You are on page 1of 285

Materials Research and Engineering

Edited by B. llschner and N. 1. Grant

Ismail C. Noyan Jerome B. Cohen

Residual Stress
Measurement by Diffraction
and Interpretation

With 160 Figures and 31 Tables

Springer-Verlag
New York Berlin Heidelberg
London Paris Tokyo

Dr. ISMAIL CEVDET NOY AN


Thomas J. Watson Research Center, IBM
P.O. Box 218
Yorktown Heights, NY 10598/USA

Prof. JEROME B. COHEN


Dept. of Materials Science and Engineering
Dean, The Technological Institute, Northwestern University
Evanston, IL 60201/USA

Series Editors

Prof. BERNHARD ILSCHNER


Laboratoire de Metallurgie Mecanique
Departement des Materiaux, Ecole Polytechnique Federale
CH-1007 Lausanne/Switzerland

Prof. NICHOLAS J. GRANT


Dept. of Materials Science and Engineering, MIT
Cambridge, MA 02139/USA

Library of Congress Cataloging in Publication Data.


Noyan, I.C. (Ismail Cevdet), 1956-- . Residual stress.
(Materials research and engineering)
Includes bibliographies and index.
1. Residual stresses - Measurement. 2. Nondestructive testing.
I. Cohen, J. B. (Jerome Bernard), 1932- . II. Title. III. Series.
TA417.6.N68
1986.
620.1'123'0287
86-21919
ISBN-13: 978-1-4613-9571-3
e-ISBN-13: 978-1-4613-9570-6
001: 10.1007/978-1-4613-9570-6
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.

Springer-Verlag New York Inc. 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.
Typesetting: With a system of Springer Produktions-Gesellschaft, Berlin
Dataconversion: Briihlsche Universitiitsdruckerei, Giessen
Offsetprinting: Ruksaldruck, Berlin. Bookbinding: Liideritz & Bauer, Berlin
2161/3020-543210

Preface

As industries find that the market for their goods and services is often as closely
connected to their quality as to their price, they become more interested in
inspection and quality control. Non-destructive testing is one aspect of this topic;
the subject of this book is a sub-field of this domain. The techniques for measuring
residual stresses have a long history for a technological subject. Yet, in the last
decade or so there has been renewed and vigorous interest, and, as a result of this,
there has been considerable progress in our understanding and in our methods. It
seemed a proper time to bring the new material together in an organized form
suitable for a course or for self-teaching, hence this book.
After an initial introduction to the qualitative ideas concerning the origin,
role, and measurement of residual stresses, we follow with chapters on classical
elasticity and the relatively new subject of microplasticity. These are primarily
introductory or review in nature, and the reader will find it important to consider
further the quoted references if he is to be involved in a continuing basis in this
area. There follows a chapter on diffraction theory, and then we fuse these subjects
with a chapter on diffraction techniques for measuring stresses and strains which
at present is our most general tool for non-destructive evaluation in this area.
In Chap. 6 we explore how to evaluate errors in such a measurement, a topic
almost as important as the stress itself and certainly vital in automating the
measurement. Practical examples are described in Chap. 7 and means for
measuring the strain distribution, in Chap. 8, in contrast to the measurement of
average values described in the earlier chapters.
The Appendices include solutions to the problems given at the end of some of
the chapters, Fourier analysis, and sources of useful data.
We emphasize diffraction because in our opinion it is the most well understood and reliable tool available for the measurement of stresses. Of course our
own background in niffraction colors our opinion, but we hope this book will
reveal to the reader why we feel this way.
Dr. M. James provided Appendices E and F, and Mr. W. P. Evans supplied
many practical examples for Chap. 7. Both gave us helpful comments on the
manuscript.
Both authors thank our many co-workers in this field, Drs. R. Marion,
M. James, H. Dolle,W. Schlosberg, Prof. Jai Wen Ho, Mrs. Rui Mai Zhong,
Messrs. T. Devine, W. Evans, and P. Rudnik. The two of us have enjoyed (most
of the time) working together over the past five years on research in this field and
on this book. One of us (I. C. N.) especially acknowledges the support of his
family.
I. C. Noyan
Evanston, IL, April 1987
J. B. Cohen

Dedication
We dedicate this book to the U.S. Office of Naval Research and particularly to
Dr. Bruce MacDonald of this office. His foresight led us into this field and his
helpful encouragement and support led us along many interesting avenues of
research on residual stresses.

Contents

1 Introduction . . . . . . . . . . . . . .

1.1 The Origin of Stresses . . . . . . . .


1.2 Methods of Measuring Residual Stresses
1.3 Some Examples of Residual Stresses
References. . . . . . . . . . . . . . .

1
4
7
12

2 Fundamental Concepts in Stress Analysis .

13

2.1 Introduction . .
2.2 Definitions . . . .
2.3 Stress and Strain. .
2.4 Forces and Stresses
2.5 Displacements and Strains
2.6 Transformation of Axes and Tensor Notation
2.7 Elastic Stress-Strain Relations for Isotropic Materials.
2.8 Structure of Single Crystals . . . . . . . . .
2.9 Elastic Stress-Strain Relations in Single Crystals
2.10 Equations of Equilibrium. . . . . . .
2.11 Conditions of Compatibility. . . . . .
2.12 Basic Definitions in Plastic Deformation
2.13 Plastic Deformation of Single Crystals .
2.14 Deformation and Yielding in Inhomogeneous Materials.
Problems . .
Bibliography. . . . . . . . . . . . . . . . . . . . . .

13
13
14
15
17
20
25
28
32
37
38
39
41
44
45
46

3 Analysis of Residual Stress Fields Using Linear Elasticity Theory

47

3.1
3.2
3.3
3.4
3.5
3.6
3.7
3.8
3.9
3.10
3.11
3.12

47
47
51
52
54
56
61
62
63
65
66
69

Introduction . . . . . . . . . . . . .
Macroresidual Stresses . . . . . . . . .
Equations of Equilibrium for Macrostresses
Microstresses . . . . . . . . . . . . .
Equations of Equilibrium for Micro- and Pseudo-Macrostresses
Calculation of Micro- and PM Stresses . . . . . . . . . . .
The Total Stress State in Surface Deformed Multiphase Materials
Macroscopic Averages of Single Crystal Elastic Constants .
The Voigt Average. . . . . . . . . . . . . . . .
The Reuss Average . . . . . . . . . . . . . . .
Other Approaches to Elastic Constant Determination.
Average Diffraction Elastic Constants . . . . . . .

VIII

Contents

Summary
References.

73
73

4 Fundamental Concepts in X-ray Diffraction.

75

4.1 Introduction . . . . . . . . . . . .
4.2 Fundamentals of X-rays . . . . . . .
4.3 Short-wavelength Limit and the Continuous Spectrum
4.4 Characteristic Radiation Lines .
4.5 X-ray Sources. . . .
4.6 Absorption of X-rays.
4.7 Filtering of X-rays. .
4.8 Scattering of X-rays .
4.9 Scattering from Planes of Atoms.
4.10 The Structure Factor of a Unit Cell
4.11 Experimental Utilization of Bragg's Law
4.12 Monochromators . . . . . . . . . .
4.13 Collimators and Slits. . . . . . . . .
4.14 Diffraction Patterns from Single Crystals
4.15 Diffraction Patterns from Polycrystalline Specimens
4.16 Basic Diffractometer Geometry . . . . . .
4.17 Intensity of Diffracted Lines for Polycrystals
4.18 Multiplicity. . . .
4.19 Lorentz Factor . .
4.20 Absorption Factor .
4.21 Temperature Factor
4.22 X-ray Detectors . .
4.23 Deadtime Correction for Detection Systems
4.24 Total Diffracted Intensity at a Given Angle 28 .
4.25 Depth of Penetration of X-rays . . . . . . .
4.26 Fundamental Concepts in Neutron Diffraction.
4.27 Scattering 'and Absorption of Neutrons
Problems . . . . . . . . .
Bibliography and References. . . . . . .

75
75

76

77

80
82
84
84
86
87
89

90
91

93
94
95
97
98
98
100
103
103

108
109

110
111
114
116
116

5 Determination of Strain and Stress Fields by Diffraction Methods

117

5.1 Introduction . . . . . . . . . . . . . . . . . .
5.2 Fundamental Equations of X-ray Strain Determination
5.3 Analysis of Regular "d" vs. sin 2 \jJ Data. . . . .
5.4 Determination of Stresses from Diffraction Data.
5.5 Biaxial Stress Analysis . . . . . . . . . . . .
5.6 Triaxial Stress Analysis. . . . . . . . . . . .
5.7 Determination of the Unstressed Lattice Spacing.
5.8 Effect of Homogeneity of the Strain Distribution and Specimen
Anisotropy . . . . . . . . . . . . . . . . . .
5.9 Average Strain Data from Single Crystal Specimens . . . . . .

117
117
119
120
122
125
126
130
131

IX

Contents

5.10 Interpretation of the Average X-ray Strain Data Measured from


Polycrystalline Specimens. . . . . . . . . . . . . . . . .
5.11 Interpretation of Average Stress States in Polycrystalline Specimens .
5.12 Effect of Stress Gradients Normal to the Surface on d vs.
sin 2 tp Data . . . . . . . . . . . . . . . . . . . .
5.13 Experimental Determination of X-ray Elastic Constants.
5.14 Determination of Stresses from Oscillatory Data.
5.15 Stress Measurements with Neutron Diffraction.
5.16 Effect of Composition Gradients with Depth
5.17 X-ray Determination of Yielding.
..
5.18 Summary .
Problem . .
References . . .

140
145
153
154
157
159
160
162
162

6 Experimental Errors Associated with the X-ray Measurement of


Residual Stress. . . . . . . . . . . . . . . . . . . . .

164

6.1
6.2
6.3

Introduction . . . . . . . . . . . . . . . . . . . .
Selection of the Diffraction Peak for Stress Measurements .
Peak Location. . . . . . . . . . . . . . . .
6.3.1 Half-Value Breadth and Centroid Methods.
6.3.2 Functional Representations of X-ray Peaks.
6.3.3 Peak Determination by Fitting a Parabola .
6.3.4 Determination of Peak Shift . . . . . . .
6.4 Determination of Peak Position for Asymmetric Peaks
6.5 Statistical Errors Associated with the X-ray Measurement 'of
Line Profiles . . . . . .
6.6 Statistical Errors in Stress.
6.6.1 The sin 2 tp Technique
6.6.2 Two-Tilt Technique.
6.6.3 Triaxial Stress Analysis
6.6.4 Statistical Errors in X-ray Elastic Constants
6.7 Instrumental Errors in Residual Stress Analysis .
6.7.1 Variatiol). of the Focal Point with 9 and tp .
6.7.2 Effect of Horizontal Divergence on Focusing.
6.7.3 Effect of Vertical Beam Divergence . . . . .
6.7.4 Effect of Specimen Displacement . . . . . .
6.7.5 Effect oftp-axis not Corresponding to the 29-axis .
6.7.6 Error Equations for the tp-Goniometer . . . . .
6.7.7 Effect of Errors in the True Zero Position of the tp-axis
6.7.8 Alignment Procedures. . . . .
6.8 Corrections for Macrostress Gradients
6.9 Corrections for Layer Removal
6.10 Summary.
Problems .
References. . .

136
137

164
164
166
167
168
171
175
178
181
186
186
187
187
189
190
191
192
195
196
199
200
202
. 203
. 205
206
208
209
209

Contents

7 The Practical Use of X-ray Techniques.

211

7.1
7.2
7.3
7.4
7.5

Introduction . . . . . . . . . .
The Use of Ordinary Diffractometers .
Software and Hardware Requirements
Available Instruments . . . . . . .
Selected Applications of a Portable X-ray Residual Stress Unit
(By W. P. Evans) .
Reference . . . . . . . . . . . . . . . . . . . . . .

211
211
212
213

8 The Shape of Diffraction Peaks - X-ray Line Broadening.

230

8.1 Introduction .
8.2 Slit Corrections
8.3 Fourier Analysis
Problem. .
References. . . . .

. . . . . . .

230
233
238
245
247

Appendix A: Solutions to Problems.

248

Appendix B . . . . . . . . . .
B.l Introduction . . . . . . . .
B.2 The Marion-Cohen Method. .
B.3 Dolle-Hauk Method (Oscillation-free Reflections)
B.4 Methods of Peiter and Lode. .
B.5 Use of High Multiplicity Peaks
References. . . . . . . . .

252
252
252
254
256
257
257

Appendix C: Fourier Analysis . . .

259

Appendix D: Location of Useful Information in "International Tables


for Crystallography" . . . . . . . . . . . . . . . .

266

Appendix E: Values of G x for Various Materials


(8y Dr. M. James) . . . . . . .

267

Appendix F: A Compilation of X-ray Elastic Constants


(By Dr. M. James)
References. . .

270
271

Subject Index

273

. . . . . . . . . .
. . . . . . . . . .
of Peak Broadening.

214
229

1 Introduction

In 200 B.C., Chinese artisans manufactured thick bronze discs that were flat and
polished on one side and had a relief cast on the other face. These were heated and
quenched. When such a "magic mirror" was illuminated on the flat face, the
reflection showed the pattern of the relief on the other side of the disc! Due to the
different cooling rates of the sections with various thicknesses, distortions occurred
on the flat side, which mimicked the pattern of the relief. To our knowledge, this is
the first deliberate use of residual stresses and strains. Today we know these playa
key role in the behavior of welded structures (and hence in ship construction,
pipelines and oil rigs) , in the response of heat treated or finished parts (ground
gears, shot-peened or sand blasted pieces, or material subjected to laser heat
treatments, or quenched after a heat treatment). These stresses are also a key
factor in the fatigue response of solids, and in the phenomenon known as stress
corrosion.

1.1 The Origin of Stresses


Residual stresses can arise in materials in almost every step of processing, as, for
example, when a material is subjected to heat treatment or machining. Consider
first a material that undergoes no change in crystal structure during heat
treatment. If aluminium is cooled quickly from the heat treatment temperature,
the surface and the interior contract at different rates, as illustrated in Fig. lola. At
some time the difference, coupled with low material yield strength associated with
the high temperature, induces plastic flow or permanent yielding. The surface
regions, which because of the temperature gradient want to contract on cooling
more than the interior, are extended by the interior and vice versa. (Note the
increase in length near the surface at time A in Fig. l.la.) This is a real effect; for
iron-based materials the product of Young's modulus and coefficient of
expansion yields a stress of 3.5 MPa (O.5ksi) per DC difference in temperature
between two such regions.
On continued cooling to room temperature, point B in Fig. l.la, the surface
regions have been eXtended relative to the interior and consequently end up in
compression. These residual surface stresses are important because notches,
scratches, sharp changes in cross-section, etc. concentrate additional applied
tensile stresses near the surface and can act to initiate a crack. Residual
compressive stresses in the surface must be overcome by the applied load to initiate
cracks, and thus the presence of surface compressive stresses is a highly favorable
condition.

