973) and

by

Broek

(1982).
In Mode III, the stress components are derivable from the Mode-III stress function ~, which near the crack tip has the form
stress components are
KIll

I2"1iT
and 013

sin~
2

a~

ar

~ =

__

KIll (2r/w)

KIll
6
- - cos..
12wr
2

1/2

cos(6/2). The

(B.9)

395

Appendix B: Notches, Cracks and Triple Points

B.2 The Stress Singularity at a'lliple Junction


of Sliding Grain Boundaries
The triple junction of sliding boundaries under tensile loading
Fig.

B.1.

The

shown

wa~

in

angular part of the Airy stress function, f(a), satisfying eq.

(B.2) and having the symmetry of the triple junction must have the form
f( 9)

a 1 cos(sa)

a 2 cos[(2-s)aJ.

= b 1 cos[s(1[-a)J

b 2 cos[(2-s)(1[-a)]

for lal < 1[-CX


for

Ia I

(B.10)

> 1[-CX.

Mode II will not be considered here. The eigenvalue s and the constants of integration

are determined (again apart from a constant) from the conditions that

the shear stress on the inclined boundaries is relaxed (are


of

0 on either

side

e = 1[-CX) , that 0aa must be continuous at a = 1[-CX, and that the displacement

component u a must be continuous. The boundary conditions which refer to stress


imply: f' = 0, f and f' continuous at 1[-CX. The evaluation of the displacement
continuity condition requires a knowledge of the relation between
and

displacement

stress function. As a first step, displacement is related to strain by eq.

(3.2), which can be integrated to give:


(B.11)

where 90

related

to

lal < 1[-CX and ao = 1[ for lal > 1[-CX. Strain is


stress by Hooke's law. The author has carried out the analysis for

0 in the domain

incompressible material only (v

1/2).

There results

3 (2-s)
4 (1-s) E
Inserting stresses from eq. (B.3) and recalling that f' is
across

(B.12)

already

continuous

the sliding boundary leads to the requirement that ff(a)da must be con-

tinuous also. Together with the three stress-related boundary conditions,

this

establishes a system of four linear homogeneous equations for a 1 , a 2 , b 1 and b 2


with the eigenvalue s. Setting the determinant equal to zero leads to the
eigenvalue

equation

also tabulated there.

given in eq. (7.20) in the main text. The eigenvalues are

Appendix C: Calculation of C* forrest Specimen


Configurations
For power-law viscous materials described by the material law
(C.l)

the C*-integral depends on load and specimen geometry according to

the

relat-

ions given in Section 20.2.1. These expressions contain dimensionless functions


of the specimen geometry, which have been calculated by Kumar, German and

Shih

(1981) using the finite element method. Some of their results are reported now.
The numerical data were fitted by functions of the general form
(C.2)

F(x)

Here, x

a/W with a and W being defined in Fig. C.1, F(x)

represents

any

of

the functions of the specimen geometry, h 1 (a/W,n), h 2 (a/W,n), h 3 (a/W,n) or


g2(a/W,n), which were defined for compact specimens in Section 20.2.1, and the
ai's

and

are adjustable parameters. Depending on which of the functions hi

and g2 is fitted, it is convenient to choose fixed values for part of the parameters and to let the rest be determined by a least-square-fit computer program. In this way, the parameters shown in the tables were obtained. The values
that

were

prescribed,

rather

than determined by the best-fit procedure, are

marked with an asterix. The resulting functions reproduce the numerical


of

Kumar

values

et al tb within better than 3%. The results for the compact specimen

in plane strain are shown in Fig. C.2.


The formulas to calculate C*, the load-line deflection rate,
placement

rate

at

the edge of the specimen,

6e ,

A,

and

the

dis-

are given in the captions to

the tables. The function n appearing in those formulas is defined as

[4a2 + 4a + 2J 1/2 - 2a - 1

n
where a

for the compact specimen


for the SENT-specimen,

a/(W-a) and the single-edge cracked plate in tension (SENT) is

inFig.C.1.

(C.4)

shown

397

Appendix C: C* in Test Specimens

0'00

tL112

1L
1

l.d/2
FIg. C.l. Test specimen configurations.

=
N

0.50
o/W

0.75

0.25

1.00

-7

0.50
o/W

-7

0.75

0.50
o/W

--'>

1.00

0.25

0.50
o/W

0.75
-7

1.00

0.75

1.00

Fig. C.2. Geometrical functions for CT-specimen in plane strain.


Square symbols from Kumar et al (1981).

398

Appendix C: C* in Test Specimens

10

16

13

20

for h,:
2.374 3.512 1.374 0.120 0.090 0.118 1.564 1.317 0.831
ao
al
0.039 -7.581 -2.867 -0.442 -0.260 -0.431 -10.25 -8.842 -5.634
-2.674 9.116 4.849 1.040 0.115 0.159 24.40 21.36 13.75
a2
a3
1.837 -3.597 -2.007 0.461 1.134 1.103 -23.79 -21.56 -14.43
8.926 8.457 6.115
0*
0*
0*
0*
0*
0*
a4
a 5 , a6
0*
0*
0*
0*
0*
0*
0*
0*
0*
p
-0.027 0.469 2.090 2.373 2,541 1.363 1.629 2.142
0*
for h 2 :
a
20.14 12.00 11.95 2.293 0.314 2.491 2.118 1.686 0.576
0
-41.43 -25.65 -36.27 -7.804 -0.684 -15.38 -13.94 -11.42 -4.374
al
a2
39.69 24.16 45.51 11.65 -0.348 35.41 33.66 28.24 12.59
-13.02 -6.777 -18.10 -3.710 2.833 -34.11 -34.86 -30.42 -16.65
a3
a4
0*
0*
0*
0*
0*
13.39 14.60 13.29 8.999
0*
0*
0*
0*
0*
0*
0*
0*
0*
a 5 , a6
p
0.285 0.419 0.552 1.707 2.995 2.412 2.767 3.095 4.179
for h3:
-0.467 8.026 8.103 1.554 0.283 0.237 0.225 1.323 0.467
ao
al
4.098 -16.'03 -24.34 -5.352 -0.601 -0.641 -0.633 -9.017 -3.563
-2.082 14.73 30.93 8.008 -0.434 -0.449 -0.522 22.46 10.30
a2
a3
2.758 -3.733 -12.22 -2.260 2.526 2.261 2.132 -24.43 -13.65
10.74 7.357
0*
0*
0*
0*
0*
0*
0*
a4
a 5 , a6
0*
0*
0*
0*
0*
0*
0*
0*
0*
p
2.194 0.401 0.560 1.756 2.868 3.282 3.487 3.064 4.135
for g2:
1.082 1.886 -0.338 -0.282 -0.227
a0
-0.860 -4.244 1.135 1.006 1.132
al
1.192 7.262 -1.027 -1.263 -1.731
a2
0*
0*
-0.611 -3.839
a3
0*
a4
0*
0*
0*
0*
0*
-0.299 -0.287 -0.220
0*
0*
a5
a6
0*
0*
-0.891 -1. 117 -1.368
1*
1*
0*
0*
1*
P

Table

C.,.

-0.182
0.675
-1.480
0*
0*
-0.283
-1.396
1*

-0:158
0.691
-1.589
0*
0*
-0.264
-1.443
1*

-0.188
1.058
-1.867
0*
0*
-0.211
-1.448
1*

-0.124
0.471
-1.413
0*
0*
-0.279
-1.427
1*

Parameters from eq. (C.2) for CT-speclmen in plane strain.


C*

hl (W-a) B [onet/(1.455 n)]n+l

6e = h2
A = h3

a B [onet/(1.455 n)]n
a B [onet/(1.455 n)]n

g2 1:. 0net

399

Appendix C: C* in Test Specimens

10

13

for h 1 :
3.235 2.252 2.009 0.498 0.554 0.041 0.103
ao
a1
-1.863 -3.069 -0.732 -1.075 -1.571 .0008 -0.250
-0.177 3.088 0.775 1.999 2.967 -0.095 -0.136
a2
a3
0.885 -0.698 0.302 -1.013 -1.247 0.941 2.477
0*
0*
0*
0*
0*
0*
0*
a
a4' 5
a6
0.836 0.558 1.949 -0.399 0.082 0.444 3.465
2*
2*
-0.25*
0*
0*
0.75* 0.75*
P

16

20

0.113
0.367
0.021
1.909
0*
2.616
2*

0.074
0.248
0.076
1.041
0*
1.283
2*

for h2 :
a
3.813 1.179
0
1.347 0.650
a1
a2
-2.705 -2.565
3.384 4.528
a3
0*
0*
a 4 a6
0.104 0.073
a5
1.5*
2.5*
P

0.574 0.248 0.240 0.297 0.512 0.270 0.131


0.133 0.274 -0.060 -0.587 -2.098 -0.954 -0.365
0.214 -0.206 0.051 0.439 3.404 2.285 1.357
1.583 1.639 1.630 1.455 -0.576 -0.545 -0.313
0*
0*
0*
0*
0*
0*
0*
0.029 .0062 .0023 .0036 .0095 -0.092 -0.209
1.44*
0*
2.5*
2.5*
2.5*
2.5*
2.5*

for h3:
3.469 1.641
ao
a1
1.087 -1.338
-2.079 -0.200
a2
a3
2.821 3.109
0*
0*
a
6
a 4
a5
0.252 0.137
p
1.5*
2.5*

0.398 0.141 0.137 0.694 0.226 0.107 0.090


0.044 0.156 0.129 -2.554 -1.625 -0.341 -0.286
0.386 -0.336 -0.358 3.151 2.226 0.244 0.126
1.174 1.594 1.502
0*
0*
0.909 0.876
0*
0*
0*
0*
0*
0*
0*
0*
0*
0.031 0.010 -.0027 0.049 0.026
2.826 2.824
2.5*
2.8*
2.5*
3*
3*

for g2:
a
0.031 -.0077 0.093 0.087 0.197 1.274 1.488 1.247 1.558
0
0.776 0.367 -0.312 0.343 -0.878 3.209 3.467 3.205 3.738
a1
a2
0.155 -1.461 0.345 2.521 2.152 -5.276 -6.404 -5.146 -7.231
-0.162 2.186 1.155 -1.630
0*
2.247 2.936 2.142 3.457
a3
0*
0*
0*
0*
0*
a
0*
0*
0*
0*
6
a 4
0*
0.074
0*
0.018 0.034
0*
0*
0*
0*
a5
4.108 3.423
-0.242 -0.309 -0.183 -0.310
1*
1*
3*
P

Table C.2. Parameters from eq. (C.2) for CT-specimen in plane stress.
n+1

C* = h1 (W-a) B [Onet / (1.071 n)]


= g2 /). 0net
n

~e

h2 a B [Onet / (1.071 n)]

i;

h3 a B [Onet/(1.071 n)]n

Appendix C: C* in Test Specimens

400

10

13

16

20

for h 1 :
0.344 3.826 2.095 1.520 0.837 0.655 0.190 0.058 0.044
ao
a1
-1.151 -6.064 -3.982 -2.713 -0.749 -1.142 -0.481 -0.096 -0.142
5.676 3.041 2.172 -0.521 -3.212 -1.300 0.025 -0.194 0.093
a2
a3
-3.322 1.624 1.567 3.260 4.451 2.619 0.520 0.353 0.047
0*
0*
0*
0*
0*
0*
0*
0*
0*
a4' a6
a5
-0.084
-.0042
-.0033
0.406
0.147
-0.173
0.052
-0.023 -0.073
2.5*
2.5*
p
1*
1*
0.76*
2*
2.5*
1.5*
1.5*
for h 2 :
a
0.307 2.301 3.642 -0.407 1.091 0.534 0.109 0.089 0.057
0
-0.561 -4.439 -7.349 8.298 -2.109 -0.747 -0.687 -0.194 -0.183
a1
a2
2.089 3.511 4.392 -22.44 -1.039 -1.803 -0.742 -0.150 0.114
2.634 -0.154 2.735 17.32 3.462 2.914 1.030 0.417 0.069
a3
a 4 , a6
0*
0*
0*
0*
0*
0*
0*
0*
0*
0.049 -0.631 0.245 -0.092 -0.280 -0.321 -.0093 -0.012 -0.014
a5
2*
2*
1*
0.5*
0.5*
2*
2.75* 0.455* 1*
P
for h3:
0.521 0.734 0.696 0.339 0.449 0.389 0.054 0.023 0.014
ao
a1
1.766 -0.285 1.117 -0.226 -0.897 -1.002 -0.106 -0.040 -0.038
-4.594 -2.686 -0.949 -1.854 -0.678 0.077 -0.130 -0.082 .0038
a2
a3
4.492 3.821 2.636 2.678 1.821 0.984 0.297 0.152 0.040
0*
0*
0*
0*
0*
0*
0*
0*
0*
a 4 , a6
a5
-0.114
-.0026
-.0029
-.0039
0.010 0.010 .0061 -.0038 -0.030
2*
2*
2*
2*
1.5*
1*
2.5*
2.75* 2.75*
P

Table C.3. Parameters from eq. (C.2) for SENT-specimen in plane strain.
C* = hl (W-a) (a/W) B [Onet/(1.455 n)]

6e

= h2

~c =

6C is the remote load


6c = A- ~c' where Ais

n+l

[Onet / (1.455 n)]


n
h3 a B [onet / (1.455 n)]

point

a B

displacement

rate

due

to

the

crack,

i.

the total displacement rate and ~c =13 B L [l3o~/2


1s the displacement rate in the absence of the crack; o~ is the remote

applied stress.

References
Abou Zahra, A.A. and Schroder, H. (1982) J. Nucl. Mater. 101, 97-103.
Abramovitz, M. and Stegun, LA. (1968) Handbook of Mathematical Functions,
Dover Publications, Inc., New York.
Ainslie, N.G. and Seybolt, A.U. (1960) J. Iron Steel Inst. 194, 341-350.
Antlerson, P.M. and Rice, J.R. (1985) Acta Metall. 33, 409-422:
Argon, A.S. (1975) (Ed.) Constitutive Equations inlPlasticity, The MIT Press,
Cambridge, Massachusetts and London, England.
Argon, A.S. (1982) In: Recent Advances in Creep and Fracture of Engineering
Materials and Structures, B. Wilshire and D.R.J. Owen, Eds., Pineridge
Press, Swansea, U.K., pp. 1-52.
Argon, A.S., Chen, I.-W. and Lau, C.W. (1980) In: Creep-Fatigue-Environment
Interactions, R.M. Pelloux and N.S. Stoloff, Eds., The Metallurgical
Society of AIME, pp. 46-85.
Argon, A.S., Chen, I.-W. and Lau, C.W. (1981) In: Three-Dimensional
Constitutive Relations and Ductile Fracture, S. Nemat-Nasser, Ed.,
North-Holland Publishing Company, Amsterdam New York Oxford, pp. 23-49.
Argon, A.S. and 1m, J. (1975) Metall. Trans. 6A, 839-851.
Arzt, E., Ashby, M.F. and Verrall, R.A. (1983r-Acta Metall. ll, 1977-1989.
Ashby, M.F. (1969) Scripta Metall. 3, 843-848.
Ashby, M.F. (1972) Acta Metall. 20,-887-897.
Ashby, M.F. (1977) In: Fracture 1977, D.M.R. Taplin, Ed., University of
Waterloo Press, Waterloo, Canada, Vol. 1, pp. 1-14.
Ashby, M.F., Bahk, S., Bevk, J. and Turnbull, D. (1980) Progr. Mater. Sci. 25,
1-34.
Ashby, M.F. and Dyson, B.F. (1984) In: Advances in Fracture Research '84 Proceedings of ICF6, Vol. 1, S.R. Valluri et al., Eds., Pergamon Press,
Oxford, pp. 3-30.
Ashby, M.F., Gandhi, C. and Taplin, D.M.R. (1979) Acta Metall. 27, 699-729.
ASTM-E 399-78: Standard Method of Test for Plane-Strain Fracture-Toughness of
Metallic Materials. Annual Book of ASTM Standards, 512-533.
ASTM-E 813-81: Standard Test for J I8 , a Measure of Fracture Toughness. Annual
Book of ASTM Standards, Part 1 , pp. 810-828.
B<aik, S. and Raj, R. (1982a) Metall. Trans. 13A, 1207-1214.
Baik, S. and Raj, R. (1982b) Metall. Trans. 13A, 1215-1221.
Baik, S. and Raj. R. (1983) Scripta Metall. ~ 1087-1090.
Bailey, R.W. (1926) J. Inst. Metals 35, 27-4~
Balluffi, R.W. (1980) In: Grain-Boundary Structure and Kinetics, ASM Materials
Science Seminar, 1979, American Society for Metals, Metals Park, Ohio, pp.
297-330.
Banthia, V. and Mukherjee, S. (1985) Int. J. Fracture 28, 83-101.
Bassani, J.L. (1981) In: Creep and Fracture of Engineering Materials and
Structures, B. Wilshire and D.R.J. Owen, Eds., Pineridge Press, Swansea,
pp. 329-344.
Bassani, J.L. (1983) In: Elastic-Plastic Fracture: Second Symposium, vol. I.
Inelastic Crack Analysis, ASTM STP 803, C.F. Shih and J.P. Gudas, Eds.,
American Society for Testing and Materials, pp. 1-532-1-550.
Bassani, J.L. and McClintock, F.A. (1981) Int. J. Solids Structures 17,
479-492.
Bassani, J.L. and Vitek, V. (1982) In: Proceedings of the 9th U.S. National
Congress on Theoretical and Applied Mechanis, L.B. Freund and C.F. Shih,
Eds., pp. 127-133.
Bathe, K.-J. and Wilson, E.L. (1976) Numerical Methods in Finite Element
Analysis, Prentice Hall, Englewood Cliffs, New Jersey.
Bazant, Z.P. and Estenssoro, L.F. (1977) In: Fracture 1977, (Proceedings of

402

References

ICF4), D.M.R. Taplin, Ed., Vol.3, University of Waterloo Press, Waterloo,


Canada, pp. 371-383.
Becker, R. and Doring, W. (1935) Ann. Phys. 24, 719-752.
Beere, W. (1978) Phil. Mag. A 38, 691-696. -Beere, W. (1980a) Acta Metall.:28, 143-150.
Beere, W. (1980b) J. Mater. Sci:-15, 657-669.
Beere, W. (1981) Res Mechanica Letters 1, 79-83.
Beere, W. (1982) Metal Sci. 16, 223-227~
Beere, W. and Roberts, G. (1982) Acta Metall. 30, 571-580.
Beere, W. and Speight, M.V. (1978) Metal Sci. 12, 172-176.
Bendick, W. and Weber, H. (1984) In: Arbeitsgemeinschaft fUr warmfeste Stahle
und Arbeitsgemeinschaft fur Hochtemperaturwerkstoffe, 7. Vortragsveranstaltung "Langzeitverhalten warmfetser Stahle und Hochtem~erat\;lr
werkstoffe", Verein Deutscher Eisenhuttenleute, Dusseldorf, pp. 1-12.
Benthem, J.P. (1977) Int. J. Solids Structures 13, 479-492.
Benthem, J.P. (1980) Int. J. Solids Structures 16, 119-130.
Benthem, J.P. and Douma, T.H. (1980) Graphs of the 3-D State of Stress at the
Vertex of a Quarter-Infinite Crack in a Half-Space. Report of Delft
University of Technology, WTHD 123, Delft, The Netherlands.
Bernardini, J., Gas, P., Hondros, E.D. and Seah, M.P. (1982) Proc. Roy. Soc.
Lond. A 379, 159-178.
Bieber, C.G. and Decker, R.F. (1961) Trans. AIME 221 629-636.
Bilby, B.A. and Eshelby, J.D. (1968) In: Fracture:-An Advanced Treatise, vol.
1, Microscopic and Macroscopic Fundamentals, H. Liebowitz, Ed., Academic
Press, New York and London, pp. 99-182.
Binder, K. and Stauffer, D. (1976) Advances in Physics 25, 343-396.
Boettner, R.C. and Robertson, W.D. (1961) Trans. AIME 221, 613-622.
Bohnenkamp, K. and Engell, H.-J. (1962) Arch. Eisenhuttenwes. 33, 359-367.
Bor isov, V. T., Gol1kov, V. M. and Scherbedinskiy (1964) Fiz. metal. metalloved.
17, 881-885.
Brailsford, A.D. and Bullough, R. (1972) J. Nuclear Mater. 44, 121-135.
Braski, D.N., Schroeder, H. and Ullmaier, H. (1979) J. Nucl:-Mater. 83 265-277.
Bricknell, R.H. and Woodford, D.A. (1982) Acta Metall. 30, 257-264. -Bricknell, R.H. and Woodford, D.A. (1984) Metal Sci. 18-;-265-271.
-Brinkman, J.A. (1955) Acta Metall. 3, 140-145.
Broek, D. (1982) Elementary Engineering Fracture Mechanics, Martinus Nijhoff
Publishers, The Hague Boston London.
Budiansky, B. and Hutchinson, J.W. (1978) J. Appl. Mech. 45, 267-275.
Budiansky, B., Hutchinson, J.W. and Slutsky, S. (1982) In:-Mechanics of Solids,
The Rodney Hill 60th Anniversary Volume, H.G. Hopkins and M.J. Sewell,
Eds., Pergamon Press, Oxford, pp. 13-45.
Budiansky, B. and Rice, J.R. (1973) J. Appl. Mech. 40, 201-203.
Burke, M.A. and Nix, W.D. (1979) Int. J. Solids Structures 15, 55-71.
Burt, H., Dennison, J.P. and Wilshire, B. (1979) Metal Sci.-'3, 295-300.
Burt, H., Elliott, I.C. and Wilshire, B. (1981) Metal Sci. 1~ 421-424.
Burton, B. (1973) Mater. Sci. Eng. 11, 337-343.
-Burton, B. (1977) Diffusional Creep-of Polycrystalline Materials; Diffusion and
Defect Monograph Series, Trans. Tech. Publications, Ohio.
Burton, B. and Beere, W.B. (1981) Phil. Mag. 43, 1561-1568.
Burton, B. and Heald,' P.T. (1975) Phil. Mag. 32, 1079-1081.
-Cane, B.J. (1976) Metal Sci. 10, 29-34.
Cane, B.J. (1979) Metal Sci. 13, 287-294.
Cane, B.J. (1981) Metal Sci. 15, 302-310.
Cane, B.J. and Greenwood, G.W:-(1975) Metal Sci. ~, 55-60.
Cane, B.J. and Middleton, C.J. (1981) Metal Sci. 15, 295-301
.Chang, H.C. and Grant, N.J. (1953) Trans. AIME 19~ J. Metals, 1175-1180.
Chang, H.C. and Grant, N.J. (1956) Trans. AIME 206, J. Metals, 544-551.
Chastell, D.J. and Flewitt, P.E.J. (1979) Mater:-Sci. Engin. 38, 153-162.

References

Chen,
Chen,
Chen,
Chen,
Chen,
Chen,

403

C.W. (1956) Trans. AIME 206, J. Metals, 1416-1411.