1 Introduction

Heat treatment does not always produce surface compressive stresses. If a


material undergoes a phase transformation, as in the hardening of steel to form
martensite, the local yielding is essentially masked by the large volume expansion
associated with the austenite to martensite phase change. The result is illustrated
schematically in the cooling curve in Fig. 1.1 b. At temperature A, the surface
regions transform to martensite and expand since they reach the transformation
temperature first. The interior, composed of low strength austenite, deforms
plastically to partially accommodate this change. At B, the interior transforms to
martensite producing an expansion which is resisted by the high strength
martensitic surface. At C, near room temperature, the surface is thrown into
tension by the interior, producing surface residual tensile stresses, which can
contribute to crack initiation and propagation. Stress relief annealing at moderate
temperatures is often employed to allow local yielding to occur, thereby
minimizing or eliminating such stresses. However, care is needed because even
differences in the coefficients of expansion between the carbides and ferrite phases
in steel can lead to significant stresses if the parts are not cooled slowly after this
treatment.
One way to replace surface tensile stresses with compressive stresses is to shot
peen the surface. In this process, high velocity shot causes local plastic yielding in
the surface, which IS thereby extended relative to the interior. The interior acts to
constrict the surface, resulting in high, local, compressive residual stresses in the
surface balanced by tensile stresses within the interior, as illustrated in Fig. l.lc.
(However, in certain cases, overpeening, or use of too high a peening velocity can
cause relaxation at the surface. )
Even releasing after a tensile extension of a specimen into the plastic region can
produce residual stresses. If the surface is harder than the interior because of defect
pile-up occurring during plastic extension, then on release of the load (at A in
Fig. l.ld), the elastic recovery tries to leave the surface shorter than the interior,
resulting in surface tensile stresses. The reverse occurs when the surface is softer
than the interior. In a forming operation, such as rolling, the surface can be
extended more than the interior due to friction at the rolls, as illustrated in
Fig. 1.1e, resultIng in compression on the surface. In this operation the magnitude
of the stress is a function of the thickness of the piece, the roll diameter and the
degree of reduction [1].
Another important cause of residual stress is welding, as illustrated in Fig. 1.1 f.
Contraction of molten weld metal during solidification is resisted by colder
surrounding metal, resulting in the stresses illustrated in the figure. (For further
information on stresses in this process see Refs. [2] and [3].) From these
examples (see also Refs. [4 - 6] ), it is clear that residual stresses in materials
arise, not only in processing, but also in use. A rapid, easy-to-use, nondestructive
method to measure these stresses is highly desirable. Unfortunately, in too many
cases, residual stresses are ignored or it is assumed that additional treatment has
either eliminated them or introduced compressive stresses. To further complicate
the issue, the macrostresses (that arise because of one part of a body acting on
another) can develop differently in different regions of the same piece. Microstresses can arise in microscopic regions, such as between the microscopically sized
phases in a multiphase material. The magnitudes of these various kinds of stresses

1.1 The Origin of Stresses

Cooling curves
.:

.:

c;,
c:

OJ
-I

Q,
c

local
yielding

OJ
-I

Phase
transition

I
I

Time

Time
b

Strain

Fig. 1.1 a,b. Schematic cooling curves during heat treatment showing the difference in contraction
of the surface and interior. In a there is no phase transition, whereas one occurs in b. c Surface
deformation by peening elongates surface regions. 1, Surface element elongated by peening put
into compression by interior. d The harder surface regions contract more (on release on the load
at A) than the softer interior. 1, Hard region; 2, soft region. e Friction at the rolls in a rolling mill
tends to extend surface regions more than the interior. f Stresses due to welding. 1, Longitudinal
shrinkage; 2 weld metal; 3 lateral shrinkage; 4 stresses

1 Introduction

can be a significant portion (half or more) of the ultimate tensile strength of the
annealed material. Furthermore, in addition to residual stresses, it is sometimes of
interest to measure applied stresses.

1.2 Methods of Measuring Residual Stresses


In view of their importance, it is not surprising that there has been a continuing
interest in developing methods for measuring residual stresses. In fact, over the
past few years, one or more conferences have been held annually on this subject
[7 -11]. What is surprising is how seldom these stresses are actually measured
and monitored in fabrication or in service!
One method for measuring residual stresses in a selected component is to drill a
hole and measure subsequent relaxation around the hole (the distortion) with
electrical-resistance strain gauges. Care is needed in making the hole to avoid
introducing new stresses, and to avoid or correct for distortion due to the stress
concentrating effects of the hole. Drilling is destructive, and furthermore,
microstresses cannot be evaluated in this way [12].
Another technique receiving great interest at the moment is based on acoustic
wave propagation [13]. The velocity of a wave in a solid is relatively easy to
measure using simple equipment; it depends upon the square root of the elastic
constant (M). In the elastic region, stress (cr) and strain (E) are usually written
as a proportionality:
cr=ME.

(1.1 )

But actually there are higher order elastic constants:


cr=ME+CE2 +DE 3 + ....

(1.2 )

Therefore, taking the first two terms of (1.2), a better approximation is:
cr~

(M+CE)E=M'E.

(1.3 )

Consequently, the velocity of a wave V is:


V=K]IM'=KVM+CE.

(1.4 )

The wave velocity depends on the state of strain, E, in the material, and hence on
the residual stresses. Unfortunately, the distribution of phases present can
produce effects that so far have prevented general use of this technique. Some
progress is being made however (see Ref. [14]) but, so far, only for uniaxial
loading.
Still another stress measuring method is associated with the Barkhausen
"noise" in magnetic materials, produced when magnetic domains are moved by a
field close to the material [15]. This noise is sensitive to the stress fields in the
material since the fields affect the ease of rotation of the domains. The measuring
equipment is simple, but unfortunately the technique is limited to ferrous alloys.
Furthermore, the noise signal saturates at about 500 MPa (70,000 psi) in either

1.2 Methods of Measuring Residual Stresses

28

,...

I \
I \
I

____ J

28

Specimen

Fig. 1.2. a Schematic of diffractometer. The incident beam diffracts X-rays .ofwavelength A. from
planes that satisfy Bragg's law in crystals with these planes parallel to the sample's surface.
If the surface is in compression, because of Poisson's ratio these planes are further apart than in
the stress-free state. The d spacing is obtained from the peak in intensity versus scattering angle 20
and Bragg's law A. = 2d sin O. b After the specimen is tilted, diffraction occurs from other grains,
but from the same planes, and these are more nearly perpendicular to the stress. These planes are
less separated than in a. The peak occurs at higher angles of20. c After the specimen is tilted, the
stress is measured in a direction which is the intersection of the circle of tilt and the surface of the
specimen

tension or compres~ion so that the method is most useful (after calibration) for
determining the sign of the stress, but not its magnitude. More details on magnetic
and acoustic technique are reviewed in Ref. [16].
Plastic models can be employed to simulate a part, or a production method,
and the stress obtained from the birefringence oflight passed through the plastic
[17]. Progress is also being made in calculating the stress pattern, see papers in
Ref. [11] by Ericsson and Hildenwall, and Yu and Macherauch.

1 Introduction

X-ray methods of measuring residual stresses in crystalline materials have been


tested and compared to other methods [17 -18] and are in use throughout the
world. Manuals have been written for their use in the United States and Japan
[18 - 20] (although these are now outdated they still contain much useful
information). Usually the measurements are made manually on a conventional
powder x-ray diffractometer or with a special unit. But there are other units that
can make measurements in seconds (in the field) with reasonable precision. These
will be discussed in Chap. 7.
The basis of the technique is straight forward and was first used by Aborn at
the U.S. Steel Co. [21]. First consider the method of x-ray powder diffraction
using a single wavelength (A) and a flat specimen in a conventional diffractometer. With reference to Fig. 1.2, assume that the x-ray detector is moved over a
range of angles, 2e, to find the angle, e, of the diffraction from grains that satisfy
Bragg's law: that is, grains that have planes of atoms with interplanar spacing "d"
such that A= 2d sin e. Grains that have planes with this spacing which are parallel to
the surface will diffract as shown in Fig. 1.2a. This diffraction takes place from a thin
surface layer ( ~ 20 ~m). If the surface of the specimen is in compression, the "d"
spacing of these planes is larger than in the unstressed state, because of Poisson's
effect. If the specimen is now tilted with respect to the incoming beam (Fig. 1.2b ) ,
new grains diffract and the orientation of the diffraction planes is more nearly
perpendicular to the stress direction.
The result is that, with the tilt, the "d" spacing decreases and the angle 2 e
increases, as shown in the figure. In effect, the interplanar spacing acts as an
internal strain gage. Since the spacing of lattice planes (the "strain gages") is
extremely small they will be affected by both micro and macro stresses. The x-ray
method measures the sum of these stresses.
The fact that x-rays penetrate only a shallow depth is in fact useful, because in
many cases there are steep gradients in the macrostress near the surface. By using
neutrons (which can penetrate 2 - 3 cm of steel), we can average out the
macrostresses and sample only the microstresses, if this depth covers the entire
gradient. Neutrons from a reactor can also be employed to examine the stress at

Slit

Fig. 1.3. Neutron scattering. The beam penetrates the entire volume, but slits define a
local volume. By moving the slits as shown,
or the sample vertically, the stress in different volumes may be measured

1.3 Some Examples of Residual Stresses

various depths, using a slit as shown in Fig. 1.3. The resolution in this case is the
order of 1 - 2 mm, depending on the neutron intensity (that is the reactor
power).
Finally, it is worth mentioning that it is possible to monitor the development or
change in stresses in use, by first measuring the value with diffraction, and then
attaching strain gages or grids. Reference [22] describes some of these procedures.

1.3 Some Examples of Residual Stresses


It is appropriate at this point to give some examples of some actual stress

measurements and the useful information such results provide. References to the
recent literature will be given, so that the reader can obtain more details. (We rely
largely on our own work, because we know it best.)
In preparing samples for mechanical testing, surfaces are often ground, or
sometimes even polished metallographically. ,(S Fig. 1.4 shows, for an HSLA steel
such methods can produce quite large stresses, even at significant depths [23].
Because these techniques of surface preparation are not usually carefully
monitored, different stresses may develop in each sample, and could be the source
of scatter in the data. Although it is not generally done, measurement of such
stresses prior to testing could be quite revealing!
Most parts in machines, engines, etc. are subject to rapidly varying alternating
stresses as they move. This alternating stress is described as fatigue. If the part fails
in only a small number of cycles, this is called low-cycle fatigue. (The stresses in
this case are relatively large, but may be below the static yield.) As Fig. 1.5 shows,
the residual stresses have a different sign, depending on whether the part is
released from the tensile or compressive portion of the alternating load [23].
In high-cycle fatigue, at much lower stresses the behavior of residual stresses is
not yet understood, although changes do occur. In a part without any initial
stresses, stresses may develop, as shown in Fig. 1.6 for an annealed 1040 steel [24]
in tension-tension fatigue (i.e. the applied load is tensile, then reduced, then tensile
etc.). Note that these stresses are compressive and therefore not harmful. In fact,
300.---------------------------~

MPa

200

III
III

100

C:i

::l

"tl

ID -100

n:

-200
-3000~--~~--~~--~~--~~~

0.05

Thickness removed

Fig. 1.4. Residual stresses in an


HSLA steel due to surface preparation. Bulk plate specimens: 1::;., surface ground; 0, 600 grit metallographic polishing. From Ref. [23]

1 Introduction

III
III

300r----------------------------.
MPa
0 1 cycle. compo
200 I- " 100 cycles. compo
1 yccle. tension
100
... 100cycles . tension

Iii

O~"''''''

:J

Vi -100

&!

'f.
,....... --.,
...

.~~.~.~_r--~----~--

':I.

-&

-200 t11- - - - 11- - . - J


-300 0L-----..,.L1 --..L
--1
0.05
0.10
0.15
0.20 mm
Thickness removed

Fig. 1.5. Residual stresses in low


cycle fatigue of an HSLA steel.
Tension-compression fatigue released from either tension or compression, as indicated. From
Ref. [23]

-240
MPo

-200 l-

-160 r--

III
III

Cl>

';;-120
1III

a:J
u

Vi

-80

I-

-40

I-

Cl>

a:

01-40
1

~'

/1
I

10

10 2

10 3 101. 10 5 10 6 10 7 108
Number of cycles

Fig. 1.6. Residual stresses in an annealed 1040 steel as a result of


tension-tension
fatigue.
From
Ref. [24]

they impede crack propagation and undoubtedly add to the life of the part.
However, if stresses are induced by processing, they "fade" at a rate which
increases with the cycling stress. An example is shown in Fig. 1.7. In this case
(peened 1040 steel, tension-tension fatigue), the initially beneficial compressive
stresses not only decrease, but reverse in sign early in the stress cycling, and
decrease the life. (This reversal occurs only in tension-tension fatigue but not in
bending, in which case only fading occurs.)
A two-dimensional map of the stresses ahead of a fatigue crack are shown in
Fig. 1.8 [25J; the component shown is normal to the crack face. Such stresses exist
well beyond the plastic (shaded) region, caused by the stress concentration ahead
of the crack, due to plastic upset in the region. Note the stresses behind the crack
front, and that the stress is compressive at the tip, impeding crack propagation.
During machining (milling, turning, or grinding), the near-surface region of
the workpiece is deformed plastically. As volume elements in this region are
extended when the tool passes by, the constraints of the bulk should introduce
compressive stresses near the surface. Indeed, this does occur for shallow, slow,

1.3 Some Examples of Residual Stresses

-600 r-------------------------~

MPo (:--Uncycled value

-500 1"'.

-1.00 f- \

1Il - 300fIII

'"
~

-200 f-

-100

1:)

'iIi

0-

'"
cr:

Fig. 1.7. Residual stresses in a peened


1040 steel as a result of tensiontension fatigue. The two symbols
represent two samples. From Ref. [24]

80
~

100 -

200 -

102 10 3 10'

10

105 106

Number of cycles

2 .40
2.00
1.60

. 1.20
0.80
0 .40

If>

48

-~9

.,3

~2

16

- 05

0">

,"

~3

16

o.
- 0 .95 -0.55 -0 .1 5

0 .2 5

0 . 65

105.BI

\.05 1.45

1.9
1.0

1.85

(mm)

Fig. 1.8. Residual stresses around a fatigue crack produced by tension-tension fatigue of an HSLA
steel (in MPa). From Ref. [25]

well-lubricated cuts with a sharp tool. But, extensive studies of the stress pattern
have shown that the actual situation is usually much more complex. If strong work
hardening occurs only near the surface, this produces tensile residual stresses due
to the greater elastic relaxation of this region compared to the bulk. Local
compressive plastic deformation due to the pressure by the tool will cause tensile
residual stresses, and, if the sum of extermil stresses and residual stresses exceed
the local yield strength, plastic recovery will take place in this region. Even when
there is no cross-feed of the tool, a biaxial stress state has been observed. Heating,
due to the lack of lubrication or a dull tool or high downfeed, produces tensile
stresses because locally-heated regions are upset by the cooler surroundings. Thus,

1 Introduction

10
Table 1.1. Residual stress tensors (in MPa) for different samples

Specimen

Armco-iron

A1

151

A2

A6

n.c.

C1

567

-65

508

-4

-7

11

-7

28

-4

267

-12

-12

367

611

55

-8

507
4

74
39

80

90

10
n.c.

12
6
4

380
16

16
351

18
9

18

12

A5

Steel

-7

-8
A4

Specimen

-7

-4

A3

Parameters
of grinding'

645

39

C2

-65

-1

137

199

-10

-63

86
5

84

541

-20

-38

-20
-38

565

1
86

-17

59
6

-10
-63
C3

C4

408

-17
59

416

+-

390
14

14
306

-28

63

557

8
8

82

+-

5
C5

+-

534
-2

10

69

nc, no cooling; c, cooling; +t, final grinding direction,

C6

108
63
-1

-1

92

-2

69
-3

-3

95

468

Pi; 5,10, downfeed in J.UD

the stress patteni is altered by many parameters, such as the depth of cut, cooling
and cutting direction. For example, after a deeper cut, the stress pattern extends to
a depth greater than after a light cut, although the value of the stress near the
surface can be lower. The level of residual stress is also strongly affected in steel by
the carbon content, which influences the microplastic behavior of the material
significantly. Also, these residual stresses can affect dimensional stability.
We show some results in Table 1.1 for grinding Annco iron and medium
carbon steel [26], with a table motion of 0.33 ms - 1. The entire stress tensor is
shown in the fonn:

(Directions 1 and 2 are in the surface of the specimen, 3 is nonnal to the surface.
The stress cr 13 is in the direction 3 and acts on the face perpendicular to axis 1.)