C.W. and Machlin, E.S. (1956) Acta Metall. 4, 655-656.
C.W. and Machlin, E.S. (1951) Trans. AIME 209, J. Metals, 829-835.
I.-W. and Argon, A.S. (1919) Acta Metall. ~ 185-191.
I.-W. and Argon, A.S. (1981a) Acta Metall.~9, 1321-1333.
I.-W. and Argon, A.S. (1981b) In: Creep and~racture of Engineering
Materials and Structures, B. Wilshire and D.R.J. Owen, Eds., Pineridge
Press, Swansea, U.K., pp.289-302.
Chen, I.-W. and Argon, A.S. (1981c) Acta Metall. 29, 1159-1168.
Chen, I.-W. and Yoo, M.H. (1984) Acta Metall. 32,-'499-1508.
Chen, R.T and Weertman, J.R. (1984) Mater. Sci:-Eng. 64, 15-25.
Chen, S.-H., Takasugi, T. and Pope, D.P. (1983) Metal~ Trans. 14A, 511-580.
Chuang, T.-J. (1982) J. Amer. Ceramic Soc. 65, 93-103.
--Chuang, T.-J., Kagawa, K.I., Rice, J.R. and Sills, L.B. (1919) Acta Metall. 21,
265-284.
Chuang, T.-J. and Rice, J.R. (1913) Acta Metall. 21, 1625-1628.
Claeys, S.F and Jones, J.W. (1984) Metal Sci. 18,~32-438.
Coble, R.L. (1963) J. Appl. Phys. 34, 1619-168~
Cocks, A.C.F. (1985) Acta Metall. 33, 129-131.
Cocks, A.C.F. and Ashby, M.F. (1980) Metal Sci. 14, 395-402.
Coffin, L.F. (1912) Metall. Trans. 3, 1111-1188.-Coffin, L.F. (1911) In: Fracture 1971, vol. 1, D.M.R. Taplin, Ed., University
of Waterloo Press, Waterloo, Canada, pp. 263-292.
Collins, A.L.W., Sidey, D. and Taplin, D.M.R. (1919) Canad. Metall. Qu. ~,
231-238.
Cottrell, A.H. (1958) Trans. Metall. Soc. AIME 212, 192-203.
Crank, J. (1956) The Mathematics of Diffusion, Oxford at the Clarendon Press,
Oxford.
Crossland, I.G. and Harris, J.E. (1914) Metal Sci. 13, 55-59.
Crossman, F.W. and Ashby, M.F. (1915) Acta Metall. 23, 425-440.
Davidge, R.W. (1919) Mechanical Behaviour of Ceramics, Cambridge University
Press, Cambridge.
Davies, P.W. and Dennison, J.P. (1915) Metal Sci. 9, 319-323.
Davies, P.W., Dennison, J.P. and Evans, H.E. (1966) J. Inst. Metals 94,
210-295.
Davies, P.W., Dennison, J.P. and Evans, H.E. (1961) J. Inst. Metals 95,
231-234.
Davies, P.W., Dennison, J.P. and Sidey, D. (1913) J. Inst. Metals ~, 153.
Davies, P.W., and Dutton, R. (1966) Acta Metall. 14, 1138-1140.
Davies, P.W. and Evans, R.W. (1965) Acta Metall. 13, 353-361.
Davies, P.W. and Williams, K.R. (1969) Metal Sci.~. 1, 220-221.
Davies, P.W., Williams, K.R. and Wilshire, B. (1968) Phil. Mag. 18, 191-200.
deLorenzi, H.G. and Shih, C.F. (1983) Int. J. Fracture 21, 195-220.
Delph, T.J. (1980) Trans. ASME, J. Eng. Mater. Technology 102, 321-336.
Dennison, J.H. Holmes, P.D. and Wilshire, B. (1918) Mater.-SCi. Eng. 33, 35-41.
Dennison, J.P. and Wilshire, B. (1911) In: Fracture 1911, Vol. 2, D.M~.
Taplin, Ed., Waterloo University Press, Waterloo, Canada, 635-639.
Detampel, V. (1986) PhD-dissertation, RWTH Aachen.
Don, J. and Majumdar, S. (1986) Acta Metall. 34, 961-961.
Dowling, N.E. and Begley, J.A. (1916) In: Mechanics of Crack Growth, ASTM STP
590, American Society for Testing and Materials, pp. 82-103.
Dunlop, G.L., Twigg, R.J. and Taplin, D.M.R. (1918) Scand. J. Metall. I,
152-160.
Duva, J.M. (1984) J. Eng. Mater. Technology 106, 311-321.
Duva, J.M. and Hutchinson, J.W. (1984) Mechanics of Materials 1, 41-54.
Dyson, B.F. (1916) Metal Sci. 10, 349-353.
Dyson, B.F. (1919) Canad. Metall. Qu. 18, 31-38.
Dyson, B.F. (1982) Acta Metall. 30, 1639-1646.

404

References

Dyson, B.F. (1983) Scripta Metall. 17, 31-37.


Dyson, B.F. and Hondros, E.D. (1984~In: Advances in Fracture Research
'84 - Proceedings of ICF6, Vol. 6, S.R. Valluri et aI, Eds., Pergamon
Press, Oxford and New York, pp. 3753-3772.
Dyson, B.F. Loveday, M.S. and Rodgers, M.J. (1976) Proc. Roy. Soc. Lond. A 349,
245-259.
Dyson, B.F. and McLean, D. (1972) Metal Sci. 6, 220-223.
Dyson, B.F. and McLean, D. (1977) Metal Sci. 11, 37-45.
Dyson, B.F. and McLean, M. (1983) Acta Metall. 31, 17-27.
Dyson, B.F. and Rodgers, M.J. (1974) Metal Sci.-a, 261-266.
Dyson, B.F. and Rodgers, M.J. (1977) In: Fracture 1977, Proceedings of ICF4,
vol. 2, D.M.R. Taplin, Ed., University of Waterloo Press, Waterloo,
Canada, pp. 621-626.
Ebi, G., Riedel, H. and Neumann, P. (1986) In: Fracture Control of Engineering
Structures (ECF6), Vol. 3, H.C. van Elst and A.Bakker, Eds., Engineering
Materials Advisory Services Ltd., Warley, England, pp. 1587-1598.
Edmunds, T.M. and Willis, J.R. (1977) J. Mech. Phys. Solids 25, 423-455.
Edward, G.H. and Ashby, M.F. (1979) Acta Metall. 27, 1505-1508.
Ehlers, R. (1981) PhD thesis. Fakultat fur BergbaU-und Hilttenwesen der
Rheinisch-Westfalischen Technischen Hochschule Aachen.
Ehlers, R. and Riedel, H. (1981) In: Advances in Fracture Research, Proceedings
of ICF5, vol. 2, D. Francois et aI, Eds., Pergamon Press, Oxford and New
York, pp. 691-698.
Elber, W. (1970) Eng. Frature Mech. 2, 37-45.
Elliott, I.C. and Wilshire, B. (1979) J. Microscopy 116, 39-45.
Ellison, E.G. and Harper, M.P. (1978) J. Strain Analysis 13, 35-51.
EI-Magd, E. and Jager, K. (1984) Arch. Eisenhilttenwes. 55~277-280.
Erhart, H. and Grabke, H.J. (1981) Metal Sci. 15, 401-408.
Ericsson, T. (1979) Canad. Metall. Qu. 18, 177=195.
Ermi, A.M. and Moteff, J. (1983) Trans.-XSME, J. Eng. Mater. Technology lQ2,
21-30.
Eshelby, J.D. (1957) Proc. Roy. Soc. Lond. A 241, 376-396.
Estrin, Y. and Mecking, H. (1984) Acta Metall~2, 57-70.
Evans, A.G. (1978) Acta Metall. 26, 1845-1853. -Evans, A.G. (1982) In: Recent Advances in Creep and Fracture of Engineering
Materials and Structures, B. Wilshire and D.R.J. Owen, Eds., Pineridge
Press, Swansea, U.K., pp. 53-133.
Evans, A.G., Rice, J.R. and Hirth, J.P. (1980) J. Amer. Ceramic Soc. 63,
368-375.
Evans, H.E. (1971) Phil. Mag. 23, 1101-1112.
Evans, H.E. (1984) Mechanisms of Creep Rupture, Elsevier Applied Science
Publishers, London and New York.
Evans, H.E. and Knowles, G. (1980) Metal Sci. 14, 262-266.
Evans, H.E. and Skelton, R.P. (1969) Metal Sci~J. 3, 152-155.
Evans, R.W., Beden, I. and Wilshire, B. (1984) In: Creep and Fracture of
Engineering Materials and Structures, D. Wilshire and D.R.J. Owen, Eds.,
Pineridge Press, Swansea, Part II, pp. 1277-1290.
Evans, R.W., Parker, J.D. and Wilshire, B. (1982) In: Recent Advances in Creep
and Fracture of Engineering Materials and Structures, B. Wilshire and
D.R.J. Owen, Eds., Pineridge Press, Swansea, pp. 135-184.
Exner, H.E. and Arzt, E. (1983) In: Physical Metallurgy, Third Edn., R.W. Cahn
and P. Haasen, Eds., Elsevier Science Publishers BV, Amsterdam Oxford New
York Tokyo, Chapter 30, pp. 1885-1912.
Feder, J., Russell, K.C., Lothe, J. and Pound, G.M. (1966) Adv. Phys. lZ,
111-179.
Feltham, P. and Meakin, J.D. (1959) Acta Metall. I, 614-627.
Fields, R.J. and Ashby, M.F. (1976) Phil. Mag. 33, 33-48.
Fields, R.J., Weerasooriya, T. and Ashby, M.F. (1980) Metall. Trans. ~,

References

405

333-347.
Fisher, J.C. (1948) J. Appl. Phys. 19, 1062-1067.
Fleck, R.G., Taplin, D.M.R. and Beevers, C.J. (1975) Acta Metall. 23, 415-424.
Floreen, S. (1983) In: Elastic Plastic Fracture: Second Symposium, Vol.
I-Inelastic Analysis, ASTM STP 803, C.F. Shih and J.P. Gudas, Eds.,
American Society for Testing and Materials, pp. 1-708-1-720.
Floreen, S. and Kane, R.H. (1980) Fatigue Eng. Mater. Structures 2, 401-412.
Floreen, S. and Kane, R.H. (1982) Metall. Trans. 13A, 145-152.
Floreen, S. and White, C.J. (1981) Metall. Trans.-r2A, 1973-1979.
Floreen, S. and Raj, R. (1985) In: Flow and Fracture-at Elevated Temperatures,
R.Raj, Ed., American Society for Metals, Metals Park, Ohio, pp. 383-405.
Frenkel, J. (1946) Kinetic Theory of Liquids, Oxford University Press, Oxford.
Friedel, J. (1967) Dislocations, Pergamon Press, Oxford and London, Chapter 8.
Frost, H.J. and Ashby, M.F. (1977) In: Fundamental Aspects of Structural Alloy
Design, R.I. Jaffee and B.A. Wilcox, Eds., Plenum Press, New York and
London, pp. 27-58.
Frost, H.J. and Ashby, M.F. (1982) Deformation-Mechanism Maps, The Plasticity
and Creep of Metals and Ceramics, Pergamon Press, Oxford, New York.
Gandhi, C. and Ashby, M.F. (1979a) Scripta Metall. 13, 371-376.
Gandhi, C. and Ashby, M.F. (1979b) Acta Metall. 27,-'565-1602.
Gandhi, C. and Raj. R. (1982) Acta Metall. 30, 505-511.
Gao, Y.-C. and Hwang, K.-C. (1981) In: Advances in Fracture Research,
Proceedings of ICF5, vol.2, D. Francois et al., Eds., Pergamon Press,
Oxford and New York, pp. 669-682.
Gas, P., Poize, S. and Bernardini, J. (1986) Acta Metall. 34, 395-403;
George, E.P. (1985) The Effects of Impurities on the Hot Tensile Ductility of
Iron. PhD-dissertation, University of Pennsylvania, Philadelphia.
Gharemani, F. (1980) Int. J. Solids Struct. 16, 847-862.
Gibeling, J.C. and Nix, W.D. (1980) Mater. Sci. Eng. 45, 123-135.
Gifkins, R.D. (1956) Acta Metall. 4, 98-99.
-Gittins, A. and Williams, H.D. (1967) Phil. Mag. ~, 849-851.
Gittus, J. (1975) Creep, Visoelasticity and Creep Fracture in Solids, Applied
Science Publishers, London.
Gittus, J. (1982) In: Mechanics of Solids, The Rodney Hill 60th Anniversary
Volume, H.G. Hopkins and M.J. Sewell, Eds., Pergamon Press, Oxford, pp.
227-263.
Gleiter, H. and Chalmers, B. (1972) Progr. Mater. Sci. ~, 77-112.
Gooch, D.J. (1981) Metal Sci. 15, 45-54.
Gooch, D.J. (1982) Metal Sci. 16, 79-89.
Goods, S.H. and Brown, L.M. (1979) Acta Metall. 27, 1-15.
Goods, S.H. and Nix, W.D. (1978) Acta Metall. 26:-739-752.
Grabke, H.J. and Martin, E. (1973) Arch. Eisenhtittenwes. 44, 837-842.
Grabke, H.J., Paulitschke, W., Tauber, G. and Viefhaus, H~(1977) Surface Sci.
63, 337-389.
Grant:-N.J. (1971) In: Fracture, An Advanced Treatise, vol. 3, H. Liebowitz,
Ed., Academic Press, New York and London, pp. 483-533.
GreenWOOd, J.N. (1952) J. Iron Steel Inst. 171, 380.
Greenwood, G.W. (1978) Phil. Mag. 19, 423-4~
GreenWOOd, G.W. (1978) Phil. Trans~Roy. Soc. Lond. A288, 213-227.
Greenwood, J.N., Miller, D.R. and Suiter, J.W. (1954)~ta Metall. 2, 250-258.
Gurson, A.L. (1977) Trans. ASME, J. Engrg. Mater. Technology 99, 2-15.
Guttmann, M. (1975) Surface Sci. 53, 213-227.
-Guttmann, M. (1980) In: Residuals:-Additives and Materials Properties, Phil.
Trans. R. Soc. Lond. A295, University Press, Cambridge, pp. 151-164.
Guttmann, M. and McLean, D:-11979) in Interfacial Segregation, W.C. Johnson and
J.M. Blakely, Eds., American Society for Metals, Metals Park, Ohio, pp.
261-348.
Hales, R. (1980) Fatigue Eng. Mater. Structures 1, 339-356.