1.3 Some Examples of Residual Stresses

11

The results of the study can be summarized as follows:


1) Residual stresses in a ground workpiece are confined to near surface layers; the
bending of such a piece cannot be employed to estimate the stresses.
2) Higher shear residual stresses developed in a medium carbon steel during
grinding than in Armco iron, probably due to the more random dislocation
distribution in the former.
3) High tensile residual normal stresses due to grinding can be produced in both
Armco iron and medium carbon steel.
4) The residual stress parallel to the grinding direction is directed opposite to the
horizontal direction of feed of the grinding wheel-in the last pass.
5) Increasing depth of cut increases the residual normal stresses.
6) Cooling decreases the tensile residual stresses in a medium carbon steel, but
increases them in Armco iron.
7) In steel (but to a lesser extent in iron) principal stresses in grinding are tilted
around an axis transverse to the grinding direction.
8) Flattening a ground specimen does not change the residual shear stresses,
although it alters the normal stresses.
Are induced stresses useful in reducing the wear rate? In some senses, wear is
like fatigue, as regions see alternating loads. In a recent study of this
possibility [27], it was found that unless the induced residual stresses were
very large, the wear process quickly eliminated them, and produced a pattern
of stresses versus depth characteristic of the wear process itself. Thus the time and
money involved in introducing such stresses could be wasted.
Residual stresses can be important in stress corrosion. One recent study in this
area is Ref. [28]. These stresses have an important effect on the chemical,
electrical and mechanical behavior of the thin films used so extensively in modem
solid-state electronics [29]. Laser heat treatment is becoming of increasing
importance, but the stresses produced are large and tensile [30].
While ceramics are normally thOUght to be brittle, significant residual stresses
can develop during processing and finishing. For example, compressive stresses of
135 -170 MPa develop during diamond grinding of Al 2 0 3 and extend to a depth
of 15 J.1m [31]. Even larger stresses ( 1000 MPa) , due to a phase transformation in
zirconia [32], have been reported in A1 2 0 3 /Zr0 2 composites.
Polymers are often amorphous or partially so and generally do not give a high
angle diffraction peak. However, the stresses can still be sampled by including a
small amount of metal powder in the preparation [33].
Further reviews of the stresses resulting from processing and use can be found
in Ref. [34].
Measurement of residual stresses via diffraction requires a familiarity and an
understanding of their origin, that is of mechanics and micromechanics, as well as
diffraction. These topics will be covered in subsequent chapters.
In Chaps. 2 and 3 we review the fundamental concepts of stress, strain, and
elasticity theory, with particular attention to those aspects pertinent to our
measurement. In Chap. 4, the fundamental aspects of the diffraction process are
covered, with particular attention to those factors that affect the position of a
diffraction peak. Then, in Chap. 5, we derive the fundamental equations relating

1 Introduction

12

the peak position to stress and strain. Various sources of error are evaluated in
Chap. 6. Portable stress analyzers and several examples of actual measurements
and their use in understanding a material's behavior are .described in
Chap. 7. In Chap. 8, we consider what additional information can be obtained
from the shape of a diffraction peak, which often changes due to stresses in
addition to shifting of the peak.

References
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34

W.M. Baldwin, Jr., Proc. ASTM 4a, 539 (1949)


V. Papazoglou and K. Masubuchi, Final Technical Report, ONR Contract No.
NOOO14-75 -C-0469
E. Macherauch, Welding Research Institute of London Reprint 11
G.M. Rassweiller and W.L. Grube (eds.), "Internal Stresses and Fracture in Metals", Van
Nostrand, New York (1959)
O.J. Horger, "Residual Stresses", Metals Eng. Design, ASME Handbook, 2nd ed., McGraw
Hill (1965)
F.A. McClintock and A.S. Argon, "Mechanical Behavior of Materials", Addison-Wesley,
Reading, MA (1966), pp. 420-442
Proceedings of a Workshop on Nondestructive Evaluation of Residual Stresses, AFML,
NTIAC-76-2, Southwest Research Center, San Antonio, TX (1976)
S. Taira (ed.), "X-ray Study on Strength and Deformation of Metals", The Society of
Materials Science, Kyoto, Japan (1971)
Hiirterei-Technische Mitteilungen 31, nos. 1 and 2 (1976)
Advances In X-ray Analysis 27, 29 (1984)
E. Kula and V. Weiss, Eds., "Residual Stresses and Stress Relaxation" 28, Sagamore Army
Materials Research Conference Proceedings, Plenum Press, New York (1982)
R.G. Bathgate, J. Brit. Soc. for Strain Measurement 4, No.2, 0 (1968)
G.A. Alers in Ref. 7
D.R. Allen and C.M. Sayers, Ultrasonics p.174 (July 1984)
e.G. Gardner, AMMRC CTR 72-22, Army Materials and Mechanics Research Center,
Watertown, MA. (1972)
M.R. James and O. Buck, CRC Critical Reviews in Solid-State and Materials Science 9
(No.1), 61 (1980)
M. Hetenyi, J. App!. Mech., Trans. ASME 60, A149, (1938)
A.L. Christenson, ed., "The Measurement of Stress by X-rays", SAE Technical Report 182,
Society of Automotive Engineers, New York, 1960
M.E. Hilley, ed., "Residual Stress Measurement by X-ray Diffraction", SAE Information
Report J784a, Society of Automotive Engineers, New York, Aug. 1971
Committee on Mechanical Behavior of Materials, "Standard Method for X-ray Stress
Measurement", The Society of Materials Science, Japan 1973
H.H. Lester and R.M. Aborn, Army Ordnance 6, 120, 200, 283, 364 (1925 - 26)
F.P. Chiang and C.e. Kin, J. of Metals, May, p. 49 (1983)
D. Quesnel, M. Meshii and J.B. Cohen, Mat. Sci. and Eng. 36(2), 207-215 (1978)
M. McClinton and J.B. Cohen, Mat. Sci. & Eng. 56, 259-263 (1982)
W.H. Schlosberg and J.B. Cohen, Met. Trans. 134, 1987 -1995 (1982)
H. Dolle and J.B. Cohen, Met. Trans. 11 A, 159 -164 (1980)
J.W. Ho, e. NQyan, J.B. Cohen, V.D. Khanna and Z. Eliezer, Wear 84, 183-202 (1983)
P. Doig and P.E.J. Flewitt, Met. Trans. 10, p. 413 (1979)
M. Murakami, CRC Critical Reviews of Solid State and Materials Science 11 No.4, p. 317
(1984 )
M.R. James, D.S. Gnanamuthu and R.J. Moores, Scripta Metal!. 18, p. 357 (1984)
F.R. Lange, M.R. James and DJ. Green, J. Am. Ceramic Soc. 66, C-16 (1983)
D.J. Green, F.F. Lange and M.R. James, J. Am. Ceramic Soc. 66, p.623 (1983)
P. Predecki and C.S. Barrett, J. Composite Materials 13, 61 (1979)
M.R. James and J.B. Cohen, Treatise on Materials Science and Technology 19A, p.l,
Academic Press, New York (1980)

2 Fundamental Concepts in Stress Analysis

2.1 Introduction
In this chapter the basic theorems of linear elasticity are reviewed, and some
fundamental definitions in plasticity theory are discussed. These concepts form the
basis of all types of stress analysis, and will be utilized in subsequent chapters for
the analysis of the deformation distributions that cause residual stress fields.

2.2 Definitions
All solid materials deform when subjected to external loads. The deformation is
manifested in displacements of the points in the body under load from their initial
(unloaded) positions. As long as the forces set up in the body due to external
loading are below a certain limit, the deformation is recoverable; that is, when the
load is removed the displacements vanish and the body returns to its unloaded
configuration. Such behaviour is termed elastic deformation. If, however, this load
limit is exceeded, the material undergoes plastic deformation. In this case some
permanent deformation remains after the load is removed. If the deformation
suffered by the material is uniformly distributed in the material volume, such that
all the points in the material volume have the same amount of deformation, the
deformation distribution is termed "homogeneous". It is also possible to have a
deformation distribution where the deformation varies from point to point along
any direction in the material volume. This type of distribution is called
"inhomogeneous" .'
It is well established that, in most of the elastic range, the deformation is
directly proportional to the load. This is known as Hooke's law. The proportionality constant, similar to thermal or electrical conductivity, etc., is a physical
property of the substance under load. If such properties do not change with
direction, i.e., the same displacement is observed for the same load for all testing
directions, the body is said to be isotropic with respect to that particular property.
However this is not always the case. If a property varies with orientation in a given
coordinate system (but is constant at all points in the material volume along a
given orientation), the body is termed to be anisotropic with respect to that
particular property. If, on the other hand, a given property varies along a given
direction in the material volume, the distribution of the particular property is
inhomogeneous.

14

2 Fundamental Concepts in Stress Analysis

2.3 Stress and Strain


The examination of the relationships between loads and displacements in solids
subjected to external loading is greatly facilitated by the definition of two
normalized quantities. These quantities are stress and strain.
Consider the axially loaded bar shown in Fig. 2.1, where the loading forces are
uniformly distributed over the ends. Under the action of these external forces the
material deforms and an internal force field is set up which exactly balances the
external forces when equilibrium is achieved. For a homogeneous bar these
internal forces will also be uniformly distributed ac;ross any cross section nn. The
intensity of this force distribution is the stress acting on nn, which is defined as
0"=

F
A'

(2.1 )

where F is the applied load and A is the cross sectional area of the bar. For the
example showI\.in Fig. 2.1, the stress distribution is homogeneous since the stress
at any point on un is given by Eq. (2.2). However this may not always be the case.
For example, if the load distribution on the external boundary of the bar is not
homogeneous, the stress distribution on nn will be inhomogeneous. Equation
(2.1 ) will yield an average stress value in this case. The actual stress acting on any
infinitesimal element oN on nn must be calculated by defining the actual force of
acting on such an element. It should be noted that the actual stress acting on oN
may be quite different than the average stress given by Eq. (2.1 ).
The second normalized quantity used in load-displacement analysis, strain, is
defined as the deformation per unit length:
~u
L-Lo
e=Y=-L-

(2.2)

Here ~u is the total deformation suffered by the bar and Lo is the initial length of
the bar, as shown in Fig.2.1. If the bar is an isotropic, homogeneous body,
subjected to a homogeneous load distribution at its boundary, the strain at a point
will be given by Eq. (2.2). For other cases, the strain for an infinitesimal element
oL will be different than the average strain given by Eq. (2.2), resulting in an
inhomogeneous strain distribution.
F

n
L
' - - - - - - - ' ..-

6U

Fig.2.1. Axially loaded homogeneous bar of initial


length L

2.4 Forces and Stresses

15

2.4 Forces and Stresses


The loading forces, discussed in the previous example, are vector quantities. A
vector F requires a direction and a magnitude to completely specify it in an
orthogonal coordinate system (Fig. 2.2). Alternatively, the resolved components
of the vector along the three orthogonal axes can be given. In this case the vector is
expressed as
(2.3 )
where F 1,F2 ,F3 are the components ofF along the X1,X 2 ,X 3 directions respectively.
The force component along any axis is defined by a single subscript indicating the
axis. Force is also called a tensor of the first rank. Tensors are physical quantities
which are completely defined with the specification of 3n components, where n is
the rank. Each tensor component is identified by a set of subscripts, where the
number of subscripts is equal to the rank of the tensor. For example, n= 1 for
vectors, and each vector component F j is completely defined by a single subscript, i.
The stresses caused by the loading forces, on the other hand, require the
specification of nine components, six of which are independent in an orthogonal
coordinate system. Consider a unit cube within a homogeneously stressed body at
static equilibrium which has no body force (or torque) distribution within the
body. On all six faces of this cube there will be forces, transmitted by the material
surrounding the cube. Since the body is at static equilibrium, the forces on
mutually parallel faces will be equal in magnitude but opposite in sign. Thus, once
the coordinate system is specified (right-handed or left-handed) only three forces,
acting on the three mutually non-parallel cube faces need be considered (Fig. 2.3).
The force acting on any face can be resolved into three components: two within the
plane they are acting on, and the third normal to this plane. The force component
acting in the normal direction is called the normal force, while the components
acting in the plane are called shear forces, and similarly for the stresses.
Various types ot: notation are in use for the indexing of stresses. Three of the
most used types are shown below:
(2.4 )

In all notations the stresses with mixed indices are the shear stresses. By
convention, the second index defines the normal of the plane in which the stress
acts, and the first index indicates its direction. Thus 0"13 is in the plane defined by
the line Ox 3, and acts in the Xl direction. 0"23 is also in the same plane, but acts
along the X 2 direction.
Normal stresses act along the direction of the axis indicated by the repeated
index. Positive values of stress indicate tensile stresses, while negative values
indicate a compressive stress state.
Thus, stress is a tensor of the second rank and stress components along any
axis require the specification of two indices. It is also seen from Eq. (2.4) that

16

2 Fundamental Concepts in Stress Analysis

F3 ___ _____

F1
Xl

--- ------

',

Fig. 2.2. The vector F and its resolved components F I , F 2 , F3 in an orthogonal


coordinate system

Fig. 2.3. Resolved forces, and the


stresses they cause, on the faces of a
homogeneously loaded unit cube at
static equilibrium
O"jj = O"jj. Such tensors are called symmetric second rank tensors. An elegant proof
that stress is a symmetric second rank tensor is given in Nye's book (see references
at the end of the chapter). However, a simple proof that this must be so for any O"jj'
for the case considered here, can be obtained by taking moments around an axis
passing through the center of the cube, parallel to the Ok axis, where k is the
missing index. For example, consider the case for 0"13' Taking the moments
around the Ox 2 line, (Fig. 2.3) it is seen that the only components contributing to
the moment are the stress components 0"31' 0"13' Since the body is at equilibrium,
and the stress distribution is homogeneous, and remembering that the sides of the
cube are of unit length:

(2.5 )

2.5 Displacements and Strains

17

2.5 Displacements and Strains


The displacements of points in a solid subjected to an external load from their
original positions may be due to rigid body translation and rotation in addition to
deformation. Since strain is produced only by deformation, the other two
components must be subtracted from the total displacement before strain can be
determined.
Consider a line segment in a homogeneous body, subjected to a uniform load
in the elastic range (Fig. 2.4) . In the unloaded position the distance between the
points NN' is .1x. After deformation this distance becomes .1x + .1u, where .1u is
the total displacement between Nand N'. The strain for the line element .1x is
[from Eq. (2.2)]
EAx =

(.1x+.1u) -.1x .1u


.1x
= .1x .

(2.6)

The strain at a point can be found by shrinking .1x to the limit:


e = lim .1u = du
x Ax .... O.1X dx'

(2.7)

Since the deformation is homogeneous, ex is constant for all x. The displacement at


any x is obtained by integrating (2.7):
(2.8 )
where U o is the rigid body translation component.
In this example in one dimension, the line segment .1x elongated in response to
the applied load, but did not change its shape and stayed linear. However in the
more general cases (plane or volume elements in two or three dimensions) there is
usually shape change as well as dimensional change. Deformation leading to shape
change is termed distortion, while dimensional change, without a change in shape,
is termed dilatation:
Consider the rectangular element ABCD (Fig.2.5a) with AB=.1x 1 ,
BC = .1x 2 , which undergoes translation, distortion and dilatation. The point B
F

N
r--I---i
T
~x

-'

Nt --I-----.

- I - N'

6U+6X

N't--I---'-'

Fig. 2.4. Deformation of a line segment in a homogeneously


loaded body

2 Fundamental Concepts in Stress Analysis

18

x,

x,

0'

:h:_,
17
X,

x,

x',

x,
b

Fig. 2.5. a Displacement, deformation and rotation in two dimensions. b Displacement and
rotation without deformation

moves to B' and suffers displacements in both Xl and X2 directions. The


translational displacement can be eliminated by defining the local coordinate
system X'l'X~ which moves with A. Defining the quantities in x;:
(2.9)

which describe the variation of the displacement .::lu l along x~, x~ respectively. The
total displacement at B' along X'l is:
OUl

.::lu l = ~ Lix l
uX l

OUl

+~
.::lx 2
uX
2

(2.10 )

Similarly:
(2.11 )

where
(2.12 )

Thus, eij is also a second rank tensor.


It is possible to define eij for i =+= j geometrically. Consider the triangle A'B'B":
(2.13 )

2.5 Displacements and Strains

19

which, for small displacements, becomes:


(2.14 )
Here e is measured in the direction of rotation of the line segment AB to its new
position (counterclockwise for this case).
It is seen from (2.14) that the eij describe the rotation of AB around A. Such
rotation may contain rigid body rotation as well as distortional components. In
order to obtain the strain associated with the distortioncomponent, the rigid body
rotation must be removed. This may be accomplished as follows:
Assume that a rectangular element ABeD undergoes translation and a rigid
body rotation <I> around A, where <I> is small (Fig. 2.5b ). In the local coordinate
system x;, only the rotation <I> will be observed. The eij tensor for this case is:

'

o -<I

(eij) rot. = wij = { <I>


>

(2.15 )

where wij contains only the rotational components.


Since the displacement ofD to D' is in the negative x~ direction, the rotational
components of eij (wij ) is antisymmetric, i.e. wij = - wji. Thus the quantity
( eij + eji ) /2 will contain only components due to distortion since the rotational
parts will cancel out. The strain tensor is then defined as:

(2.16 )

From (2.16) Eij = eji. Thus strain is also a symmetric second rank tensor.
The strain in three dimensions is also obtained by a similar procedure:

C' '" ' ')

Eij = E12 E22 E23


E13 E23 E33

ell

- (e 12 +e 2d

- (e13 +e 3d

e22

-(e23+ e32)

= -(e 12 +e 2d
2

-(e13+ e3d - (e 23 +e 32 )

(2.17 )

e 33

The strain tensors described by Eq. (2.17) contain both the dilatation and
distortion terms. The normal strains Ejj contribute to the dilatation part,1, which is
defined as
(2.18 )

.113 is called the mean or hydrostatic strain component along each axis.

2 Fundamental Concepts in Stress Analysis

20

The distortion terms in the strain tensor are obtained by subtracting 11/3 from
each Eii :

(2.19 )

where

E;j

are called the deviatoric strains.

2.6. Transformation of Axes and Tensor Notation


In Sect. 2.4 th~ stresses acting on a unit cube within a homogeneously loaded
isotropic material were determined from the components of the forces acting on
the surfaces of this cube. These components, in tum, were obtained by resolving
the forces, transmitted by the material surrounding the cube, in a coordinate
system whose axes were parallel to the three mutually perpendicular sides of the
cube. If a different unit cube, for example cube B shown in Fig. 2.6, is selected, a
new coordinate system is defined. Consequently the resolved forces, and thus the
stresses, acting on the surfaces of cube B will be different than those acting on cube
A. The resolved forces and the stresses acting on the surfaces of both cubes A and B
are related to one another since these are caused by the same physical quantity,
namely the load distribution at the boundary, and the difference arises solely from
the specification of different coordinate systems.
If the forces acting on the surfaces of cube A are known, the forces acting on
the surfaces of cube B can be determined by using the rules applicable to
transformation of axes for vectors. This procedure is examined next.
In Fig. 2.7 the coordinate systems Sand L, defining, respectively, the unit
cubes A and B are shown. Here Xl ,X2,X 3 belong to S and X'l'X~X~ belong to L. The
angular relationships between Sand L are given by the direction cosines between
all the axes Xi and x;:
X2

Xl

X3

Xl COS(X'lX l ) cos (X'lX 2 ) COS(X'lX 3 )


X1 cOS(X~Xl) COS(X~X2) cOS(X~X3)
X3 cOS(X~Xl) COS(X~X2) cOS(X~X3)

(2.20a)

or:
Xl

X2

X3

Xl a 11 a 12 a 13
X1 a 21 a 22 a 23
X3 a 31 a 32 a 33

(2.20b)

21

2.6 Transfonnation of Axes and Tensor Notation

X;\
\

\
\

L71
)Lp"2
x,

(0'.

x,

It

Fig. 2.6. Two arbitrary unit cubes A,


B in a homogeneously stressed body

- ---x;

'/

x,
/

/
,

x,

Fig. 2.7. The "new" an d "old" coordinate


systems L( X'I) and S (x;) and the definition
of the direction cosines between them

Here the first subscript refers to the "new" axis (in L) and the second to the "old"
axis (in S), following Nye.
Let F be a force vector with components F I,F 2,F 3 in the S coordinate system.
The components of F along L,F'I ,F~,F~ may be obtained by resolving each Fj
along X'I ,x~,x~ and then summing up the components for each x;. For example the
new force component along x~ will be:
F~=F 1cos (Xl X'd +F 2COS (x2x'd +F3 cos (x3x'd.

(2.21 a)

The force F; has components from all F j Equation (2.21 a) may be written more
compactly by utilizing Eq. (2.20b):
(2.21b)

F'I =allF l +a 12F 2 +a 13 F 3


Similarly for the force components along
F~ =a 21 F l + a 22 F 2 +a 23 F 3 ,

F~ =a 31 F 1 a32F 2 + a 33 F3 .

x~

and

x~:

(2.21c)

In Eqs. (2.21a,b,c) the leading subscript for each of the direction cosines a jj is the
same as the subscript for the new direction along which the new force component
is being determined. Thus Eqs. (2.21) can be written more compactly as
F; =

L Fj"a jj .

j= 1

(2.22 )

2 Fundamental Concepts in Stress Analysis

22

This equation can be simplified further through the use of the Einstein suffix
notation:
(2.23 )
Here i occurs singly on the left hand side of the equation and denotes the new
direction along which the force component is required. It is termed a free suffix
and is defined by the user of the equation. The index "j", on the other hand, is
repeated on the right hand side of the equation. Such repetition implies summation
of the terms in the equation over all possible values ofj. An index such as "j" in this
example is called a dummy suffix.
The transformation of stresses, caused by the loading forces, from S to L
utilizes the transformation rule for second rank tensors:
(2.24 )
where ij are the free sufflXes and define the stress O";j in S. Since both k and I are
repeated on the right hand side of the equation they are both dummy suffixes and
summation of 3.;k a jl ( over all possible combinations of k and I) is required. For
example, expanding Eq. (2.24) for arbitrary ij:
O"ij= 3.;1aj10" 11 + 3.;1 aj20"12 +ail aj3O"13
+ 3.;2aj1 0" 12 + 3.;2 a j20"22 + ai2a j30"23
+ a i3 aj1 0"13 + a i3 aj20"23 + ai3aj30"33 .

(2.24b)

For each ij pair, one such equation exists. Since stress is a symmetric second rank
tensor, six such equations, each with nine components on the right hand side, are
required for the complete transformation of O"ij in L to O";j in S.
The transformations given by Eqs. (2.23), (2.24) are valid for all first and
second rank tensors, respectively, and are not restricted to the transformations of
forces and stresses. They can be applied, for example, to the electric field vector (a
tensor of the first rank) , or to electrical conductivity (a tensor of second rank) .
For symmetric second rank tensors it is possible to define a unique coordinate
system in which the tensor has no off-diagonal components. That is, a tensor of the
form

dermed in S, becomes

when transformed into the new set of coordinate axes P. The set of axes P, thus
defined, are the principal axes, and the tensor components referred to them are
called the principal tensor components.

23

2.6 Transformation of Axes and Tensor Notation

B
Fig. 2.8. The principal unit cube and an
arbitrary unit cube in a homogeneously
stressed solid

F
Since both stress and strain are symmetric second rank tensors, it is possible to
determine the principal axes associated with them (sometimes called eigenvectors
in the literature). Physically, this means that it is possible to define a unit cube in a
generally stressed material on whose faces only normal stresses (Fig. 2.8), or
normal strains, act. These stresses and strains are also termed principal stresses
and strains.
Principal stresses and the direction of the principal axes can be easily
determined in terms of the stresses existing in an arbitrary cartesian coordinate
system. Consider the homogeneously stressed, tetrahedron-shaped, free body in
Fig. 2.9, which is at static equilibrium. Assume that the normal N to the plane ABC
is a principal direction, along which the principal stress crii acts, and which makes
the direction cosines a, b,c with the coordinate system Xi' ( a 2 + b 2 + c 2 = 1 ), in
which a general stress tensor cr;j is defined. Writing the force balances of the free
body for the X 1,X 2,X3 directions:
( crii - cr ~ l)'a - cr'12b- cr'l3'C

=0

-cr'12a + (crii-cr~2)b-cr~3c=O
- cr ~3'a- cr~3b+. (cr ii - cr~3) 'c=O,

(2.25 )

the Eqs. (2.25) may be written in matrix form


(2.26)

24

2 Fundamental Concepts in Stress Analysis

Fig. 2.9. A tetrahedron-shaped frcc body, onc of whose faces (A Be) is a


principal plane

which is satisfied for a = b = c = 0 (which is a trivial solution), or for


(2.27 )

Expanding the determinant [Eq. (2.27)] one obtains a cubic equation:


(O"ii)3-l r '( O"ii)2+1 2' (O"ii) -13=0.

(2.28 )

Here the coefficients 11,12,1 3 are:


11 =0"~1 +0"~2+0"~3
12 = 0" ~ 10"~2 + 0"~20"~3 + 0"~30"'11 - 0" l2 - O"A - O"l3

(2.29)

13 = 0" ~ 1'0"~2'0"~3 + 20" ~20"~30" ~3 - 0"'11 O"l3 - 0"~20" ~23 - 0"~30"'/2 .

The roots of Eq. (2.28),0"11,0"22,0"33' are the principal stresses. The direction
cosines a,b,c, can then be obtained from Eq. (2.25).
It must be noted that, since the coefficients 11,12,1 3 of the cubic Eq. (2.28)
define the principal stresses, they are independent of the stresses O";j, defined in the
arbitrary coordinate system Xi' These quantities are called the stress invariants.
From the first stress invariant 11 , which is the sum of the leading diagonal of the

2.7 Elastic Stress-Strain Relations for Isotropic Materials

25

stress tensor, it is seen that the sum of the normal components in any coordinate
system is invariant and equal to the sum of the principal stresses.
An equivalent discussion can be also given for the strain tensor. In this case
Eq. (2.28) becomes
( Eii ) 3- I~ ( Eii ) 2 + I~ ( Eii )

I~ =

(2.30)

0,

where
I~ =E~l +E~2 +E~3

(2.31 )

I~ = E~ 1E~2 + E~2E~3 +E~3E'11 -El2 -E'/3 -E~23


I~ = E'l1 E~2E~3 + 2E'12E~3E'13 - E'l1 El3 - E~2El3 - E~3El2 .

The terms I;, defined by Eq. (2.31) are the strain invariants.

2.7 Elastic Stress-Strain Relations for Isotropic Materials


Consider a unit cube within an isotropic material subjected to a homogeneous
tensile elastic stress cr 11 along the x 1 axis. The stress cr 11 will cause a tensile strain
Ell along the Xl axis and compressive strains E22 ,E 33 along the transverse
directions X2'X 3. The transverse strains occur in part because of the cross-bonds
between the atoms. Consider Fig. 2.1 0, and assume that the atomic bonds can be
represented by coil springs. When the stress cr 11 is applied along AB, AB elongates
and suffers a tensile strain. The bonds AC, BC however, pull on atom C, resulting
in a net contraction, and thus a compressive strain along the transverse direction
CC'.
Hooke's law states that these strains are linearly related to the applied stress:
1
E11 ="E cr11

(2.32a)
(2.32b)

-+
A

Fig.2.10. Schematic depiction of transverse contraction in a solid loaded in uniaxial tension

26

2 Fundamental Concepts in Stress Analysis

From Eqs. (2.32a,b)


(2.33)
The proportionality constants E and v are properties of the material being stressed
and are termed Young's modulus of elasticity and Poisson's ratio, respectively.
Since the coordinates are arbitrarily selected, the Eqs. (2.32) may be generalized
for any stress O"jj:

1
Ejj = "O"jj,

v
Ejj = - E O"jj,

i=l=j.

(2.34)

Thus for a unit cube subjected to three normal stresses 0"11 ,0" 22,0" 33' the strain
along any direction Xi is the sum of the strain caused by the stress in that direction,
and the strains caused by the stresses acting in the transverse directions:

1
v
E11 =EO"l1-E (0"22+0"33)
1 ..
V
E22= E 0"22-" (0"11 +0"33)

(2.35)

1
v
E33 ="0"33-E (0"11 +0"22).
From Eqs. (2.17) and (2.35) it is seen that for isotropic materials, normal stresses
O"jj cause only dilatation. Pure distortion (shape change) occurs when a shear
stress O"ij' i =1= j, is applied to the material. The shear strain in this case is given by
1
2Eij =Y= ~O"ij,

i=l=j.

(2.36)

Here the quantity y = 2Eij is called the "engineering shear strain" and the
proportionality constant "11" is called the shear modulus. For isotropic materials 11
is related to E and v by 211 = Ej ( 1 + v ) .
Equations >( 2.35) and (2.36) can be expressed compactly in the suffix
notation as
(2.37 )
where k is the dummy suffix and implies summation for all k, and ()ij is
Kronecker's delta, which is defined as:
1, i=j
()ij = { 0 . =1=
,
1 J.

(2.38)

For example, for i=j=l, Eq. (2.38) becomes

v
1 +v
Ell = 0"11-" (0"11 +0"22+0"33)
which is identical to Eq. (2.35).

(2.39a)

2.7 Elastic Stress-Strain Relations for Isotropic Materials

27

Table 2.1. Young's modulus and Poisson's ratio for some materials (at room temperature)
Material

E(GPa)

Aluminum
Copper
Gold
Iron
Molybdenum
Tungsten
Titanium Carbide
Lead

70.3
129.8
78.0
211.4
320.0
411.0
436.9
24.5

0.345
0.343
0.44
0.293
0.293
0.280
0.199
0.410

Data adapted from R.W. Hertzberg, Deformation and Fracture Mechanics of Engineering
Materials, Wiley, New York, 1976, p.8, and G. Simmons and Herbert Wang, Single Crystal
Elastic Constants and Calculated Aggregate Properties, MIT Press, Massachusetts, 1971

IfEq. (2.37) is inverted, stresses are expressed in terms of strains. Summing up


Eq. (2.37), for i=j=1,3:
E

0"11+0"22+0"33= (1-2v) (1':11+1':22+1':33)

(2.39b)

and substituting for any two stresses O"jj'O"Il in (2.39b) (i.e. the stresses 0"22,0"33'
expressed in terms of strains I':ij from equations similar to (2.39a), would be
substituted in (2.32b) if we wanted to determine 0"11 ) :
E
vE
O"ij = (1 + v) I':ij + (1 + v) ( 1 _ 2v) oihk

(2.40)

The term vE/[ (1 +v) (1- 2v)] is usually represented by the symbol A and is
called Lame's constant.
From Eq. (2.18) and Eqs. (2.35), (2.39) it can be seen thatthe dilatation A is
linearly related to the hydrostatic component of stress that causes it:
A=

(1-2v)
1
E
.30"Hyd. = K O"Hyd.

(2.41 )

where O"Hyd. = ( 1/3) O"kk and K is called bulk modulus. It can be seen by inspection
that for isotropic materials, K = E/3 ( 1 - 2v ) .
From an examination of the above equations, it can be seen that the stressstrain relations for an isotropic solid in the elastic range require the specification of
only two constants E and v. Typical values of these constants for some important
materials are given in Table 2.1.
Up to this point elastic behaviour in homogeneous isotropic solids has been
discussed. Single crystals, or the grains in a polycrystalline material, are
homogeneous (in a macroscopic sense) but anisotropic solids. Thus, the simple
equations between stress and strain given in this section are no longer applicable
and a more general treatment that takes into account the particular structure of
single crystals is required.

28

2 Fundamental Concepts in Stress Analysis

2.8 Structure of Single Crystals


Single crystals are materials that exhibit long-range order (periodicity) in the
position and stacking sequence of their constituent atoms throughout their total
volume. This long range order may be visualised as a three dimensional
arrangement of (lattice) points in space, where each lattice point has identical
surroundings. Associated with each lattice point is an atom or group of atoms,
depending on the solid under consideration. Thus, once a lattice and the atom
arrangement associated with a point in the lattice is defined, the atom arrange-

!11J
1

acto
1

10

II

-.-

o '.

12

13

14

Fig. 2.11. Unit cells of the 14 Bravais lattices: 1, Simple triclinic; 2, simple monoclinic; 3, basecentered monoclinic; 4, simple orthorhombic; 5, base-centered orthorhombic; 6, body-centered
orthorhombic; 7, face-centered orthorhombic; 8, hexagonal; 9, trigonal (rhombohedral);
10, simple tetragonal; 11, body-centered tetragonal; 12, simple cubic; 13, body-centered cubic
( BCC); 14, face-centered cubic (FCC)

2.8 Structure of Single Crystals

29

Table 2.2. Definitions of the seven different crystal systems


System

Axes and interaxial angles

Triclinic

Three axes not at right angles, of unequal lengths.


a 1 *a2 *a 3 , cx* ~*y*900.