406

References

Hancock, J.W. (1976) Metal Sci. 10, 319-325.


Hanna, M.D. and Greenwood, G.W. (1982) Acta Metall. 30, 719-724.
Hansen, M. (1958) Constitution of Binary Alloys, McGraw-Hill, New York.
Harris, J.E. (1965) Trans. AIME 233, 1509-1516.
Harris, J.E. (1976) J. Nuclear Mater. 59, 303-306.
Harris, J.E., Tucker, M.O. and Greenwood, G.W. (1974) Metal Sci. 8, 311-314.
Harrison, C.B. and Sandor, G.N. (1971) Eng. Fracture Mech. 3, 403~420.
Hart, E.W. (1967) Acta Metall. 15, 1545-1549.
Hart, E.W. (1976) Trans. ASME, ~ Eng. Mater. Technology 98, 193-202.
Hart, E.W. (1980) Int. J. Solids Strutures 16, 807-823. -Hart, E.W. (1983) In: Elastic-Plastic Fracture: Second Symposium, Vol. I Inelastic Crack Analysis, ASTM STP 803, C.F. Shih and J.P. Gudas, Eds.,
American Society for Testing and Materials, pp. 1-521-1-531.
Hartweck, W.D. (1979), Ph.D. Thesis, University of Dortmund (in German).
Hawk, D.E. and Bassani, J.L. (1986) J. Mech. Phys. Solids 34, 191-212.
Hayhurst, D.R., Brown, P.R. and Morrison, C.J. (1984) Phil~Trans. Roy. Soc.
Lond. A ~, 131-158.
Hayhurst, D.R. and Leckie, F.A. (1984) In: Mechanical Behaviour of Materials,
Proceedings of ICM4, Vol. 2, J. Carlsson and N.G. Ohlson, Eds., Pergamon
Press, Oxford, pp. 1195-1212.
He, M.Y. (1983) In: International Symposium on Fracture Mechanics, Tan Deyan
and Chen Daning, Eds., Science Press, Beijing, China, pp. 31-41.
He, M.Y. and Hutchinson, J.W. (1981) J. Appl. Mech. 48, 830-840.
He, M.Y. and Hutchinson, J.W. (1983) In: Elastic-Plastic Fracture, Second
Symposium;-~Vol. 1, Inelastic Crack Analysis, ASTM STP 803, C.F. Shih and
J.P. Gudas, Eds., American Society for Testing and Materials, pp.
1-291-1-305.
Heald, P.T. and Williams, J.A. (1970) Phil. Mag. 22, 1095-1100.
Heald, P.T. and Williams, J.A. (1971) Phil. Mag. 24, 1215-1220.
Heitmann, H.H., Vehoff, H. and Neumann, P. (1984)-rn: Advances in Fracture
Research '84 - Proceedings of ICF6, Vol. 5, S.R. Valluri et aI, Eds.,
Pergamon Press, Oxford and New York, pp. 3599-3606.
Henderson, P.J. and McLean, M. (1984) In: Creep and Fracture of Engineering
Materials und Structures, B. Wilshire and D.R.J. Owen, Eds., Pineridge
Press, Swansea, Part I, pp. 319-332.
Henderson, P.J. and McLean, M. (1985) Scripta Met. 19, 99-104.
Herring, C. (1950) J. Appl. Phys. 21, 437-445.
-Herring, C. (1951)'In: The Physics-of Powder Metallurgy, W.E. Kingston, Ed.,
McGraw-Hill, New York, p. 143.
Hill, R. (1956a) Plasticity, Oxford at the Clarendon Press, Oxford.
Hill, R. (1956b) J. Mech. Phys. Solids 5, 66-74.
Hillert, M. (1957) Jernkont. Annaler 141, 67-89.
Hillert, M. (1976) 'In: Physical Chemistry in Metallurgy, R.M. Fisher, R.A.
Oriani and E.T. Turkdogan, Eds., United States Steel Corporation,
Monroeville, pp. 445-462.
Hippsley, C.A., Knott, J.F. and Edwards, B.C. (1980) Acta Metall. 28, 869-885.
Hippsley, C.A., Knott, J.F. and Edwards, B.C. (1982) Acta Metall. 30, 641-654.
Hippsley, C.A., Rauh, H. and Bullough, R. (1984) Acta Metall. 32, 1381-1394.
Hirth, J.P. and Rice, J.R. (1980) Metall. Trans. 11A, 1501-151~
Hirth, J.P. and NiX, W.D. (1985) Acta Metall. 33,~9-368.
Hoenig, A. (1978) Int. J. Solids Structures 14~925-934.
Hoff, N.J. (1953) J. Appl. Mech. 20, 105-108~
Hoff, N.J. (1954) Quart. Appl. Math. 12, 49-55.
Hollstein, T. (1985) z. Werkstofftech~16, 223-228.
Hollstein, T., Blauel, J.G. and Voss, B~(1984) In: Elastic-Plastic Fracture
Test Methods: The User's Experience, ASTM STP 856, E.T. Wessel and F.J.
Loss, Eds., American Society for Testing and Materials, pp. 104-116.
Holt, R.T. and Wallace, W. (1976) Int. Metals. Rev. ~, 1-24.

References

407

Hondros, E.D. (1965) Proc. Roy. Soc. Lond. A206, 479-498.


Hopkins, B.E., Tipler, H.R. and Branch, G.D.-c1971) J. Iron Steel Inst. 209,
745-746.
Horii, H. and Nemat-Nasser, S. (1983) J. Mech. Phys. Solids 31, 155-171.
Hsueh, C.H. and Evans, A.G. (1981) Acta Metall. 29, 1907-191~
Hui, C.Y. (1983) In: Elastic-Plastic Fracture: Second Symposium, Vol. I Inelastic Crack Analysis, ASTM STP 803, C.F. Shih and J.P. Gudas, Eds.,
American Society for Testing and Materials, pp. 1-573-1-593.
Hui, C.Y. (1986) Int. J. Solids Structures 22, 357-372.
Hui, C.Y. and Banthia, V. (1984) Int. J. Fracture 25, 53-67.
Hui, C.Y. and Riedel, H. (1981) Int. J. Fracture 1~ 409-425.
Hui, C.Y. and Wu, K.-C. (1986) Int. J. Fracture 31: 3-16.
Hull, D. and Rimmer, D.E. (1959) Phil. Mag. 4, 673-687.
Hult, J. (1962) In: Proceedings of the International Symposium on Second Order
Effects in Elasticity, Plasticity and Fluid Mechanics, (Haifa, April
1962), M. Reiner and D. Abin, Eds., Pergamon Press, Oxford, and Jerusalem
Academic Press, p. 352.
Hult, J.D. and McClintock, F.A. (1957) In: 9th International Congress on
Applied Mechanics, Vol. 8, Brussels, pp. 51-58.
Hutchinson, J.W. (1968a) J. Mech. Phys. Solids 16, 13-31.
Hutchinson, J.W. (1968b) J. Mech. Phys. Solids 16, 337-347.
Hutchinson, J.W. (1983) Acta Metall. 31, 1079-1088.
Hutchinson, J.W. and Obrecht, H. (1977) In: Fracture 1977, D.M.R. Taplin, Ed.,
University of Waterloo Press, Waterloo, Canada, vol. 1. pp. 101-113.
Hutchinson, J.W. and Paris, P.C. (1979) In: Elastic-Plastic Fracture, ASTM STP
668, American Society for Testing and Materials, pp. 37-64.
Ilschner, B. (1973) Hochtemperaturplastizitat, Springer Verlag, Berlin,
Heidelberg, New York.
Ilyushin, A.A. (1946) Prikladnaia Mat. Mekh. P.M.M. 10, 347.
Intrater, J. and Machlin, E.S. (1959) Acta Metall. 7:-140-143.
Intrater, J. and Machlin, E.S. (1959/60) J. Inst. Metals 88, 305-310.
Ishida, Y. and McLean, D. (1967) Metal Sci. J. 1, 171-172:James, L.A. (1979) Trans. ASME, J. Pressure Vessel Technology 101, 171-176.
Jager, W., Grabke, H.J. and Jin Yu (1984) In: Creep and Fractur-e-of Engineering
Materials, D.R.J. OWen and B. Wilshire, Eds., Pineridge Press, Swansea,
Part II, pp. 649-660.
Jenkins, C.H.M., Bucknall, E.H. and Jenkinson, E.A. (1943) J. Inst. Metals 70,
57-79.
Johannesson, T. and Tholen, A. (1969) J. Inst. Met. 97, 243-247.
Kachanov, L.M. (1960) The Theory of Creep. English translation edited by A.J.
Kennedy, Boston Spa, Wetherby.
Kelly, A. (1966) Strong Solids, Clarendon Press, Oxford.
Kettunen, P.O., Lepisto, T., Kostorz, G. and Goltz, G. (1981) Acta Metall. 29,
969-972.
Khare, A.K. (1983) (Editor) Ferritic Steels for High-Temperature Applications,
American Society for Metals, Metals Park, Ohio.
Kikuchi, M. and Miyamoto, H. (1984a) In: Proceedings of the Third International
Conference on Numerical Methods in Fracture Mechanics, A.R. Luxmoore and
D.R.J. Owen, Eds., Pineridge Press, Swansea, U.K., pp. 205-217.
KikUChi, M. and Miyamoto, H. (1984b) Int. J. Pres. Ves. Piping 16, 1-16.
King, B.L. (1980) in ReSiduals, Additives and Materials Properties, Phil.
Trans. R. Soc. Lond. A295, University Press, Cambridge, pp. 151-164.
Knauf, G. and Riedel, H. (1981) In: Advances in Fracture Research, Vol. 5, D.
Francois et aI, Eds., Pergamon Press, Oxford and New York, pp. 2547-2553.
Knott, J.F. (1973) Fundamentals of Fracture Mechanics, Butterworths, London.
Koeller, R.C. and Raj, R. (1978) Acta Metall. 26, 1551-1558.
Koterazawa, R. and Mori, T. (1977) Trans. ASME:-J. Eng. Mater. Technology 99,
298-305.
-