Monoclinic

Three axes, one pair not at right angles, all of unequal lengths.
a 1 *a2 *a3 , cx* ~*y=900

Orthorhombic

Three axes at right angles, all unequal in length.


a 1 *a2 *a 3 , cx= ~=y=900

Tetragonal

Three axes at right angles, two of equal length.


a 1 =a 2 *a 3 , cx= ~=y=900.

Hexagonal

Two axes of equal length at 120, the third axis perpendicular to


these.
a 1 =a 2 *c, ~=y=90. cx=1200.

Trigonal

Three axes of equal length, all equally inclined but not mutually
perpendicular.
a 1 =a2 =a 3 . cx= ~=y*900.

Cubic

Three axes at right angles, all equal in length.


a 1 =a 2 =a 3 , cx=~=y=900.

ment at any point in the crystal is uniquely defined by the symmetry associated
with this particular lattice. There are only fourteen different ways of arranging
points in space such that each has identical surroundings. These point lattice
arrangements are called the Bravais lattices. They are described simply by a unit
cell which reflects their symmetry. This unit cell can be repeated periodically (like
stacking bricks in a wall) to form the complete lattice. In Fig. 2.11 and Table 2.2
. the fourteen point lattices and the seven crystal unit cells that describe them are
shown. Each of these cells has several possible arrangements of atoms around the
points consistent with periodicity, (but because of periodicity the number of such
arrangements is limited), resulting in 230 "space groups", or possible patterns for
atomic structures: We say patterns because the dimensions and actual distances
between the atoms vary for the same pattern, depending on the material. (For a
complete discussion of symmetry and point groups, the reader is referred to the
standard texts on the subject. See, for example, "Diffraction from Materials" by
Schwartz and Cohen. )
The position of an atom within the unit cell is given by its fractional coordinates
u,v,w, where u,v,w are defined by the vector r=ua 1 +va 2 wa 3 , that connects the
atom to the origin ofthe unit cell. For example an atom at the body center of a unit
cell, referred to an orthogonal set of axes has the coordinates ~,~, ~, while the one
at the origin is located at 000.
The direction of any line in a unit cell is described by drawing a line through the
origin parallel to the given line and then expressing the coordinates of this line as
the smallest set of integers that is a multiple of these coordinates. Directions are

2 Fundamental Concepts in Stress Analysis

30

[001]

,A

[101]

[021]

[010]

[100]

Fig. 2.12. Va rio us c rystallographic directio ns in a


cubic unit cell

usually expressed in square brackets [hkl]; this defines a family of parallel


directions. In Fig. 2.12 various directions in a cubic lattice are shown. In this figure
it can be seen that the directions [100], [010], [001] are equivalent directions which
are rela\ed to each other by symmetry, (for example [100] becomes [010] if the
cube is rotated 90 around the AA' axis) . Such equivalent directions belong to a
form of directions which is denoted by carats (hkl). [100], [010], [001], as well as
the directions [100], [010], [001] (which have negative intercepts denoted by
"-" along x 1 ,X Z,X3 respectively), belong to the (100) form of directions.
The orientations of planes of atoms are expressed through the use of Miller
indices, which are the reciprocals of the fractional intercepts the plane makes with
the unit cube axes. For example, the Miller indices of a plane intersecting the unit
[001]

[100]

Fig. 2.13. a Various crystallographic planes in a cubic unit cell. b Various crystallographic planes
and directions in a hexagonal unit cell

2.8 Structure of Single Crystals

31

cube axes at l/h,1/k,1/1 are (hkl). Alternately, the intercepts of the plane (211) is
0.5 along Xl and 1 along X 2 and x3 . If a plane is parallel to one of the axes, its intercept is at infinity and the Miller index for that axis is o. In Fig. 2.13a various planes in
a cubic lattice are illustrated. Using a reasoning similar to the one given for the
equivalent directions, it can be seen tha the planes (100), (010), (001), (TOO),
(010), (001) are equivalent planes related by symmetry. Such planes are called
planes of a form and are denoted by {hkl}. In the case discussed above the planes
belong to the {100} form.
For hexagonal systems a slightly different system of plane indexing is used
instead of the one discussed above. In this scheme a third axis a 3 , which makes
angles of 1200 with the basal vectors a 1 ,a 2 , is defineq in the basal plane of the
hexagonal unit cell. The Miller indices of any plane are then determined by
obtaining the inverse of the fractional intercepts of the plane on the three basal
axes a1>a 2,a 3 and the vertical axis c. Thus four numbers [hkil] are specified as the
plane indices. Since the intercepts on a 1,a2 uniquely determine the intercept on a 3,
the index i is a function of the indices hand k, where i is given by
(2.42 )

h+k= -i.

In Fig. 2.13b, various planes and directions in a hexagonal unit cell are shown. It
must be noted that, even though the plane indices are referred to four indices as
described above, the lattice directions are referred to the three crystallographic
axes a 1 ,a 2 ,c, in keeping with the convention for all crystal systems.
A single crystal of finite dimensions contains a very large number of unit cells,
each with its own set of {hkl} planes. For an unstressed single crystal, the family of
(hkl) planes throughout the volume are parallel (due to symmetry) , and the plane
spacing between them is constant. This spacing oflattice planes is a function of the
indices (hkl) of the plane and the unit cell constants a 1,a2,a 3,Cl,(3;y. A summary of
these relationships for some unit cells is given in Table 2.3. In general, all planes of
a family have the same lattice spacing.
Table 2.3. Formulas relating the plane spacing, dhkb to the indices h, k, I, and unit cell parameters
for various unit cells '

Cubic
Tetragonal
Hexagonal

Rhombohedral
Orthorhombic

d2

(h 2 + k 2 + 12) sin 2 cr+ 2(hk + kl + hI) (cos 2 cr-cos cr)


a 2 (1-3 cos 2 cr+ 2 cos 3 cr)

1 h 2 k 2 J2
-=-+-+d2 a~ a~ a~

2 Fundamental Concepts in Stress Analysis

32

2.9 Elastic Stress-Strain Relations in Single Crystals


In a crystal, any atom within the unit cell occupies a position where its potential
energy, due to the attractive and repulsive forces from the surrounding atoms, is a
minimum. This is shown schematically in the potential energy vs. nearest neighbor
plot (Fig. 2.14) where, at equilibrium, the atom occupies the bottom of the
potential energy well, separated from its neighbours by the equilibrium distance
roo If an external force is applied to the crystal, the atoms will suffer a displacement
u to a new interatomic spacing r at higher potential energy, where the sum of the
applied and interatomic forces is zero. For small displacements the potential
energy can be expressed as a Taylor series around ro:
v(u)=vro +
where

Vro

(dV)
dr

ro'

1(ddrv)

u +2

ro'

u 2 + ... ,

(2.43a)

is the potential energy at ro and all derivatives are evaluated at roo

Noting that

(.~:) ro = 0 and neglecting higher order terms:


(2.43b)

The applied force is the derivative of the potential energy with respect to
displacemen t:
F=

ov (u) = (02V )
OU

or2

.U
ro

'

( 2.44a)

or
F=K-u
where K,

(2.44b)

(~:~ ) ro' is the curvature at the bottom of the potential well.

Equation (2.44b) is the original form of Hooke's law. If, instead of forces and
displacements, it is expressed in terms of stresses and strains, the proportionality
constant is direc;tly proportional to the elastic modulus of the material.
From the above discussion, the elastic modulus will be the same for both
tension and compression since K is independent of the sign ofu. As it is dependent
only on the variation of the interatomic forces with interatomic distance about the
equilibrium position, the modulus is a basic material property and is not
significantly affected by processes such as heat treatment or plastic deformation
for most common engineering materials. Furthermore, since along various
directions in the point lattices the spacing between the atoms is different, the
variation of potential energy with distance, and hence the elastic modulus, will
change with direction, causing anisotropy. However it must be noted that, because
a crystal is built up by repeating the unit cell periodically, the elastic modulus
along a given direction will be constant through the crystal volume.

2.9 Elastic Stress-Strain Relations in Single Crystals

>

33

r-

Fig. 2.14. The potential well for an atom at equilibrium in a solid

Hooke's law for such anisotropic materials is of the form

(2.45 )

Eij = SjjklO"kl
or, inverting the equation and expressing stresses in terms of strains:

(2.46)
Here Sjjkl,C jjkl are the compliance and stiffness moduli of the crystal under
consideration, and are both tensors of the fourth rank.
Equation (2.45) means that if a general homogeneous stress is applied to a
single crystal, the resulting homogeneous strain is linearly related to each
component of the stress tensor through the appropriate component of the
compliance tensor. For example the strain in the Xl direction, Ell' is given by:
Ell = Sl1l1 0" 11 + Sl112 O" 12 + Sl113O" 13 +S1121 0" 12 + Sl1220"22

(2.47)

+ S11230"23:+ Sl131 0" 13 + Sl132 0"23 + S11330" 33 .

It is seen that, for a general anisotropic crystal, in contrast to an isotropic solid,


the shear stresses, as well as the normal stresses, contribute to the strain in a
normal direction and vice versa.
For the most general case, Eqs. (2.45) and (2.46) describe nine linear
equations, each with nine terms. Thus, there are 81 constants in the the Sjjkl' and
C jjkl tensors. It can be shown that (Nye) for both of these tensors C jjkl = C jj1k and
Sjjkl = Sjjlk' Similarly Cjjkl = Cjjkl and Sjjkl = Sjjkl' These relationships bring the
number of distinct constants to 36 for both tensors.
The elastic stiffness and compliance terms are usually expressed with two
subscripts instead offour. This notation is called the contracted notation or matrix
notation and it is obtained by contracting the first pair of suffixes and the last pair
of suffixes into a single number, each according to the following set of rules:

tensor notation
matrix notation

11

22

33

23,32

31,13

12,21

2 Fundamental Concepts in Stress Analysis

34

In addition, for the Sjjkl the terms

~ and ~ are introduced as follows:

when m and n are 1,2, or 3,


when either m and n are 4, 5, or 6,
when both m and n are 4, 5, or 6.
These terms are not required for the C jjkl.
For example, (2.47) written in matrix notation becomes
1

ell =Sl1O"l1 + S12O"22+ S 13 0"33+ 2(S140"23+S150"13+S160"12).

(2.48 )

The matrices Sjj and Cjj are 6 x 6 matrices and contain 36 terms each. However not
all of these terms are finite and independent. It can be shown, from consideration
of the strain eJlergy of a crystal, that for any crystal Sij = Sji" Similarly C jj = Cjj.
Thus the number of independent constants is reduced to 21 (for i = j, there are six
terms, and for the remaining 30, only one half is independent, for a total of 21 )
for a general anisotropic crystal. A particular crystal symmetry or atomic
arrangement can further reduce the number of independent constants (see Nye
for a complete discussion). For example, for cubic materials there are only three
independent constants. The compliance matrix in this case becomes

Sjj=

Sl1
S12
S12
0
0
0

S12
Sl1
S12
0
0
0

S12
S12
Sl1
0
0
0

0
0
0
S44 0
0
S44
0
0

0
0
0

0
0
0
0
0
S44

(2.49)

It should be noted that, even though there are three independent constants for
cubic crystals, the compliance matrix contains nine non-zero components. In
Table 2.4 the number of independent constants and total number of non-zero
components for each of the seven crystal unit cells is given. It can be seen that as
the symmetry o(the crystal increases, the number of independent constants in the
stiffness and compliance matrices decrease. Typical values of stiffnesses and
compliances for various crystals are shown in Table 2.5.
The stiffnesses and compliances given in Table 2.5 are defined in the unit cube
axes, (100), which are also termed crystal axes. The compliance (or stiffness)
tensor in any otqer coordinate system can be calculated from the transformation
law for fouith rank tensors:

(2.50)
where Sjjkl is the elastic compliance in the new coordinate system, ajj are the
direction cosines between this system and the crystal axes, and Smnop are the

2.9 Elastic Stress-Strain Relations in Single Crystals

35

Table 2.4. Number of independent constants and number of non-zero elements in the stiffness and
compliance matrices of the seven unit cells
Crystal system

Independent constants

No. of non-zero elements

Triclinic
Monoclinic
Orthorhombic
Tetragonal
Trigonal
Hexagonal
Cubic
Isotropic

21
13
9
7'

21
13
9

11'

7-

1499

3
2

, For certain symmetries this lattice has 6 independent constants and 12 finite elements
- For certain symmetries this lattice has 6 independent constants and 9 finite elements. See Nye
for details

Table 2.5. Elastic stiffness and compliance values for selected materials
Material
(cubic)

C ll

C 12

C 44

10.82
16.84
18.60
23.70
46.00
50.1
27.9
51.3

Material
( hexagonal)

C ll

Titanium
Zinc

16.0
16.1

C I2

6.13
12.14
15.70
14.10
17.6
19.8
15.3
10.6
C 13

C 33

2.85
7.54
4.20
11.60
11.0
15.14
15.3
17.8

1.57
1.50
2.33
0.80
0.28
0.26
0.585
0.21

C 44

Sll

(x10 Io Pa)
9.0
3.42

S44

(x 10- 11 Pa -I)

(x10 10 Pa)
Aluminum
Copper
Gold
Iron
Molybdenum
Tungsten
Spinel
Titanium carbide

SI2

SII

-0.57
-0.63
-1.07
-0.28
-0.08
-0.07
-0.208
-0.036
SI2

SI3

3.51
1.33
2.38
0.86
0.91
0.66
0.654
0.561
S33

S44

(x 10- 11 Pa -1)
6.6
5.01

18.1
6.1

4.65
3.83

.97 -.47
.84
.05

-.18 .69
-.73 2.84

2.15
2.61

Data adapted from R. W. Hertzberg, Deformation and Fracture Mechanics of Engineering


Materials, Wiley, New York, 1976, p. 13f

compliances defined, in the crystal axes. Equation (2.50) can also be used to
determine Young's modulus for any arbitrary direction: Assume that a uniaxial
stress 0' is applied to a cubic single crystal along the direction [hkIJ. The strain
along [hklJ is linearly related to the stress through Hooke's law:
E=

1
-(E) .0'.
hkl

( 2.51)

36

2 Fundamental Concepts in Stress Analysis

Table 2.6. Relative anisotropy of various cubic materials


Material
Aluminum
Copper
Gold
Iron
Molybdenum
Tungsten
Titanium Carbide
Spinel

1.22
3.20
2.86
2.51
0.79
1.00
0.877
2.43

E(100)

E(lIl)

(GPa)

(GPa)

63.7
66.7
42.9
125.0
357.1
384.6
476.2
170.0

76.1
191.1
116.7
272.7
291.6
384.6
429.2
364.5

However, since the material is anisotropic, Eq. (2.51) can also be written through
the general form of Hooke's law. Arbitrarily defining [hkl] as the x~ of a
coordinate system,
(2.52 )
where S' 1111 are referred to the x; coordinate system. From Eq. (2.50), (2,51) and
( 2.52 ), Young's modulus along [hkl] can be written in terms of elastic
compliances S;jkl defined in the crystal axes:
1
.
-(E) =Slill =almalnalOalpSmnop'
hkl

(2.53)

Here alj are the direction cosines between the crystal axes and the direction [hkl].
Expanding Eq. (2.53), and noting that the terms almalnalOalpsmnop indicate
summation over all mnop, with m,n,o,p = 1,3 (i.e., all nine components of the
compliance tensor (2.49) will contribute to the summation), one obtains:

1S44 ](alla12+aI2a13+alla13)
22
(E1).
hkl =Sllll=Sll-2 [(Sll-S12)-2
22

22

(2.54 )
Equation (2.54) shows that Young's modulus in cubic crystals is anisotropic,
and is a function of the direction cosine term (aIlaI2+aI2aI3+aIlaI3)' This
term varies between 0 for the (100) family of directions, and

for the (111)

family. Thus the direction with the maximum Young's modulus depends on the
term Sll - 2 [( Sll - S12) -

~ S44

J.