408

References

Krishnamohanrao, Y., Kutumbarao, V.V. and Rama Rao, P. (1986) Acta Metall. 34,
1783-1806.
Kubashewski, O. and Alcock, C.B. (1979) Metallurgical Thermochemistry, 5th
Edition, Pergamon Press, Oxford.
Kubo, S. (1981a) Combined Effect of Creep Recovery-Hardening and Elastic
Straining on the Stress and Strain-Rate Fields near a Crack Tip in
Creeping Materials, Brown University Report MRL E-133, Providence, RI.
Kubo, S. (1981b) In: Creep in Structures, A.R.S. Ponter and D.R. Hayhurst,
Eds., Springer-Verlag, Berlin Heidelberg New York, pp. 606-610.
Kubo, S. (1983) In: Elastic-Plastic Fracture: Second Symposium, Volume I Inelastic Crack Analysis. ASTM STP 803, C.F. Shih and J.P. Gudas, Eds.,
American Society for Testing and Materials, pp. 1-594-1-614.
Kubo, S., Ohji, K. and Ogura, K. (1979) Eng. Fracture Mech. 11, 315-329.
Kumar, V., German, M.D. and Shih, C.F. (1981) An Engineering~pproach for
Elastic-Plastic Fracture AnalYSiS, Report NP-1931 on Project 1237-1 for
Electric Power Research Institute, Palo Alto, California.
Lagneborg, R. (1981) In: Creep and Fatigue in High Temperature Alloys, J.
Bressers, Ed., Applied Science Publishers, London, pp. 41-72.
Lai, J.K. and Wickens, A. (1979) Acta Metall. 27, 217-230.
Laird, C. and Smith, G.C. (1962) Phil. Mag. 7,-S47-857.
Landau, L.D. and Lifshitz, E.M. (1966) ElastTzitatstheorie, Course on
Theoretical Physics, vol. 7, Akademie-Verlag, Berlin. English translation
available from Pergamon Press, Oxford.
Landes, J.D. and Begley, J.A. (1976) In: Mechanics of Crack Growth, ASTM STP
590, Amerinan Society for Testing and Materials, pp. 128-148.
Landolt-Bornstein (1961) Zahlenwerte und Funktionen, 6. Ausgabe, II. Band, 4.
Teil, Springer-Verlag, Berlin Gottingen Heidelberg.
Langdon, T.G. (1981) In: Deformation of Polycrystals: Mechanisms and
Microstructures, N. Hansen, A. Horsewell, T. Leffers and H. Lilholt, Eds.,
Riso National Laboratory, Roskilde, Denmark, pp. 45-54.
Langdon, T.G. (1982) Metal Sci. 16, 175-183.
Langdon, T.G. and Vastava, R.B. (1982) In: Mechanical Testing for Deformation
Model Development, ASTM STP 765, R.W. Rohde and J.C. Swearengen, Eds.,
American Society for Testing and Materials, pp. 435-451.
Larson, F.R. and Miller, J. (1952) Trans. ASME 74, 765-775.
Larsson, S.G. and Carlsson, A.J. (1973) J. Mech~Phys. Solids 21, 263-277.
Lau, C.W. and Argon, A.S. (1977) In: Fracture 1977, ProceedingS-of ICF4, vol.
2, D.M.R. Taplin, Ed., Waterloo University Press, Waterloo, Ontario, pp.
595-601
Lau, C.W., Argon, A.S. and McClintock, F.A. (1983) In: Elastic-Plastic
Fracture: Second Symposium, Vol. I-Inelastic Crack AnalysiS, ASTM STP 803,
C.F. Shih and J.P. Gudas, Eds., American Society for Testing and
Materials, pp; 1-551-1-572.
Law, C.C. and Blackburn, M.J. (1980) Metall. Trans. 11A, 495-507.
Leeuwen, H.P. van (1977) Eng. Fracture Mech. 9, 951-974.
Leipholz, H. (1968) Einfilhrung in die ElastizTtatstheorie, G. Braun, Karlsruhe.
Lewis, M.H. and Karunaratne, B.S.B. (1981) In: Fracture Mechanics Methods for
Ceramics, Rocks and Concrete, ASTM STP 745, S.W. Freiman and E.R. Fuller,
Eds., American Society for testing and Materials, pp. 13-32.
Lim, L.C. and Raj, ,R. (1984a) Acta Metall. 32,1183-1190.
Lim, L.C. and Raj, R. (;984b) Acta Metall. 32, 727-733.
Liu, C.T., White, C.L., Koch, C.C. and Lee,~.H. (1983) In: High Temperature
Materials Chemistry-II, Z.A. Munir and D. Cubicciotti, Eds., The
Electrochemical Society, Pennington, New Jersey, pp. 32-41.
Liu, C.T., White, C.L. and Lee, E.H. (1985) Scripta Met. 19, 1247-1250.
Lloyd, G.J. (1983) In: Fatigue at High Temperature, R.P. Skelton, Ed., Applied
Science Publishers, London and New York, pp. 187-258.
Lloyd, G.J. and Wareing, J. (1981) Metals Technology, ~, 297-305.

References

409

Lonsdale, D. and Flewitt, P.E.J. (1979) Mater. Sci. Eng. 39, 217-229.
Lonsdale, D. and Flewitt, P.E.J. (1981) Proc. Roy. Soc. Lond. A373, 491-509.
Lopez, H. and Shewmon, P.G. (1983) Acta Metall. ll, 1945-1950.---Love, A.E.H. (1952) The Mathematical Theory of Elasticity, Cambridge University
Press, Cambridge.
Loveday, M.A. and Dyson, B.F. (1983) Acta Metall. 31, 397-405.
Luxmoore, A.R. and Owen, D.R.J., Eds. (1984) Proceedings of the Third
International Conference on Numerical Methods in Fracture Mechanics,
Pineridge Press, Swansea, U.K.
Machlin, E.S. (1956) Trans. AIME 206, J. Metals 8, 106-111.
Maiya, P.S. (1981) Mater. Sci. Eng:-47 , 13-21. Majumdar, S. and Maiya, P.S. (1979) Canad. Metall. Qu. 18, 57-64.
Mancuso, J.F. and Li, C.-Y. (1979) Metall. Trans. 10A, 1719-1722.
McClintock, F.A. (1968) J. Appl. Mech. 35, 363-371--.-McClintock, F.A. and Bassani, J.L. (1981) In: Three-Dimensional Constitutive
Relationships and Ductile Fracture, J. Zarka and S. Nemat-Nasser, Eds.,
North Holland Publishing Co., pp. 119-141.
McClintock, F.A. and Irwin, G.R. (1965) In: Fracture Thoughness Testing and Its
Applications, ASTM STP 381, American Society for Testing and Materials,
pp. 84-113.
McKimpson, M. and Shewmon, P.G. (1981) Metall. Trans. 12A, 825-834.
McLaughlin, R. (1977) Int. J. Engng. Sci. 15, 237-244.
McLean, D. (1956/7) J. Inst. Metals 85, 468-472.
McLean, D. Grain Boundaries in Metal~ Oxford University Press, London, 1957.
McLean, D. (1958) In: Vacancies and Other Points Defects in Metals and Alloys,
Institute of Metals Monograph and Report Series No. 23, London.
McLean, D. (1963) J.Austral. Inst. Metals 8, 45-51.
McMahon, C.J. (1968) Temper Embrittlement in Steel, ASTM STP 407, 127-167.
McMahon, C.J. (1984) Z. Metallkde. 75, 496-509.
McMahon, C.J., Weng Yu-quing, Morga~ M.J. and Meynhard, M. (1985) In: Fourth
JIM International Symposium: Grain Boundary Structure and Related
Phenomena, Minakami Spa, Japan.
McMeeking, R.M. (1977) J. Mech. Phys. Solids 25, 357-381.
McMeeking, R.M. and Parks, D. (1979) In: Elastic-Plastic Fracture, ASTM STP
668, J.D. Landes, J.A. Begley and G.A. Clarke, Eds., American Society for
Testing and Materials, pp. 175-194.
Melford, D.A. (1980) In: Residuals, Additives and Materials Properties, Phil.
Trans. R. Soc. Lond. A295, 89-103.
Merkle, J.G. and Corten, H:r:-(1974) J. Pressure Vessel Technology, Trans. ASME
96, 286-292.
Miche~ D.J. and Smith, H.H. (1980) Acta Metall 28, 999-1007.
Middleton, C.J. (1981) Metal Sci. 15, 154-167. -Miller, D.A. (1979) Sc'ripta Meta 11-:-13 , 595-598.
Miller, D.A. and Pilkington, R. (1978) Metall. Trans. 9A, 1221-1227.
Miller, K.J. (1985) In: Fundamentals of Deformation and Fracture, B.A. Bilby,
K.J. Miller and J.R. Willis, Eds., Cambridge University Press, Cambridge,
pp. 477-500.
Mills, W.J. and James, L.A. (1980) Fatigue Eng. Mater. Structures 3, 159-175.
Min, B.K. and Raj, R. (1978) Acta Metall. 26, 1007-1022.
Min, B.K. and Raj, R. '(1979a) Canad. Metal!." Qu. 18, 171-176.
Min, B.K. and Raj, R. (1979b) In: Fatigue Mechanisms, J.T. Fong, Ed., ASTM STP
675, American SOCiety for Testing and Materials, pp. 569-591.
Moller, R., Erhart, H. and Grabke, H.J. (1984) Arch. Eisenhilttenwes. 55,
543-548.
-Monkman, F.C. and Grant, N.J. (1956) Proc. American Society for Testing and
Materials 56, pp. 593-620.
Mullendore, A.W-:-and Grant, N.J. (1961), In: Structural Processes in Creep.
Special Report No. 70, Iron and Steel Institute, London.

410

References

Mura, T. (1982) Micromechanics of Defects in Solids, Martinus Nijhoff


Publishers, The Hague, Boston.
Muskhelishvili, N.I. (1977) Mathematical Theory of Elasticity, Noordhoff
Internationl Publishing, Leyden.
Nabarro, F.R.N. (1948) In: Proc. Conference on Strength of Solids, Physical
SOCiety, London, p. 75.
Nazmy, M.Y. and Duerig, T.W. (1982) Scripta Metall. 16, 209-212.
Needham, N.G. (1983) Report of British Steel Corporation to the Commission of
European Communities, Contract No. 7210.MA/802.
Needham, N.G. and Gladman, T. (1980) Metal Sci. 14, 64-72.
Needham, N.G. and Gladman, T. (1982) In: AdvanceS-in the Physical Metallurgy
and Applications of Steels, Book 284, The Metals Society, London, pp.
309-317.
Needham, N.G. and Gladman, T. (1984) In: Creep and Fracture of Engineering
Materials and Structures, D. Wilshire and D.R.J. Owen, Eds., Pineridge
Press, Swansea, Part II, pp. 1263-1276.
Needham, N.G. and GreenWOOd, G.W. (1975) Metal Sci. 9, 258-262.
Needham, N.G. and Orr, J. (1980) in Residuals, AdditIves and Materials
Properties, Phil. Trans. R. Soc. Lond. A295, University Press, Cambridge,
pp. 151-164.
Needleman, A. and Rice J.R. (1980) Acta Metall. 28, 1315-1332.
Neubauer, B. (1981) In: Recent Advances in Creep and Fracture of Engineering
Materials and Structures, B. Wilshire and D.R.J. Owen, Eds., Pineridge
Press, Swansea, pp. 617-619.
Neubauer, B. and Arens-Fischer, F. (1982) In: Kraftwerke 1982, Vortrag C10 beim
VGB-Kongre1 in Mannheim, Sept 1982, Vereinigung der
GroBkraftwerksbetreiber, Essen.
Neumann, P. (1974) Acta Metall. 22, 1155-1165, and 1167-1178.
Neumann, P., Fuhlrott, H. and Vehoff, H. (1979) In: Fatigue Mechanisms, ASTM
STP 675, J.T. Fong, Ed., American Society for Testing and Materials, pp.
371-395.
Newman, J.C. (1981): In: Methods and Models for Predicting Fatigue Crack Growth
under Random Loading, ASTM STP 748, J.B. Chang and C.M. Hudson, Eds.,
American Society for Testing and Mateials, pp. 53-84.
Nieh, T.G. and Nix, W.D. (1979) Acta Metall. 27, 1097-1106.
Nieh, T.G. and Nix; W.D. (1980a) Acta Metall.:28, 557-566.
Nieh, T.G. and Nix, W.D. (1980b) Scripta Metal~ ;4, 365-368.
Nieh, T.G. and Nix, W.D. (1981) Metall. Trans. 12~ 893-901.
Nikbin, K.M., Webster, G.A. and Turner, C.E. (1976) In: Cracks and Fracture,
ASTM STP 601, American Society for Testing and Materials, pp. 47-62.
Nilsson, G.G. and Roth, M. (1981) Mater. Sci. Eng. 50, 101-108.
Nix, W.D., Yu, K.S. and Wang, J.S. (1983) Metall. Trans. 14A, 563-570.
Norton, F.H. (1929) Creep of Steel at High Temperatures, McGraw-Hill, New York.
Oden, J.T. (1972) Finite Elements in Nonlinear Continua, McGraw-Hill, New York.
Ohji, K. (1977) Theoretical and Applied MechaniCS, 27, 3-20.
Ohji, K., Ogura, K. and Kubo, S. (1974) Preprint of-Yapan. Soc. Mech. Engrs.
No. 640-11, p. 207 (in Japanese).
Ohji, K., Ogura, K. and Kubo, S. (1980) J. Soc. Mater. Sci. Japan 29, No. 320,
pp. 465-471.
Ohji, K., Ogura, K." Kubo, S. and Katada, Y. (1980) Proc. of the International
Conference on Engineering Aspects of Creep, Sheffield, Vol. 2, The
Institution of Mechanical Engineers, London, pp. 9-16.
Okazaki, M. and KOizumi, T. (1983) Trans. ASME, J. Eng. Mater. Technology ~,
81-87.
Orowan, E. (1946) J. West Scotl. Iron Steel Inst. 54, 45-53.
Orowan, E. (1948) Rep. Progress Phys., The Physicar-Society, London, VOl. 12,
pp. 185-232.
Orr, R.L., Sherby, O.D. and Dorn, J.E. (1954) Trans. ASME 76, 113.

References

411

Owen, D.R.J. and Hinton, E. (1980) Finite Elements in Plasticity: Theory and
Practice, Pineridge Press, Swansea.
Owen, D.R.J. and Luxmoore, A.R. (Eds.) (1980) Numerical Methods in Fracture
Mechanics, Pineridge Press, Swansea.
Padmanabhan, K.A. and Davies, G.J. (1980) Superplasticity, Springer-Verlag,
Berlin Heidelberg New York.
Page, R., Weertman, J.R., and Roth M. (1982) Acta Metall. 30, 1357-1366.
Pan, J., Saje, M. and Needleman, A. (1983) Int. J. Fracture-21, 261-278.
Pandey, M.C., Dyson, B.F. and Taplin, D.M.R. (1984) Proc. RoY: Soc. Lond. A
393,1437-1441.
PandeY:-M.C., Mukherjee, A.K. and Taplin, D.M.R. (1984) Metall. Trans. l2!,
1437-1441.
Pandey, M.C., Taplin, D.M.R. and Mukherjee, A.K. (1984) Metall. Trans. 15A,
1763-1767.
Parker, C.A. and Russell, K.C. (1981) Scripta Metall. 15, 643-647.
Parker, J.D. and Wilshire, B. (1980) Mater. Sci. Engin~43, 271-280.
Parthasarathy, T.A., Lopez, H.F. and Shewmon, P.G. (1985~Metall. Trans. ~,
1143-1149.
Pavinich, W. and Raj, R. (1977) Metall. Trans. 8A 1917-1933.
Pelloux, R.M. and Huang, J.S. (1980) In: Creep-Fatigue-Environment
Interactions, R.M. Pelloux and N.S. Stoloff, Eds., The Metallurgical
Society of AIME, pp. 151-164.
Perry, A.J. (1974) J. Mater. Sci. 9, 1016-1039.
Peterseim, J. and Sauthoff, G. (1984) In: Creep and Fracture of Engineering
Materials und Structures, B. Wilshire and D.R.J. Owen, Eds., Pineridge
Press, Swansea, Part I, pp. 267-277.
Peterson, N.L. (1980) In: Grain-Boundary Structure and Kinetics, ASM Materials
Science Seminar, 1979, American Society for Metals, Metals Park, Ohio, pp.
209-238.
Pilkington, R. (1979) Met. Sci. 13, 555-564.
Pilkington, R., Miller, D.A. and-Worswick, D. (1981) Metall. Trans. ~,
173-181.
Pineau, A. (1985) In: Flow and Fracture at Elevated Temperatures, R. Raj, Ed.,
American Society for Metals, Metals Park, Ohio, pp. 317-348.
Pizzo, P.P. and Mandurrago, G.L. (1981) J. Eng. Mater. Technology 103, 62-70.
Piayer, R.L. and Brinson, G. (1975) J. Austral. Inst. Metals 20, 226-238.
Pope, D.P. (1983) In: Embrittlement of Engineering Alloys, C.~ Briant and S.N.
Banerji, Eds., Academic Press, New York, pp. 125-155. .
Pope, D.P. and Wilkinson, D.S. (1981) In: Recent Advances in Creep and Fracture
of Engineering Materials and Structures, B. Wilshire and D.R.J. Owen,
Eds., Pineridge Press, Swansea, pp. 531-544.
Porter, J.R., Blumenthal, W. and Evans, A.G. (1981), Acta Metall. 29,
1899-1906.
Prager, W. and Hodge, P.G. (1951) Theory of Perfectly Plastic Solids, John
Wiley, New York.
Presland, A.E.B. and Hutchinson, R.I. (1963/4) J. Inst. Metals 92, 264-269.
Price, C.E. (1966) Acta Metall. 14, 1781-1799.
Pugh, C.E. and Robinson, D.N. (1978) Nucl. Eng. Design 48, 269-276.
Rabotnov, Y.N. (1969) Creep Problems in Structural Members, translation from
Russian original (1966) by F.A. Leckie, North-Holland Publ., Amsterdam.
Raj, R. (1975) Metall. Trans. 6A, 1499-1509.
Raj, R. (1976) Trans. ASME, J.~ng. Mater. Technology 98, 132-139.
Raj, R. (1978a) Acta Metall. 26, 341-349.
Raj, R. (1978b) Acta Metall. 26, 995-1006.
Raj, R. (1982) Acta Metall. 3~ 1259-1268.
Raj, R. (1983) Acta Metall. 31, 29-36.
Raj, R. and Ashby, M.F. (1971) Metall. Trans. 2A, 1113-1127.
Raj, R. and Ashby, M.F. (1975) Acta Metall. 23:-653-666.

412

References

Raj, R., Shih, H.M. and Johnson, H.H. (1977) Scripta Met. 11, 839-842.
Rao, P.R, Rao, V.V.P.K. and Pandey, M.C. (1973) J. Iron Steel Inst. ~,
801-802.
Reiley, T.C. (1981) Scripta Metall. 15, 497-500.
Resnick, R. and Seigle, L. (1957) Trans. AIME 209, J. Metals 2, 87-94.
Rice, J.R. (1967a) J. Appl. Mech. 34, 287-298.--Rice, J.R. (1967b) In: Fatigue Crack Propagation, ASTM STP 415, American
Society for Testing and Materials, pp. 247-311.
Rice, J.R. (1968a) In: Fracture. An Advanced Treatise, H. Liebowitz, Ed., vol.
2, Academic Press, New York, pp. 191-311.
Rice, J.R. (1968b) ASME J. Appl.Mech. 35, 379-386.
Rice, J.R. (1974) J. Mech. Phys. SolidS-22, 17-26.
Rice, J.R. (1976) In: Effect of Hydrogenon Behavior of Materials, A.W.
Thompson and I.M. Bernstein, Eds., Metallurgical Society of AIME, pp.
455-466.
Rice, J.R. (1981) Acta Metall. 29, 675-681.
Rice, J.R., Drugan, W.J. and Sham, T.-L. (1980) In: Fracture Mechanics: Twelfth
Conference, ASTM STP 700, P.C. Paris, Ed., American Society for Testing
and Materials, pp. 189-221.
Rice, J.R. and Johnson, M.A. (1970) In: Inelastic Behavior of Solids, M.F.
Kanninen et al., Eds., McGraw-Hill, New York, pp. 641-671.
Rice, J.R. and Rosengren, G.F. (1968) J. Mech. Phys. Solids 16, 1-12.
Rice, J.R. and Thomson, R. (1974) Phil. Mag. 29, 73-97.
-Rice, J.R. and Tracey, D.M. (1969) J. Mech. Phys. Solids 17, 201-217.
Riedel, H. (1976) In: Fourth International Conference on the Strength of Metals
and Alloys, Laboratoire de Physique du Solide, Nancy, pp. 529-534.
Riedel, H. (1978) Z. Metallkunde 69, 755-760.
Riedel, H. (1981a) In: Creep in Structures, A.R.S. Ponter and D.R. Hayhurst,
Eds., Springer-Verlag, Berlin Heidelberg, pp. 504-519.
Riedel, H. (1981b) J. Mech. Phys. Solids 29, 35-49.
Riedel, H. (1983a) In: Elastic-Plastic Fracture, Second Symposium, Vol. 1,
Inelastic Crack Analysis, ASTM STP 803, C.F. Shih and J.P. Gudas, Eds.,
American Society for Testing and Materials, pp. 1-505-1-520.
Riedel, H. (1983b) In: Mechanical Behaviour of Materials-IV, Vol. 1, J.
Carlsson and M.G. Ohlson, Eds., Pergamon Press, Oxford, pp. 67-78.
Riedel. H. (1984a) In: Subcritical Crack Growth Due to Fatigue, Stress
Corrosion and Creep, L.H. Larsson, Ed., Elsevier Applied Science
Publishers, Barking, England, pp. 449-467.
Riedel, H. (1984b) Acta Metall. 32, 313-321.
Riedel, H. (1985a) In: Flow and Fracture at Elevated Temperatures, R. Raj, Ed.,
American Society for Metals, Metals Park, OhiO, pp. 149-177.
Riedel, H. (1985b) In: Fundamentals of Deformation and Fracture, B.A. Bilby,
K.J. Miller artd J.R. Willis, Eds., Cambridge University Press, Cambridge,
pp. 293-309.
Riedel, H. (1985c) Z. Metallkunde 76, 669-675.
Riedel, H. and Kochendorfer, A. (1979) Arch. Eisenhilttenwes. 50, 173-178.
Riedel, H. and Rice, J.R. (1980) In: Fracture Mechanics: Twelfth Conference,
ASTM STP 700, P.C. Paris, Ed., American Society for Testing and Materials,
pp. 112-1 30.
Riedel, H. and Wagner, W. (1981) In: Advances in Fracture Research, Proceedings
of ICF5. vol. 2, D. Francois et aI, Eds., Pergamon Press, Oxford and New
York, pp. 683-688.
Riedel, H. and Wagner, W. (1985) In: Advances in Fracture Research '84 - Proceedings of ICF6, S.R. Valluri et aI, Eds., Pergamon Press, Oxford, Vol.
3, pp. 2199-2206.
Robinson, D.N. (1978) Oak Ridge National Laboratory Report No. TM-5969, Oak
Ridge, Tenn.
Robinson, J.R. and Scott, M.H. (1980) In: Residuals, Additives and Materials

References

413

Properties, Phil. Trans. R. Soc. Lond. A295, 105-117.