If this term is positive, Young's modulus

exhibits a maximum along (111) and a minimum along (100). Ifit is negative, as
it is for Mo, the maximum is along (100), and the minimum along (111). If this
term is zero, Young's modulus is isotropic. That is, it is independent of direction.
Thus, for isotropic materials Sl1 -S12 =

~ S44' It can also be seen that the larger

(Sl1-S12)- ~ S44' the greater the anisotropy in Young's modulus. This quantity

2.10 Equations of Equilibrium

37

is frequently used as a measure of anisotropy. In Table 2.6, 2 ( S 11 - S 12 ) jS44' and


E Ill , E IOO are given for various (cubic) materials.

2.10 Equations of Equilibrium


In the discussions so far it has been assumed that the material under load is at
static equilibrium. This condition imposes certain restrictions on the stress
distributions possible within the material. These restrictions are described by the
equations of equilibrium.
Consider a unit cube with sides OX I ,OX 2,OX 3 within a body subjected to an
inhomogeneous stress distribution (Fig. 2.15). The total force acting on the
element in the Xl direction (neglecting body forces such as weight, and assuming
no body torques exist in the material volume) is
[( CT ll ) A - (CT ll hJ OX20X3 + [( CT 12 h

(CT 12 hJ OX I OX 3

+ [(CT 13 )C- ( CT13)DJOX I OX 2 =0.

(2.55)

Dividing both sides by OV = oX I ox2ox 3, and taking the limit as oV -+0;


OCT l l
oX I

+ OCT12 + OCT 13
oX 2

OX3

=0.

(2.56a)

Similarly, for the X2,X 3 directions:

OCT12
oX I

+ OCT 22 + OCT 23

=0

OCT l3
oX I

+ OCT 23 + OCT 33

=0

oX 2

OX2

oX 3
oX 3

(2.56b)

(2.56c)

Fig. 2.15. An infinitesimal cube subjected to an


inhomogeneous stress distribution

2 Fundamental Concepts in Stress Analysis

38

These equations can be written in abbreviated form as

j=!

(2.57 )

OO'ij =0,

oXj

which, in suffix notation, becomes


(2.58 )

O'ij,j=O.

Here j is the dummy suffix implying summation, and the comma before j indicates
the derivative of each term with respect to xj .
It must be noted that Eq. (2.58) describes the variation of stresses within the
body volume and must be satisfied at every point within the body if the body is at
static equilibrium. If the stress distribution is homogeneous, Eq. (2.58) is
automatically satisfied since each derivative with respect to distance is identically
equal to zero.
If the volume element has a surface boundary, any surface forces will also
contribute to the force balance. A similar treatment for this case yields
(2.59 )
Here OJ is the unit vector in the Xj direction on the boundary and Fi is the applied
force in the Xi direction. If the surface has no applied tractions, i.e., it is a free
surface, Eq. (2.59) becomes
(2.60)
Equations (2.59) , (2.60) describe the boundary conditions for the stress
distribution existing within the material required for equilibrium.

2.11 Conditions of Compatibility


In Sect. 2.5 it was shown that the components of the displacement vector at a point
are related to the components of the strain tensor at that point through the
relations
1

(2.61 )

t ij ="2 (Ui,j+Uj,i) ,

where (2.61) describes a set of six equations.


It is possible to eliminate the displacement terms Ui in these equations and
obtain a set of equations relating the components of the strain tensor to each other.
For example for the components t l1 ,t 12 ,t 22 :
02t l1

03 U1

ox~ = oX 1ox~ ,

02t 22

03 U2

oxi

ox 20xi '

02t 12

03 U!

ox! oX 2 = ox! ox~

03 U2

+ ox 2 0xi

thus:
(2.62 )

2.12 Basic Definitions in Plastic Deformation

39

The relationship between the terms ell,e13,e23,e12 may be determined in a similar


manner:
(Yell

OX Zox 3 =

1 0 {

2 oX I

Oe23

- oX I

Oe13

oX z

Oe12 }

oX 3

(2.63 )

Four other equations similar to Eqs. (2.62) and (2.63) may be obtained through
the cyclic permutation of the suffixes in these equations. Written in the suffix
notation the compatibility conditions are expressed as
epkjeqljeij,kl = 0 ,

(2.64 )

where epkj is the permutation tensor. It is defined as .


for even permutation of pki,
for off permutation of pki,
for all other combinations.
For example, eI23~1, e132=e321 = -1, e112=e313=0.
The compatibility conditions restrict the types of displacement functions that
may exist in a material for a given type of loading. For a homogeneous elastic
stress distribution, for example, the displacement components U j must be
continuous, single valued functions of the coordinates Xj.

2.12 Basic Definitions in Plastic Deformation


Materials suffer plastic, or permanent, deformation when subjected to loads in
excess of the elastic limit. The plastic strains caused by such deformation are
generally much larger than elastic strains and there is no simple linear relation
between stresses and strains in the plastic regime similar to Hooke's law, which is
only valid in the range of elastic loading.
The flow curve, which is the stress-strain plot in tension over the entire loading
range, is generally used as a measure of the plastic deformation characteristics of
materials. In contrast to the elastic region, where for all metals strain varies
linearly with stress, with a slope dependent on the particular material, the shapes
of the stress-strain curves are different for various groups of materials. In
Fig. 2.16a - c the stress-strain plots for grey cast iron (a), mild steel (b), and pure
aluminum (c) are depicted schematically. The plot for grey cast iron, which is
representative of brittle materials, exhibits no plastic deformation and obeys
Hooke's law until fracture. The other two curves are from ductile materials which
deform plastically. The curve for mild steel exibits a sharp "yield point" which
separates the elastic and plastic deformation regions. Aluminum, on the other
hand, undergoes a gradual transition from elastic to plastic deformation. Most of
the engineering materials with BCC structures behave in a fashion similar to mild
steel, while metals with FCC or HCP structures exhibit flow curves similar to
aluminum.

40

2 Fundamental Concepts in Stress Analysis

Fig. 2.16. Schematic flow curves for (a)


grey cast iron, (b) mild steel and (c)
pure aluminum

The actual flow curves for ductile materials shown above are hard to model
mathematically and can not be incorporated easily in plasticity theories. Instead,
idealized flow curves which simplify the mathematics without deviating too much
from physicaf'reality, are usually used. The flow curve for rigid, perfectly plastic
material, which does not deform elastically but starts to flow plastically at a
constant stress, is shown in Fig.2.17a. Another variation of this model is the
perfectly plastic material with an initial elastic region (Fig. 2.17b ). Strain
hardening (increase in stress with increasing plastic deformation) materials are
often modeled by the piecewise linear flow curve shown in Fig.2.17c.
The stress and strain plotted in all of the flow curves shown above are the "true
stress" and "true strain". Due to the large deformations usually encountered in
plasticity these values are defined with respect to the instantaneous specimen
dimensions rather than the original specimen dimensions, as shown in
Eq. (2.1 ), (2.2) for the elastic loading regime.
True strain is defined as the sum of incremental strains over the total
deformation length:
E=

l:

j=l

Lj-L(i-l) = S dL =ln~.
Lj
Lo L
Lo

For small deformations, i.e., (L - Lo)


by Eq. (2.2).

~ 1,

(2.65 )
true strain is equal to the strain given

c
Fig. 2.17 a-c. Various idealized flow curves. a Perfectly plastic material; b elasto-plastic material;
c elastic and linearly work-hardening plastic material

2.13 Plastic Deformation of Single Crystals

41

Fig. 2.1S. Unloading of stress after plastic deformation


to a total plastic strain of &2

True stress at any time t is defined as the instantaneous load intensity which is
cr=F/A.

(2.66 )

Here F is the load and A is the actual load bearing area at time t.
It must be noted that the true strain at any time t contains plastic and elastic
components. For example, upon unloading from a point A (Fig. 2.18 ) on the flow
curve, with true strl;lin E 3 , the material unloads along a line (AA'), parallel to the
elastic region of the flow curve, to residual plastic strain E 2 The elastic strain
component at A is thus Ee = E3 - E2

2.13 Plastic Deformation of Single Crystals


Plastic deformation in single crystals occurs primarily by the process of slip. In this
process one part of the crystal slides an interatomic distance relative to the other.
In Fig. 2.19 this process is shown where an entire atomic plane has slipped over the
one below it. The stress required for this translation is extremely high however,
since all the atoms on the slip plane are displaced from their equilibrium positions
simultaneously and a large number of bonds must be broken all at once. Such
movement in single crystals generally occurs due to dislocation motion
(Fig. 2.20), as the stress required. for this process, which involves successive
displacements of single atoms on the slip plane, is much lower. In either case,
however, the symmetry of the crystal is preserved after deformation. Due to this
condition the periodic arrangement of atoms strongly influences slip. The relative
motion of one part with respect to the other usually takes place on definite
crystallographic planes along definite directions. Often, the slip plane is the
plane of maximum atomic density. Such planes are the most widely separated
planes in the crystal and offer less resistance to slip. The slip directions are the
closest packed directions within the slip plane. These directions have the shortest
path between equilibrium atom positions and thus represent the shortest
translation step, after which the atoms are back in register and the original
symmetry within the crystal is re-established.

-. .

42

.
....

A-

2 Fundamental Concepts in Stress Analysis

.. .

Fig. 2.19 a,b. Schematic depiction of slip in a


perfect lattice

The slip plane and the slip direction together define the slip system. Depending
on the symmetry of the particular crystal under consideration, there may be a
number of equivalent slip systems. For example in FCC crystals the {111} planes
are the most densely populated planes, with the (110) directions the closest
packed directions in these planes. Since there are four sets of {Ill} planes in a unit
cell, each containing three (110) directions, there are 12 equivalent slip systems in
the FCC lattice. In the HCP lattice, On the other hand, the only high density plane
is the basal plane {0001} and the close packed directions within the basal plane are
the three (110) directions. Thus there are only three equivalent slip systems for
this lattice, unless slip occurs on other planes (as it sometimes does).
The slip system that will be activated during extemalloading depends on the
resolved shear stress on the equivalent slip systems. Slip begins when the shearing
stress on a particular slip system exceeds a critical value (Schmid's law) which is
called the critical resolved shear stress. For example, consider Fig. 2.21, where a
normal tensile load F is applied to a single crystal of cross sectional area Ao normal
to the load. The cross-sectional area A of a slip plane which is inclined at angle <I> to
the tensile axis is
(2.67)

A=Ao/cos <1>.
(J --

-.l..L

Slip Plane
------------ - - - - - -




---


(J--

IJ

Fig. 2.20. Schematic depiction of slip by movement of an edge dislocation, denoted by 1-, to the
material boundary

2.13 Plastic Deformation of Single Crystals

43

Slip Plane
Normal

Slip Direct ion

Fig. 2.21. Relative orientations of the loading direction, slip plane and slip direction in a sample loaded in
uniaxial tension

If the slip direction in this plane makes an angle A with the load, the resolved load
for this direction is

P=POCOSA.

(2.68 )

Thus the resolved shear stress on the slip plane is


cr =

(~: ) 'cos 'iP'cos A.

(2.69 )

The geometric term cos <I>'cos Ais called the Schmid factor and is a maximum when
<I>=A=45. When <I> or A=900, there is no resolved shear stress on the slip plane
and thus no slip will take place on this particular plane.
Once plastic deformation by slip starts, it can proceed in two ways. The crystal
sections bounded by the slip planes may simply slide over each other without
changing their orientation with respect to the load axis (Fig. 2.22a ) . In a normal
tensile test however, such lateral movement is constrained by the grips applying
the load, and the slip planes rotate to accomodate these constraints, (a similar
process occurs also in a macroscopically homogeneous body subjected to complex
loading when its movement is restrained) . The planes near the grips also undergo
bending in addition to rotation (Fig. 2.22b). In both cases however, the
~

Fig. 2.22 a,b. Plastic flow in single


crystals in uniaxial loading. a Without end constraints; b with end
constraints

Twin plane

Fig. 2.23. Schematic depiction of


atom movement during twinning
for a mirror twin

44

2 Fundamental Concepts in Stress Analysis

deformation distribution is inhomogeneous, with plastic deformation concentrated in certain regions in the crysta1.
Another process by which crystal deformation can take place is twinning. In
twinning, one section of the crystal deforms in a manner that brings it into mirror
or rotational symmetry with the rest of the crystal lattice (Fig. 2.23 ). Twinning
involves more types of atomic planes, compared to slip, and usually causes small
deformations. However, after twinning, a set of slip planes may be oriented
favorably with respect to the applied load, thus twinning may facilitate slip. The
plastic deformation distribution caused by twinning is also inhomogeneous.

2.14 Deformation and Yielding in Inhomogeneous Materials


In the discussions so far, the solid undergoing loading was treated as a
homogeneous continuum. This implies a homogeneous deformation distribution
in the materialwith plastic yield starting in all regions of the body at the same time.
These assumptions are valid, on a macro scale, for engineering materials with a
small grain size undergoing relatively small strains. On a microscopic scale (with
distances of the order of a few grain diameters), however, such materials are
inhomogeneous. That is to say, if such a material is loaded at the boundary with a
uniform stress (Jjj, the stress or strain at a point in the material is affected by the
distribution of the material properties around the particular point. Similarly, the
applied stress may exceed the yield point in certain grains, whereas the rest of the
material may still be in the elastic range, causing an inhomogeneous distribution
of plastic strains. Such inhomogeneous behaviour causes some problems in the
definitions of the "elastic limit" for real materials. Furthermore, the more sensitive
strain measurements become, the lower becomes the stress at which some plastic
deformation is found. These observations have caused the utilization of a number
of parameters describing the onset of yielding:
The "proportional limit" is the stress below which Hooke's law is obeyed by
the bulk of the material; i.e., the stress is proportional to the strain with the strain
measured to a sensitivity of 10- 4 mm/mm. This parameter can be obtained from
the stress/strain plot for a given material by observing the deviation of the plot
from linearity. This parameter is usually equal to or slightly lower than the load at
which irreversible plastic deformation starts in the bulk. It is possible, however, in
some cases such as high strength metal whiskers, to have non-linear proportionality between stress and strain at relatively large strains where the material is still
elastic.
The "elastic limit" is defined as the greatest stress a material can withstand
without exhibiting macroscopic yielding. The sensitivity of such measurements is
usually in the range of 10 -4 mm/mm. In polycrystalline materials, some localized
plastic deformation will occur below the "elastic limit" and, thus, in some cases a
more sensitive parameter is used.
The quantities "anelastic limit" and "true elastic limit (also called the precise
elastic limit)" are based on LVDT (linear variable differential transformer)
measurements of strain on the order of 10- 6 mm/mm. These values are very low

2.14 Deformation and Yielding in Inhomogeneous Materials


III
III

45

Vi
~

ClI

.c

lfl

r.

Fig. 2.24. Determination of the "true yield point ('to) "and


the anelastic limit ('t A ) during strain measurement with an
LVDT
Shear strain -

and reflect the movement of a very small number of dislocations. The true elastic
limit is associated with the movement of a few hundred dislocations. Above this
value, a load-unload cycle produces a completely closed, parallelogram shaped
mechanical hysteresis loop (OABC in Fig. 2.24 where AB is arbitrarily small) at
the strain levels involved. The loop closes on itself (OADE) until the applied
stress reaches the anelastic limit ('tA)' Any further increase in stress results in an
open hysteresis loop, with some micro plastic deformation distributed in the
material. These values are especially important in applications like the manufacture of inertial navigation systems, etc., where no plastic deformation can be
tolerated.
A parameter more commonly used in engineering is the "yield strength" which
is the stress required to produce a specified amount of plastic deformation. The
amount of plastic deformation ( or offset) is usually specified at 0.002 inch/inch or
0.2 % strain) and the yield stress is taken as the stress at which a parallel drawn to
the elastic portion of the stress-strain curve from 0.2 % strain intersects the curve.
This value is also n;:ferred to as the "offset yield stress".
It must be emphasized that whenever the applied stress exceeds the "anelastic
limit" some plastic deformation occurs in the material, and such deformation is
usually inhomogeneously distributed for polycrystalline solids. Analysis of such
behaviour is possible by the extension of the theories treated so far into the micro
scale, where the interaction between regions of different properties is also taken
into account. These procedures will be treated in detail in Chap. 3.