Robson, K. (1972) International Conference on Creep Resistance in Steel, Verein
Deutscher Eisenhuttenleute, Dusseldorf.
Rooke, D.P. and Cartwright, D.J. (1974) Stress Intensity Factors, Her Majesty's
Stationery Office, London.
Russell, K.C. (1970) In: Phase Transformations, ASM-Seminar, 1968, American
Society for Metals, Metals Park, Ohio, pp. 218-268.
Russell, K.C. (1971) Acta Metall. 19, 753-758.
Russell, K.C. (1978) Acta Metall. 26, 1615-1630.
Sadananda, K. and Shahinian, P. (1980a) Mater. Sci. Eng. 43, 159-168.
Sadananda, K. and Shahinian, P. (1980b) Metall. Trans. 11~ 267-276.
Sadananda, K. and Shahinian, P. (1981a) Eng. Fracture Mech. l2, 327~342.
Sadananda, K. and Shahinian, P. (1981b) In: Cavities and Cracks in Creep and
Fatigue, J. Gittus, Ed., Applied Science Publishers, London and New
Jersey, pp. 109-195.
Sadananda, K. and Shahinian, P. (1983) Metall. Trans. 14A, 1467-1480.
Sadananda, K. and Shahinian, P. (1984) Metall. Trans. 15A, 527-538.
Saegusa, T. and Weertman, J.R. (1978) Scripta Metall. 12; 187-191.
Saegusa, T., Weertman, J.R., Cohen, J.B. and Roth, M. (1978) J. Appl.
Crystallogr. 11, 602.
Saxena, A. (1980a)-rn: Fracture Mechanics: The Twelfth Conference, ASTM STP
700, P.C. Paris, Ed., American Society for Testing and Materials, pp.
131-151.
Saxena, A. (1980b) Fatigue Eng. Mater. Structures 3, 247-256.
Saxena, A. (1983) In: Thermal and Environmental Effects in Fatigue:
Research-Design Interface, PVP-Vol. 71, C.E. Jaske, S.J. Hudak and M.E.
Mayfield, Eds., The American Society of Mechanical Engineers, New York,
pp. 171-184.
Saxena, A. (1986) In: Fracture Mechanics: Seventeenth Symposium, ASTM STP905,
J.H. Under ynnrl et al, Eds., American Society for Testing and Materials,
pp. 185-201
Saxena, A. and Bassani, J.L. (1984) In: Fracture: Interactions of
Microstructure, Mechanisms and Mechanics, J.M. Wells and J.D. Landes,
Eds., The Metallurgical Society of AIME, pp. 357-383.
Saxena, A., Ernst, H.A. and Landes, J.P. (1983) Int. J. Fracture 23, 245-258.
s~xena, A. and Hudak, S.J. (1978) Int. J. Fracture 14, 453-468. -Saxena, A. and Landes" J.D. (1984) In: Advances in Fracture Research '84 - Proceedings of ICF6, Vol.6, P. Rama Rao et al, Eds., Pergamon Press, Oxford,
pp. 3977-3988.
Saxena, A., Williams, R.S. and Shih, T.T. (1981) In: Fracture Mechanics,
Thirteenth Conference, ASTM STP 743, R. Roberts, Ed., American Society for
Testing and Materials, pp. 86-99.
Schmidt, C.G. and Milier, A.K. (1981) Res. Mechanica 2, 109-129 and 175-193.
Schneibel, J.H., White, C.L. and Padgett, R.A. (1983)-In: Strength of Metals
and Alloys (ICSMA6), vol. 2, R.C. Gifkins, Ed., Pergamon Press, Oxford and
New York, pp. 649-654.
Seah, M.P. (1975) Surface Sci. 53, 168-212.
Seah, M.P. (1980) Acta Metall. 28, 955-962.
Seah, M.P. and Hondros, E.D. (1973) Proc. Roy. Soc. Lond. A335, 191-212.
Seah, M.P. and Lea, C: (1975) Phil. Mag. 31, 627-645.
Segerlind, L.J. (1976) Applied Finite Element Analysis, John Wiley, New York.
Shahinian, P. and Sadananda, K. (1982) In: Fatigue Thresholds, Vol. 1, J.
Backlund, A. Blom and C.J. Beevers, Eds., Engineering Materials Advisory
Services, Warley, England, pp. 527-545.
Shahinian, P. and Sadananda, K. (1984) In: Superalloys 1984, M. Gell et al,
Eds., The Metallurgical Society of AIME, pp. 741-750.
Sham, T.-L. and Needleman, A. (1983) Acta Metall. 31, 919-926.
Shewmon, P.G. (1976) Metall. Trans. 7A, 279-286. --

414

References

Shewmon, P.G. (1985) Mater. Sci. Technology 1, 2-11.


Shih, C.F. (1974) In: Fracture Analysis, ASTM STP 560, American Society for
Testing and Materials, pp. 187-210.
Shih, C.F. (1983) Tables of the Hutchinson-Rice-Rosengren Singular Field
Quantities, Brown University Report MRL E-147, Providence, RI.
Shih, C.F., deLorenzi, H.G. and Andrews, W.R. (1977) Int. J. Fracture 13,
544-548.
-Shih, H.-M. and Johnson, H.H. (1982) Acta Metall. 30, 537-545.
Shin, J. and McMahon, C.J. (1984) Acta Metall. 32,-'535-1552.
Shiozawa, K. and Weertman, J.R. (1981) Scripta Metall. 15 1241-1244.
Shiozawa, K. and Weertman, J.R. (1983) Acta Metall. 31,-g93-1004.
Sidey, D. and Coffin, L.F. (1979) In: Fatigue Mechanisms, J.T. Fong, Ed., ASTM
STP 675, American Society for Testing and Materials, pp. 528-568.
Siverns, M.J. and Price, A.T. (1970) Nature 228, 760-761.
Skelton, R.P. (1966) Phil. Mag. 14, 563-572.--Skelton, R.P. (1975) Metal Sci. ~ 192-194.
Skelton, R.P. (1978a) Mater. Sci~ Eng. 32, 211-219.
Skelton, R.P. (1978b) Mater. Sci. Eng. 35, 287-298.
Skelton, R.P. (ed.) (1983) Fatigue at High Temperature, Applied Science
Publishers, London and New York.
Sklenicka, V., Saxl, I., Cadek, J. and Rys,P. (1980) Res Mech. 1, 301-317.
Sklenicka, V., Saxl, I. and Cadek, J. (1981a) Scripta Met. l2, 349-352.
Sklenicka, V., Saxl, I. and Cadek, J. (1981b) In: Advances in Fracture
Research, ICF5, vol. 4, D. Francois et aI, Eds., Pergamon Press, Oxford,
pp. 1643-1A50.
Smith, E. (1966) Acta Metall. 14, 985-989.
Smith, E. and Barnby, J.T. (1967) Metal Sci. J. 1, 1-4.
Smith, A.P., Towers, O.L. and Smith, I.J. (1984)-In: Proceedings of the Third
International Conference on Numerical Methods in Fracture Mechanics, A.R.
Luxmoore and D.R.J. Owen, Eds., Pineridge Press, Swansea, U.K., pp.
205-217.
Smith, E. (1966) In: Physical Basis of Yield and Fracture, Institute of Physics
and Physical Society Oxford, pp. 36-45.
Snowden, K.U. (1981) In: Cavities and Cracks in Creep and Fatigue, J. Gittus,
Ed., Applied Science Publishers, London, pp. 259-289.
Snowden, K.U., Hughes, D.S. and Stathers, P.A. (1981) Metal Sci. 15 73-78.
Speidel, M.D. (1981) In: Advances in Fracture Research (ICF5), Vol: 6, D.
Francois et aI, Eds., Pergamon Press, Oxford and New York, pp. 2685-2704.
Speight, M.V. and Beere, W. (1975) Metal Sci. 9, 190-191.
Speight, M.V., Beere, W. and Roberts, G. (1978) Mater. Sci. Eng. 36, 155-163.
Speight, M.V. and Harris, J.E. (1967) Metal Sci. J. 1, 83-85.
Sritharan, T. and Jones, H. (1980) Acta Metall. 28, 1633-1639.
Stanzl, S.E., Argon, A.S. and Tschegg, E.K. (1983) Acta Metall. 31, 833-843.
Stevens, R.A. and Flewitt, P.E.J. (1981) Acta Metall. 29, 867-88~
.
Stiegler, J.~. Farrell, K., Loh, B.T.M. and McCoy, H.E:-(1967) Trans. ASME 60,
494-503.
Stowell, M.J., Livesey, D.W. and Ridley, N. (1984) Acta Metall. 32, 35-42.
Stratmann, L., Keller, H., Hansel, H. and H.J. Grabke (1983) In:-Physico-Chimie
de l'Etat Solide, Applications aux Metaux et leur Composes, 36eme Reunion
Internationale de Chimie Physique, Paris.
Stroh, A.N. (1954) Proc. Roy. Soc. Lond. A 223, 404-414.
Stroh, A.N. (1957) Adv. Phys. 6, 418-465. --Stucke, M., Khobaib, M., Majumdar, B. and Nicholas, T. (1984) In: Advances in
Fracture Research '84 - Proceedings of ICF6, Vol. 6, S.R. Valluri et aI,
Eds., Pergamon Press, Oxford and New York, pp. 3967-3975.
Suresh, S., Parks, D.M. and Ritchie, R.O. (1982) In: Fatigue Thresholds, Vol.
1, J. Backlund, A. Blom and C.J. Beevers, Eds., Engineering Materials
Advisory Services, Warley, England, pp. 391-408.

References

415

Svensson, L.-E. and Dunlop, G.L. (1979) Canad. Metall. Qu. 18, 39-47.
Svensson, L.-E. and Dunlop, G.L. (1981) Int. Metals Rev. 26~109-131.
Svensson, L.-E. and Dunlop, G.L. (1982) Metal Sci. 16, 57-64.
Swaminathan, V.P., Shih, T.T. and Saxena, A. (1982):Eng. Fracture Mech. l,
827-836.
Swindeman, R.W., Farrell, K. and Yoo, M.H. (1981) Res Mechanica Letters ~,
67 -71
Swindeman, R.W., Sikka, V.K. and Klueh, R.L. (1983) Metall. Trans. ~,
581 -593.
Swinkels, F.B. and Ashby, M.F. (1981) Acta Metall. 29, 259-281.
Swinkels, F.B., Wilkinson, D.S., Arzt, E. and AshbY~M.F. (1983) Acta Metall.
31, 1829-1840.
Tada,~., Paris, P.C. and Irwin, G.R. (1973) Stress Analysis of Cracks
Handbook, Del Research Corporation, Hellertown, Pennsylvania.
Taira, S., Ohtani,.R. and Kitamura, T. (1979), Trans. ASME, J. Eng. Mater.
Technology 101, 154-161.
Taira, S., Ohtan~R. and Komatsu, T. (1979) Trans. ASME, J. Eng. Mater.
Technology 101, 162-167.
Takasugi, T. an d-rzumi, O. (1985) Acta Metall. 33, 39-48.
Takasugi, T. and Pope, D.P. (1982) Metall. Tran~ 13A, 1471-1481.
Takasugi, T. and Pope, D.P. (1983) Mater. Sci. Eng:-57, 15-20.
Takasugi, T. and Vitek, V. (1981) Metall. Trans. 12A~659-665.
Taplin, D.M.R. and Collins, A.L.W. (1978) Ann. ReV:-Mater. Sci. 8, 235-268.
Taplin, D.M.R. and Wingrove, A.L. (1967) Acta Metall. 15, 1231-1236.
Tauber, G. and Grabke, H.J. (1978) Ber. Bunsenges. PhyS: Chern. 82, 298-302.
Thomas, G.B. and Gibbons, T.B. (1984) Mater. Sci. Eng. 67, 13-2~
Thornton, D.V. (1972) International Conference on Creep~esistance in Steel,
Verein Deutscher Eisenhuttenleute, Dusseldorf, VI 5.
Timoshenko. S. and Goodier, J.N. (1951) Theory of Elasticity, McGraw-Hill, New
York.
Tipler, H.R. (1980) in Residuals, Additives and Materials Properties, Phil.
Trans. R. Soc. Lond. A295, University Press, Cambridge, pp. 151-164.
Tipler, H.R. and Hopkins, B.E. (1976) Metal Sci. 10, 47-56.
Tomkins, B. (1981) In: Creep and Fatigue in High Temperature Alloys, J.
Bressers, Ed., Applied Science Publishers, London, pp. 73-110 and 111-143.
Towle, D.J. and Jones, H. (1976) Acta Metall. 24, 399-407.
Trinkaus, H. (1978) Ber. Bunsenges. Phys. Chem:-82, 249-253.
Trinkaus, H. (1979) phys. stat. sol. (b) 93, 293-303.
Trinkaus, H. (1983) Phys. Rev. B27, 7372-7378.
Trinkaus, H. and Ullmaier, H. (1979) Phil. Mag. A 39, 563-580.
Trinkaus, H. and Yoo, M.H. (1985) private communication, submitted to Phil.
Mag.
Turnbull, D. (1956) In: Solid State Physics, vol. 3, F. Seitz and D. Turnbull,
Eds., Academic Press, New York, pp. 268-306.
Turnbull, D. and Fisher, J.C. (1949) J. Chern. Phys. 17, 71.
Tvergaard, V. (1984a) J. Mech. Phys. Solids 32, 373-393.
Tvergaard, V. (1984b) Acta Metall. 32, 1977-1990.
Tvergaard, V. (1985) Int. J. Solids;Structures 21, 279-293.
Vitek, V. (1978) Acta Metall. 26, 1345-1356.
-Vitek, V. (1980) Metal Sci. 14~403-407.
Volmer, M. (1939) Kinetik de;-Phasenbildung, Steinkopf-Verlag, Dresden und
Leipzig.
Volmer, M. and Weber A. (1926) Z. Phys. Chern. 119, 277-301.
Wang, J.S., Martinez, L. and Nix, W.D. (1983) Acta Metall. 31, 873-881.
Watanabe, T. (1983) Metall. Trans 14A, 531-545.
-Watanabe, T. and Davies, P.W. (197ar--Phil. Mag. 37, 649-681.
Wareing, J. (1983) In: Fatigue at High Temperature, R.P. Skelton, Ed., Applied
Science Publishers, London and New York, pp. 135-185.