Problems
2.1. The stress tensor in the
0" . .
1J

= (-500 -400

30

Sj

coordinate system is given below:

30) MPa.
200

2 Fundamental Concepts in Stress Analysis

46

a) Calculate the stress tensor in the L coordinate system if: cos - 1 a 33 = 45, and x~
is in the plane defined by X l ,X 2 and is rotated 60 counterclockwise from x2 .
b) Calculate the strain tensor for both coordinate systems, assuming an isotropic
Fe sample.
c) Calculate the principal stress tensor for this system.
2.2. Assume that the stress

crij= (

-400 0)
0 0 0 MPa

is applied in the [100] direction ofa Fe single crystal. Calculate the strain tensor in
the crystal axes.

Bibliography
L.H. Schwartz and J.B. Cohen, "Diffraction from Materials', Academic Press, New York, N.Y.,
1977
G.E. Dieter, "Mechanical Metallurgy", 2nd. ed., McGraw-Hill, New York, N.Y., 1976
T. Mura, "Micromechanics of Defects in Solids", Martinus Nijhoff Publishers, Hague, The
Netherlands, 1982
J.F. Nye, "Physical Properties of Crystals", Oxford University Press, Oxford, Great Britain,
1976
S.P. Timoshenko and N.J. Goodier, "Theory of Elasticity", 3rd. ed., McGraw-Hill, New York,
1970

3 Analysis of Residual Stress Fields Using Linear


Elasticity Theory

3.1 Introduction
In one of the most comprehensive books on micro plasticity, Mura[1] defines

residual stresses as the "self-equilibrating internal stresses existing in a free body


which has no external forces or constraints acting on its boundary". These stresses
arise from the elastic response of the material to an inhomogeneous distribution of
nonelastic strains such as plastic strains, precipitation, phase transformation,
misfit, thermal eKpansion strains, etc. For example, mechanical deformation
processes that cause plastic deformation in the surface layers of the material, such
as shot-peening, grinding, machining, etc., cause residual stresses in these layers
because of the constraining effect of the bulk, where plastic deformation is
minimal. These stresses are called macros tresses. Since the surface layers will also
constrain the bulk in return, the bulk material will also have residual stresses even
though it may not have suffered deformation. It is also possible to have residual
stresses when a multi-phase body, where the phases have different yield points, is
pulled in uniaxial tension. The macrostress field in the deformed material will be
negligible since the material will have the same plastic strains at all depths. The
inhomogeneous distribution of yield points in the material volume, however,
causes an inhomogeneous partitioning of the plastic strains between the phases,
which, due to the constraining effect of the stronger phases on the weaker ones,
causes a residual stress field to form. Residual stresses of this type are called
microstresses. These stresses can be present in a single phase material also, if, for
any reason, plastic deformation is inhomogeneously distributed between the
grains of the polycrystalline body.
In this chapter the relationships between the nonelastic strains and the residual
stresses they cause will be investigated. Most of the equations given are for
macroscopic, or for micro-mechanically ideal systems, which can not be used
easily in practice. However, the fundamental concepts which link stress to strain in
all cases, and the physical boundary conditions they must obey, are important to
our understanding of what is measured by diffraction methods, and must be
known for the correct interpretation of measured stress values.

3.2 Macroresidual Stresses


Consider the case of shot peening. It is assumed that spherical shot is arriving
normal to the surface layers of a flat work piece (of dimensions L in the X 1,X 2

3 Analysis of Residual Stress Fields

48

X2

Workpiece
Fig. 3.1. Schematic description of the normal shot-peening process

directions and of initial thickness to), and plastically deforming these layers
(Fig. 3.1 ). FOlt a homogeneous, isotropic material, peened with a large number of
shot impinging randomly on the surface, the plastic strain tensor e~ in the surface
layers will be

(3.1 )

The plastic strain component e~3 is produced due to the force exerted by the shot
at the time of impact in the X3 direction. Since the material is, in effect, compressed
in the X3 direction, plastic flow also occurs in the X I ,X 2 directions, and from
conservation of volume e~2 = ell = - e~3/2. Thus, the plastic deformation distribution in the plane of the surface is homogeneous, as expected from the symmetry
of the process causing the deformation.
The plastic strain terms ell,e~2 imply a length change in the surface layers, i.e.
if the surface layer, of initial length Lo and thickness t? (t? ~ to), could be
separated from the rest of the sample and shot peened, the final dimensions would
be
L'=Lo+Jel'ldx1 =Lo+oL
tl

= tl0 + JeP33 d x3 ,

(3.2 )

where it is assumed in this example that the plastic deformation at all depths in the
surface layer is constant.
This layer is now placed on the rest of the sample (the bulk), which has
suffered no plastic strains and is still oflength Lo. In order to match the dimensions
of the surface layer and the bulk, surface tractions F must be applied at the
boundaries (Fig. 3.2) which will elastically deform both the bulk and the surface
layer to a final length L, where L' > L > Lo (Fig. 3.2b ) . The tractions, F, will cause
macro residual stress fields of opposite signs in the surface and in the bulk volume.

49

3.2 Macroresidual Stresses


-

Lo ... b L -

--I~
lt=
1 ====~I~
F ;-==~

t -11

- - L o-

,I

f--

-F

- - L - - - -.,

a
b
Fig. 3.2. Free-body deformation of the surface and bulk layers without mutual constraint (a),
and with mutual constraint (b) of a hypothetical shot-peened specimen

0'1
O~~----------------~

O~--~--------+-~

Fig. 3.3. Variation of the in-plane stress (cr 11) with depth (x 3 ) in a hypothetical specimen shotpeened on one side (a), and on both sides (b)

The magnitude of the macrostresses in the bulk and the surface can be calculated
from Hooke's law:
L
( all) s = Eln L'

L
(all h=E-ln Lo .

(3.3 )

The stress (all) s will be equal to (a 22) s due to the symmetry of deformation in
the plane of the surface. The stress distribution in the work piece after reattachment is shown in Fig. 3.3a. This stress distribution will produce a couple
which will cause bending of the specimen to a curvature co. This curvature is
proportional to the stresses in the volume and can be used to calculate the residual
stress value at the surface when the stress distribution is known. This concept
forms the basis of mechanical residual stress determination methods. If the plate is
shot peened with the same intensity on both faces, the stress distribution is
symmetric with respect to the central plane of the specimen (Fig. 3.3b ) . In this
case the stress distribution can be represented by two equal but opposite couples,
and the specimen remains flat.

3 Analysis of Residual Stress Fields Using Linear Elasticity Theory

50
E

o~----~-------------

I t,

I
I

0'11

I
I
I

I
I

O~--~I C=========~
Fig. 3.4. Variation of the in-plane plastic strain
( E) (a) and in-plane residual stress (b) with
depth in an actual specimen
b

In this example a plastic strain distribution described by a step function was


assumed. In most actual specimens the plastic deformation, and hence the residual
stress, decays gradually as one moves into the material (Fig. 3.4 ). In such cases it
is possible to treat the surface as made up of layers of infinitesimal thickness, in
which the plastic strains are constant with depth. The residual stress profile is then
obtained by considering the constraint imposed by each layer on its neighbours. It
must be noted, that the maximum plastic deformation does not occur in the
outermost surface layer in all cases. It is also possible to have unconstrained plastic
flow between the surface layers, which causes relaxation of the mutual constraint,
if the peening intensity is high enough. For these cases, the residual stress profile
will exhibit a m~ximum deeper into the material. The basic treatment for all cases
is similar, however.
Macroresidual stresses are also created in processes where parts of different
lengths are joined together by welding or fasteners in a way that mutual constraint
occurs. For example, consider the case of the shrink fitting of a tube on a solid
shaft. A tube of radius a - 0, is expanded by heating and then fitted on a shaft of
radius a + 0, after which the system cools down to room temperature (Fig. 3.5) .
At room temperature both the tube and the shaft will have residual stress
components in the tangential and radial directions:
(3.4 )

3.3 Equations of Equilibrium for Macrostresses

51

~z
Us

Fig. 3.5. Stress state set-up when a tube of initial inner


radius of a-o is fitted on a shaft of radius a+o

in the shaft, and


(3.5a)

cro) Tu

(a-o)2(crr )Sh ..
b2 -a 2

(1 + r

b2)
2

(3.5b)

in the tube. (The detailed solution for this case may be obtained from any book on
elasticity theory, for example see Timoshenko [2J, pp 65 - 75.) It may be noted
that, in contrast to the shot-peening case, in which no stress component in the
direction of the surface normal occurs, the shrink-fitting has such a component
( crr ) , which decays to zero at the surface where r = b, and thus exists as a gradient
with depth.

3.3 Equations of Equilibrium for Macrostresses


Consider the shot-peened plate discussed above. After peening the specimen is at
equilibrium, and therefore the net force acting on the total cross-sectional area
(A) of the specimen must be zero. The equation of equilibrium relating the
compressive residual macrostresses in the surface layers to the tensile residual
macrostresses in the bulk is:
(3.5 )
Similarly, for the bending moments:
( 3.6)

52

3 Analysis of Residual Stress Fields Using Linear Elasticity Theory

In addition to Eq. (3.5) the macrostresses at a point must obey the differential
equations of equilibrium given in Sect. 2.10:
O"ijnj=Fi

(3.7a)

O"ij,j=O.

(3.7b)

Since the macrostress tensor produced by shot-peening is homogeneous,


Eqs. (3.7a,b) are automatically satisfied at all points in the material volume.
Equations (3.7a,b) also limit the types of stress states possible in the material. For
example a triaxial macrostress tensor of the form
(3.8 )
is not possible in a shot-peened surface, since the force and hence the stress
(Eq. 3.7a) is zero at the surface. Even deeper down in the surface layers 0"33
cannot be finite. This can be seen by writing Eq. (3. 7b) for i = 3:
00"13

ox!

+ 00"23 + 00"33
oX 2

oX 3

=0

(3.9 )

since the terms 00"13/0Xl' 00"23/0X2 are zero, the term 00"33/0X3' which describes
the variation of 0"33 with depth is zero. Thus the value of 0"33 at any depth is equal
to its value at the surface, which, as shown before, is zero.
In the case of shrink-fitting a tube on a solid shaft there is a stress gradient
along the radial direction, even though the stress at the surface is zero. The
tangential and radial stresses caused by the shrink-fitting operation also obey the
equations of equilibrium (3.7a,b). This can be easily seen by writing the force
balance in cylindrical coordinates [2] and is left to the reader as an exercise,

3.4

~crostresses

It can be seen from the discussion given above that macrostresses are homog-

eneous on a macroscopic scale along at least one direction. For example, the radial
stress for the shrink fitted tube at any depth is independent of e. For the case of
shot-peening discussed before, the stress 0" 1 1 is independent of Xl ,x 2 and depends
only on x3 . Microstresses are, however, usually inhomogeneous on a macroscopic
scale and may also be inhomogeneous on a microscopic scale as well, such that the
components of the microstress tensor are different at all points in the material
volume. These stresses may arise from a number of causes. The inhomogeneous
distribution of plastic deformation on a micro scale, such as the partitioning of
plastic deformation between the matrix and precipitates of a two phase alloy,
causes a microstress field in and around each precipitate. This case may be
visualized as follows. Assume that initially there are no stresses of any kind in the

uu o o

3.4 Microstresses

01

Ie>

~
ol
a

53

Fig.3.6a-c. Formation of microstresses due to inhomogeneous partitioning of plastic strains in a twophase body

material volume. ' If a precipitate could be removed from the material without
disturbing its surroundings, the dimensions ofthe precipitate and the void which it
leaves behind in the matrix would be identical. However, since the yield points of
the matrix and the precipitate are different, a given applied stress will cause
different plastic deformation in the precipitate and the matrix. This means that
after plastic deformation, the dimensions of the precipitate and the void are no
longer identical. In order to insert the precipitate into the void, tractions must be
applied to the boundary of the void and the precipitate such that they match each
other at all points (Fig. 3.6). These tractions will cause a microstress field in and
around each precipitate. Such microstresses will be referred to here as "microstresses due to differential plastic deformation". An inhomogeneous distribution
of elastic constants in the material volume may also cause formation of
microstress fields when a homogeneous stress is applied at the boundary of the
material. Consider, for example, a sample consisting of two contiguous grains
with different cryst~llographic directions (and hence elastic constants Sijkl ) along
a vector V across the grain boundary. This bi-crystal is subjected to the
homogeneous stress CJ 11 at the surface (Fig. 3.7). If these grains were subjected
separately to CJ l l on their boundaries, each would have a different strain
[Eq. (2.39) J, and ~ence different displacements [Eq. (2.2) ] along V. However,
in a solid body, the grains are in rigid contact and displacements across the grain
boundary must be constant in order to avoid having voids at the boundary.
Therefore, a microstress field arises which modifies the applied stress and thus
constrains the displacements in the material surrounding the grain boundary.
Such microstresses are called "microstresses due to elastic incompatibility".

0'1

~~o,
V
1

<J

_8

I)

Fig. 3.7. Fonnation of micro-stresses


due to elastic incompatibility in a bicrystal subjected to a homogeneous
surface stress

3 Analysis of Residual Stress Fields Using Linear Elasticity Theory

54

depth below surface

Fig. 3.8. A piecewise continuous PM-stress distribution

Experimental determination of the microstresses at a point requires a probe of


infinitesimal size and is, thus, impractical for most cases. Probes of finite size, on
the other hal!d, will yield information about the average microstresses in the
measuremeni'volume. In this book the term "pseudo-macro stress (PM stress) "
will be used to denote the average microstress, where the average is taken over a
volume which contains a statistically representative number of grains. Thus, the
average is homogeneous in a given phase. This means that, for a multi-phase
material where the structure consists of a continuous matrix and discrete
precipitates of the other phases, the PM stress is piecewise continuous. As long as a
representative volume is measured, the average stress in a given phase is constant
but not equal to the average stresses in the other phases, independent of the
location of the measurement volume within the specimen. A possible di~tribution
of this type is shown in Fig. 3.8.

3.5 Equations of Equilibrium for Microand Pseudo-Macrostresses


Both types of microstresses discussed above are self-equilibrating, and obey the
differential equations of equilibrium:
O"jj,j=O

(3.7b)

in the volume, and


(3.10 )
at the boundary.
The relationship between the average stress component in the matrix and the
average stress component in the precipitates can be derived as follows. From the
equations of differential equilibrium (3. 7b ), (3.1 0), it can be shown that the
average of any stress over the whole body is zero,
(3.11a)

55

3.5 Equations of Equilibrium for Micro- and Pseudo-Macro Stresses

where D is the total volume of the body. It must be emphasized that this equation
is valid if and only if Eqs. (3.7a,b) are satisfied, as is shown below. Rewriting
Eq. (3.11a):
(3.11b)
( Ojk = Xj,k, since oxiox = 1 for i = j and = 0 otherwise)
Integrating (3.11 b) by parts 1, with "xt the coordinate in the j direction:
J O"ijdD = J O"ikxjnkds - J O"ik,kxjdD = 0 .
D

I~

(3.12 )

The first integral on the right-hand side is over the boundary of the body ID I, where
O"iknk=O [Eq. (3.10)], and the second integral on the right-hand side, evaluated
in the volume, is zero as O"ik,k=O [Eq. (3.7b)].
For a two-phase material Eq. (3.11) can be written as
J O"ijdD =
D

D-Q

O"ijdD + J O"ijdD = 0,
n

(3.13a)

where Q is the volume of the second phase. Rewriting Eq. (3.13a):


D-Q
D-Q

DLn O"ijdD+ nJO"ijdD=O

(3.13b)

or,
(3.13c)
where the carats "( )" imply averages over the appropriate volume. Multiplying
both sides of Eq. (3.13c) by liD:
(3.14a)
Here f is the volume fraction of the second phase, and (O"ij)m, (O"i)ppt. are the
average stresses (PM stresses) in the matrix and precipitates (second phase)
respectively.
.
For a multi-phase material a similar derivation yields:
N

L f;"( O"ij)i = 0

i= 1

(3.14b)

where N is the number of phases, fi is the volume fraction of the ith phase which
has the average stress (O"i)i' It can be seen from the above discussion that when
the PM stresses obey Eqs. (3.14a) or (3.14b), the actual stress distribution at all
points in the body volume obeys the differential equations of equilibrium
(3.7a,b). Thus, even though 0"33 can not exist as a macrostress gradient in the near
surface regions of a shot-peened homogeneous material [Sect. 3.2, Eq. (3.9)],
0"33' as well as 0" 13,0"23 can exist as pseudo-macro stress gradients in a shot-peened
two-phase material. In this case, the gradient for each PM stress component
(0"33),(0"13),(0"23) is balanced in the grains of the respective phases In
1 Integration by parts utilizes the formula Judv=uv-Jvdu, where u,v are functions in x.