416

References

Weber, J.H. and Gilman, P.S. (1984) Scripta Met. 18, 479-482.
Weertman, J. (1955) J. Appl. Phys. 26, 1213-1217.-Weertman, J. (1974) Metall. Trans. 5; 1743-1751.
Weertman, J.R. (1979) Canad. Metall~ Qu. 18, 73-81.
Westmann, R.A. (1965) Trans. ASME, J. AppI: Mech. 32, 411-417.
Westwood, H.J. and Taplin, D.M.R. (1975) J. Aust. Inst. Metals 20, 141-149.
White, C.L. and Padgett, R.A. (1982) Scripta Met. 6, 461-466.
White, C.L. and Padgett, R.A. (1983) Acta Metall. 3, 1005-1012.
White, C.L., Padgett, R.A. and Swindeman, R.W. (1981) Sripta Metall. 12,
777-782.
Wilkinson, D.S. and Vitek, V. (1982) Acta Metall. 30, 1723-1732.
Williams, K.R. and Cane, B.J. (1979) Mater. Sci. Eng. 38, 199-210.
Williams, K.R. and Wilshire, B. (1977) Mater. Sci. Eng:-28, 289-296.
Williams, K.R. and Wilshire, B. (1981) Mater. Sci. Eng. 47, 151-160.
Williams, T.M., Harries, D.R. and Furnival J. (1972) J. Iron Steel Inst. ~,
351-358.
Wingrove, A.L. and Taplin, D.M.R. (1969a) Scripta Metall. 3, 649-654.
Wingrove, A.L. and Taplin, D.M.R. (1969b) J. Mater. Sci. 4~ 789-796.
Wray, P.J. (1969) J. Appl. Phys. 40, 4018-4029.
Wu, F.-H., Bassani, J.L. and Vite~ V. (1986) J. Mech. Phys. Solids 34,
455-475.
Wathrich, C. (1982) Int. J. Fracture 20, R35-R37.
Wynblatt, P. and Ku, R.C. (1979) in Interfacial Segregation, W.C. Johnson and
J.M. Blakely, Eds., American Society for Metals, Metals Park, OhiO, pp.
115-136.
Yamaguchi, K. and Kanazawa, K. (1979) Metall. Trans. lOA, 1445-1451.
Yamaguchi, K., Kanazawa, K. and Yoshida, s. (1978) Mater. Sci. Eng. 33,
175-181.
Yang, W. and Freund, L.B. (1985) Int. J. Solids Structures 21, 977-994.
Yang, W. and Freund, L.B. (1986) Int. J. Fracture 30, 157-174.
Yoo, M.H., Ogle, J.C., Borie, B.S., Lee, E.H. and Hendricks, R.W. (1982) Acta
Metall. 30, 1733-1742.
Yoo, M.H., White, C.L. and Trinkaus, H. (1985) in Flow and Fracture at Elevated
Temperatures, R. Raj, Ed., American Society for Metals, Metals Park, Ohio,
pp. 349-382.
Yu, J. and Grabke, H.J. (1983) Metal Sci. 17, 389-396.
Zeldovlch, J.B. (1943) Acta physicochim. u:8.S.R. 18, 1.
Zienkiewicz, O.C. t1971) The Finite Element Method~n Engineering SCience,
McGraw-Hill, London.

Index

Airy stress function 31,32,93,95,


106,391-392
aluminum 14,53,55,84,136,248,350,
360-363
angular distribution of cavitating
boundaries 56
annealing of creep damage 156-157
anti-plane shear 29,312-314
bicrystals 56,58,154,172,200,213,
237-238
blunting (see crack opening
displacement)
Boltzmann's constant 5
Borisov 158-159
boron 117-119,390
carbide 7,16,54,57.59,60,69,102,120,
133,135,138,140-146,159,243,266,
281,325,369
cavity growth 148-246
cavity size distribution 225-227
ceramics 13,15,17,55,147,223,242,
244,346
characteristic time 81,91,99-101,112,
113,221,308-309,324,325,330,331,
335-339,341-344,348,358,374,375
chemical potential 42,45,125,132,
140,141
Chuang and Rice model 160-170
C!-integral 36,89,263-385,396-400
Ch-integra1334-337,359
cleavage 16-17,53
cohesive strength 61-62,66,67-69,88,
116,118,120,127-129,140,146,243,
266
compatibility 27,28,104,107,111,174,
204,290,303
constitutive models 5-6,27-28,185-196,
327,332,338-339,346,349-350
constrained cavity growth 172-197,200,
219,231-241,244-246,252-254,256,
273,284
constrained grain boundary sliding 8,
10-11,214,254-255
copper 23,54,56-58,60,63,65,131,154,
169,200,213,220,249,256,266,389,
350,360-363
corrosion 15,266,325-326.351.368-369,
386

cosegregation 122.123
crack opening displacement 270.
286-288.293.295.365-366.370.
372.377.379.384
crack closure 366.368.372.377.384-385
creep crack growth tests 281-283.
297-298.324-326.331.342-345.363
creep curve 3.4.25,194-196,339
creep embrittlement 117-119
creep-fatigue 247-260.364-386
creep zone 99.306-308.329-330,
334-335,357-358,375
cyclic loading 247-260.364-386
damage (different kinds of) 14
damage mechanics 349-363
damage parameter 24-26,349-350.353.359
decohesion (see cohesive strength)
deformation map 11-13
density 52,240-241
diffusion 5-7,40-48,123,130,389-390
diffusion creep 6-8,96-98,110-112,
346-348
dimensional analysis
(see similarity solutions)
dimpled fracture 18
dislocation 4,5,12-14,16,198
elastic-viscous analogy 33,94,107,
269,328,346
electron microscopy 52
equilibrium 27.121,142,143,204,205,
292,303
fatigue 247-260,364-386
finite elements 10-11,39,94-96,106,194,
210,212,216,269,292,294,296-300,
303,305,307-310,316,318,351,
354-355.362-363,376.396
Fokker-Planck equation 77-82
fracture mechanics 263-386
fracture-mechanism map 15-22
free energy 72-75
free enthalpy 121,133.134
gas bubbles 131-139,159,169,243
gas constant 5
Gibbs adsorption equation 124-126
grain boundary energy 68,69,71,124-126
grain boundary migration 21,55,58,242
grain boundary sliding 8-11,13,56,
85-115,212-214,254-255

Index

418

heat affected zone 59,63,119


helium 60,139
Hoff equation 20-21
hot shortness 22
HRR-fields 36,37,89,271-274,287,

292-293,302,306,312,316-319,329,
334-336,340,352-355,372,374-376
Hull and Rimmer model 148
hydrogen attack 117,136-138,160
hyperbolic-sine creep law 312
ideal strength 67-69
impurities 17,63,69,116-130,158
incubation time 80-82
intermetallic compounds 17,55,117
internal stress 5,140-147,338-339
J-integral 34,35,38,327-331
Kachanov equations 24-26,349-350
Langmuir-McLean isotherm 122
Larson-Miller parameter 24
ledge 57,114
load parameter map 336-337,341-342,
347,356
magnesium alloys 54,59,60,249
Monkman-Grant product 21,23,26,91,
154,193,195-197,200,201,
210-212,230,244,273,353
multiaxial loading 28,86,173-174,
186,191,203,205-209,216,349
necking 18-21,196,197
neutron scattering 52
nickel alloys 7,14-15,54,55,57,60,
63 -6 6 , 70 ,84 , 117 -1 18 , 131 -1 36 , 157 ,
159,172,182-184,239-240,243,
255-256,266,287-288,324-326,
368,369,389
Norton's creep law 5,28-31,105,186,
205,268,270, 28 1,389-390
notch 311,392-393
notch strengthening/weakening 361
nucleation of cavities 57-60,62-147,
194-196,225-241,277-278,280-281
nucleation stress 65-66,78-80,83,
129,130,277,28i,284
oxide particle or layer 15,55,56,59,
60,69,118,133,136,243,266,367-369
oxygen attack 131-136
particle coarsening 14,124,281,325
penny-shaped crack 87,90,173,186,190
plane strain 32,93,95,97,293,307,308
354-355,361,396-400
plane stress 31,307,308,396-399
plastiC zone 328
power-law breakdown 312
primary creep 4,332-345,359
process zone 351,354-355
recovery 4,338-339
recrystallization 21

replication technique 53
rupture lifetime 154-156,167-169,

181-185,193-196,200,209-211,
214,224,228-240,244-246
by creep crack growth 344,359-363
scaling law 33,34,269
secondary creep 4,26,267-271,359
segregation 116-130,158
self-consistent method 11,188-189,212
side grooves 282,296-300
silver 7,55,169-170,224
similarity solutions 38,46,80-81,
99 -100,113,164,201 ,215,226,
302-304,315,316,318,329,335,
340,352,357,358,375,376
slip band 16,57,84,86,243,250
small-scale creep 302-306,315,321,356
small-scale damage 351-356,359,
362-363
s tee 1 s 7, 1 4-1 8 ,21 ,22, 5 3-5 6 , 59 , 60 , 6 3-6 5 ,
118-120,123,126,136-138,158,160,
172,175-179,234-238,241 ,257-258,
266,281-284,324-325,331,342-345,
350,368-369,389
strain hardening 4,327,332
stress concentrations
at cracks 36-38,263-385
at ledges 114
at particles 102-114
at slip bands 86-91
at triple grain junctions 92-101
stress-enhancement factor 10,11,

96,192

stress intensity factor 38,87,88,

102,105,222,301-326,346-348,
356-358,364,372-377,393-394
stress relief cracking 59,62,119-121
Stroh-McLean transition 61
sulfur and sulfide 59,63,69,118-121
superplasticity 18-21
surface diffusion 45,160-171
surface tension 71,151,153,156-157
temper embrittlement 116-117
tertiary creep 4,26,194-196,339
thermal expansion 146,147
threshold stress 7,8,112
threshold stress intensity 364
titanium 55
trace elements (see impurities)
triple junction 59,61,92-102
tungsten 55
vacancy 6,40,69,70,74,75,129,139,

148-149,198

wedge crack 61-62,254-255


welding 22,59,63,117,119-121
Z-integral 372-373,379-386
zirconium 55,116-118,158-159,389

Potrebbero piacerti anche