56

3 Analysis of Residual Stress Fields Using Linear Elasticity Theory

accordance with Eq. (3.14). (The shear components will be especially important
if peening is not normal to the part. An example of this is given in Chap. 7,
Example 7.7.)

3.6 Calculation of Micro- and PM Stresses 1


The solution for the stress fields in and around an inclusion in a material that is
subjected to a homogeneous stress field on its surface boundaries, or the stress
fields arising due to the inhomogeneous partitioning of plastic deformation
between the matrix and the inclusion, has been treated extensively in the literature
for various inclusion shapes and stress and deformation configurations. The
classical works on the subject are by Eshelby [3,4,5] who showed that if an
inclusion in an infinite, isotropic matrix suffers inelastic strains e;j (where e;j may
be plastic strains e~, transformation strains elj etc: such strains are also termed
eigenstrains) while the matrix remains undeformed, the elastic strain in the
inclusion due to the constraining effect of the matrix is homogeneous (i.e.
constant in the inclusion volume) and is related to the eigenstrain e;j by
(3.15 )
where Sijkl are components of a fourth rank tensor called Eshelby's tensor. For an
ellipsoidal inclusion Sijkl are functions of the Poisson's ratio v and the aspect ratio
of the ellipsoid aJa j for all ai' where a i is shown in Fig. 3.9. Sijkl for a spherical and a
penny shaped precipitate are shown in Table 3.t.
Once the elastic strains eij in the inclusion are obtained from Eq. (3.15), the
stresses in the inclusion volume may be obtained from Hooke's law. Thus, the
stress field within the precipitate is also homogeneous.
The solution given by Eshelby [4] for the stress field at points outside the
precipitate (exterior points) is more complicated. This field too is dependent on
the shape of the precipitate but, in addition, contains terms dependent on the
position of the particular point in the matrix where the field is required. Thus, the
stress field in the matrix is inhomogeneous and varies from point to point. Even
though the solution is different for different inclusion shapes, for all shapes the
stress decreases as one moves away from the inclusion-matrix interface, decaying
to zero at the material boundary at infinity.
Similar solutions also apply when an external stress, 0"3, is applied at the boundary of an infinite material which contains an inclusion whose elastic constants C;jkl
are different than those of the matrix (i.e. both the matrix and the inclusion are
homogeneous within their respective volumes, however the composite is an
inhomogeneous material). For this case, the elastic incompatibility at the
1 The following section is given as a brief introduction for the methodology of calculating micro
and PM stresses by micro-mechanical methods. It also serves to illustrate the different effects
that contribute to such stresses and how they are taken into account. The procedure is
applicable to ideal systems only, and cannot be used easily in practice. The reader may just
examine the final equations for PM stresses, to get an idea of the terms contributing to this type
of stress.

3.6 Calculation of Micro- and PM Stresses

57

Fig. 3.9. Definition of the axes of an elliptical


precipitate
.
Table 3.1a,b. Eshelby tensor components Sijkl for (a) sphere, (b) penny shaped inclusion

7-5v
Sl1l1 =S2222 =S3333 ~ 15( 1-v)
5v-1
S1122 =S2233 =S3311 =S1133 =S2211 =S3322 = 15( 1-v)
4-5v
S1212=S2323=S3131 = 15( 1-v)
all other Sijkl =0.
b Penny Shape (a l = a 2 ~ a3 )
1-2v 1t a 3
13-Sv a3
Sl111 =S2222= 32(1-v) 1t ai' S3333=1----1-v 4 a l
Sv-1
a3
2v-1
a3
Sl122=S2211 = 32(1-v)1t al , S1133=S2233=S(1_v)1t al ,

7-Sva3
1(
V-21ta3 )
S1212= 32(1-v) 1t ai', S1313=S2323= 2 1 + 1-v"4 a l '
1 - 2v 1t a3
Skkl1 = Skk22 = -1-- -4 -val

+1 ()'
-v

1 - 2v 1t a 3
Skk33 = 1 - -1-- -2 -val

inclusion-matrix interface will cause a stress disturbance at all points within the
material volume such that the stress at any point in the matrix will be
(3.16)

where (Jij is the stress caused by the elastic incompatibility, egl is the homogeneous
strain that would occur in the matrix with elastic constants Cijkl if no

58

3 Analysis of Residual Stress Fields Using Linear Elasticity Theory

incompatibility was present and Ekl is the elastic strain due to O'ij.1t can be seen that
Eq. (3.16) is simply Hooke's law written for the matrix. Similarly Hooke's law for
the precipitate is

(3.17 )
where the terms are similar to the ones for the matrix but are defined for the
inclusion with elastic constants C;jkl'
The stress/strain values for this problem may be solved through the equivalent
inclusion method of Eshelby [5]. In this method the goal is to transform the
problem into the eigenstrain problem for a homogeneous material where both the
matrix and the inclusion have the same elastic constants C jjkb and then solve it
using the procedure discussed above. Eshelby showed that this is possible by
proving that a fictitious homogeneous inclusion having the same elastic constants
C jjk1 as the matrix will be equivalent to the actual (inhomogeneous) inclusion
when the equation

(3.18 )
is satisfied in the inclusion volume. In this equation the right-hand side is the stress
for the fictitious equivalent inclusion and E;j are the inelastic strains (eigenstrains)
that must be introduced in a homogeneous inclusion of the same shape and size as
the actual inclusion such that the equivalency is satisfied. It must be noted that
such equivalency cannot be achieved if elastic strains are introduced instead of E;j
since this would change the overall stress state through Hooke's law.
Substituting (3.15) into (3.18):

(3.19)
which may be solved for the unknowns E;j' Then Ejj is determined from (3.15) and
the stress state for the interior points obtained from Hooke's law (3.17). For
explicit solutions for a number of inclusion shapes and for solutions for exterior
points the reader is referred to the original papers by Eshelby [3,4,5] or to Mura

[1,pp 151-158].

The pseudo-macro stresses are volume averages of the microstresses discussed


in the above examples. Since most engineering materials contain precipitates or
grains that are much smaller than the probes (such as a strain gage or standard xray beam) that, are used to obtain information about the stresses, the measured
values contain pseudo-macro components rather than the individual micro-strain
values. Thus a procedure for evaluating the macroscopic average of the microstresses is required. In the following a procedure developed by Tanaka and Mori
[6,7] and Brown et al. [8,9] is used for this purpose:
Consider a two phase slab of total volume D which has "N" precipitates that
are elliptical in,shape, and are randomly distributed. The total volume of the
precipitates is O. There are plastic strains E~ that are uniformly distributed in each
precipitate. Thus the slab contains elastic strains and microstresses due to
differential plastic deformation. However the slab is not subjected to any external
stress or constraint and the equation of equilibrium for the average stresses (PM
stresses) existing in the slab is given by Eq. (3.14).

3.6 Calculation of Micro- and PM-Stresses

59

The average stress in the matrix <crij>m can be written as


( 3.20)
where e~1 is the average elastic strain in the matrix. If a single precipitate is
randomly inserted into the matrix (for large N, this will not change f), the stress in
this new precipitate will be
(3.21 )
Here crij is the stress calculated for a single precipitate, containing the plastic
strains Elj, present in an infinite matrix.
Since the precipitate can be inserted at any place in the matrix, crjj' given by
Eq. (3.21), is the average stress in all the precipitates.
Writing Hooke's law for the precipitate:
( 3.22)
where e~1 is the average elastic strain in the matrix and Ekl is the total strain in the
precipitate containing plastic strains E~I' Thus, the term in the brackets is the total
elastic strain in the precipitate inserted into the matrix.
Solution ofEq. (3.22) is possible through the equivalent inclusion method of
Eshelby. Writing Hooke's law for the equivalent inclusion which has elastic
constants C jjkl (the same as the matrix):
(3.23 )
where E~I is the nonelastic strain (eigenstrain) that must be introduced into the
inclusion such that
P)-C
.. ) ,
C'ijkl (0
ekl + Ekl - Ekl
- ijkl (0
ekl + Ekl - Ekl

(3.24 )

. where E~; = E~I + E~I'


Assuming the matrix and the precipitate both to be isotropic, Eq. (3.24)
becomes
21.1' {e~ + Eij - Elj}

+ OijA' {e~k + Ekk -

E~k}

(3.25 )
Here Jl,A,Jl' and A' are the shear modulus and Lame' constants of the matrix and
the precipitate respectively.
Equation (3.25) can be expressed in a more useful form by defining the
deviatoric components of the strains contributing to it:
'0
s: 0
eij
= ei. - uije
kk /3

Elj = Efj - OijE~k/3


,..
s: "'3
Eij
= Eij.. - uijEkk
.

( 3.26)

60

3 Analysis of Residual Stress Fields Using Linear Elasticity Theory

Substituting Eqs. (3.26) into Eq. (3.25), and re-arranging:

' - eij
'p} = 211 {'O
' - eij
,..}
2 11*{'O
eij + eij
eij + eij

(3.27 a)

K* {e~k+ekk-e~k} =K{e~k+ekk-e;k}'

(3.27b)

where K" and K are the bulk moduli of the precipitate and matrix respectively
[K = (211 + 31..) /3].
For shear strains (i =F j), Eqs. (3.27) become
(3.28 )
since all oij = 0 for this case.
The unknowns eij,e~ can be expressed in terms of each other by the property
(3.29)
where Sijkl is the Eshelby tensor components associated with the precipitate shape.
From Eqs. (3.28) and (3.29) (expressing e;; in terms of eij' where i =F j, and
noting that the terms Siiik,Siikl are zero):
.. _ {2(1l-1l*)e?j+21l*e~}!, .4-.
eij- {4( *_)
2} 10rl,J.
11 11 Sijij + 11

(3.30 )

Another equation between the unknowns e~,e~ is obtained by re-writing


Eq. (3.14). This is done by substituting the respective stress values expressed in
terms of strains [Eqs. (3.21 ) , ( 3.23 )] into Eq. (3.14) and then expressing eij,e~ in
terms of each other by substituting Eq. (3.29) into the final equation:
21le~ + oij1..e~k + f{ 211 (Sijkle~;- e;;) + Oij1.. (Sllkle;;+ S22kle;;+ S33kle~;- e~~) }

=0.

(3.31 )

For shear strains eij' i =F j, all oij = 0, and we have


(3.32 )

21le~ + f{ 211 ( Sijkle~; - e;;) } = 0

then, substituting (3.30) into (3.32):

IJ

IJ

e=e

21l*f( 1- 2Sijij )
*
4(1l-Il)sijij+21l-f(1-2sijij)(21l-1l*)

lor I=FJ.

(3.33 )

Thus the average (PM) shear stresses in the matrix are

<(J'ij)m = 2Ile~,

i =F j .

(3.34 )

Here e~ is given by Eq. (3.33). The PM shear stresses in the precipitate are then
obtained from Eq. (3.14).
Unfortunately, the solution for the normal stresses (i=j) is much harder:
Equations (3.26) and (3.29) yield three equations in the six unknowns e~ and e;;
(for i = j ). Another three equations between e~ and e;; are obtained from the
equilibrium equation (3.14). Then e~,e~ can be obtained from the simultaneous
solution of these six equations. Since the terms in the equations are too long and
cumbersome to express in closed form, explicit solutions for e~ are not given here.

3.7 The Total Stress State in Surface Deformed Multi-Phase Materials

61

However, it can be seen by inspection of Eqs. (3.33) - (3.37) that these strains
will be functions of the plastic strain E~ in the direction xj, the total plastic strain
E~k and the Eshelby tensors Sijkl associated with the shape of the second phase
particles. Once the eS are determined, the PM stresses O"ij) in the matrix can be
evaluated from Hooke's law. The stresses in the precipitate can then be obtained
from Eq. (3.14).
In this procedure the interaction between the stress fields around individual
inclusions was not taken into account. This approximation is valid for f ~ 1. For
large f, the solution will follow along the same lines. However, an average
interaction term will also contribute to the PM stresses. If the surface is near the
region of the calculation, a surface effect term (similar to the "image stress"
associated with dislocations) is also needed. These effects are complicated and will
not be treated here. However, the PM stress contributions from both of these
effects will obey the equation of equilibrium (3.14).
When an external load is applied to an inhomogeneous two-phase material
(i.e. the elastic constants of the second phase are different than those of the
matrix) which already has nonelastic strains (eigenstrains) in the second phase
grains, the PM stress tensor will have components from both types of microstresses. A procedure similar to the ones discussed above can be used to determine
the PM stresses in the matrix and precipitates (second phase particles) for this
problem. Writing. Hooke's law for the matrix and the precipitates:

<

O"S + O"ij = C;jkl ( egl + Ekl - E~I)

(in the precipitates)

O"S + O"ij = Cijkl ( egl + Ekl )

(in the matrix).

(3.35 )

Here O"ij is the residual stress created due to the differential plastic deformation
plus the elastic incompatibility. The other terms are equivalent to those discussed
in Eq. (3.16).
To solve Eq. (3.35) one can use the equivalent inclusion method discussed
above. For the actual derivation and discussion on a number of precipitate shapes
the reader is referred to Mura [l,pp 151-168,334-342].

3.7 The Total Stress State in Surface Deformed


Multiphase' Materials
In the previous sections macro- and microstresses were treated separately. In
actual practice, however, both stresses may exist together, and usually do. For
example, assume that a slab of initial length L, which consists of an isotropic
matrix, (with elastic constants E,v), and randomly distributed isotropic precipitates (with elastic constants E', v' ), is shot peened on both sides. In the surface
layers, where plastic deformation due to peening occurs, the total stress at a point
is
O"~j = ( O"ij ) DPD + O"S + ( O"ij ) EI + ( O"ij ) SE. + ( O"ij) INT. ,

(3.36)

62

3 Analysis of Residual Stress Fields Using Linear Elasticity Theory

where (O'ij) DPD are the microstresses due to differential plastic deformation
between the matrix and the precipitates, (O'~) is the macrostress which will form
due to the constraining effect of the bulk material which has suffered no plastic
deformation [the effect of the bulk can be represented by the appropriate
tractions F acting on the boundary of the surface layer (Fig. 3.2)], and ( O'ij )EI are
the microstresses due to elastic incompatibility which are caused by the tractions,
F i, imposed by the bulk on the surface layers. The terms (O'ij )INT> (O'ij )SE describe
the microstresses due to the interaction of the elastic fields of the precipitates, (this
term is zero for f ~ 1. ) , and the microstresses due to the presence of a free surface
respectively.
The stresses measured in the surface layers from a given phase will then be
<O'ij)t=O'~+ O'ij)

+ (O'ijhI+ (SiJSE+ (O'ij)INT),

(3.37)

where the terms are volume averages of the stresses contributing to Eq. (3.36).
It must be noted that the equations of equilibrium applicable to the terms
contributing to the total average stress are different. For the PM components
<(O'ij), Eqs. (3.14) apply individually or in total, whereas for the macro-stress
component, the~equations of equilibrium are
Here F are the tractions imposed by the bulk on the surface layers as discussed
before.
The stress state in the bulk is simpler. There will be no microstresses due to
differential plastic deformation between the phases (since the bulk suffers no
plastic deformation at all) , and the surface effect term (O'ij) SE may be neglected.
However, the other two components contributing to the total stress tensor at a
point will be finite.

3.8 Macroscopic Averages of Single Crystal Elastic Constants


In the discussions given above a multi-phase solid is treated as a piecewise
homogeneous body. That is, even though the total material is an inhomogeneous
solid, each precipitate is considered to be homogeneous within its volume and
equivalent to all other precipitates of the same phase, and the matrix itself is
homogeneous at 'all points outside the precipitates. Furthermore, the use of
Hooke's law to obtain the average stresses from average strains [Eqs. (3.34)]
implies that the distribution of elastic constants, Cij , within the constitutive phases
is homogeneous such that along a given direction V in the volume of the material,
the elastic constant Cij changes only when the vector V crosses phase boundaries,
as shown in Fig. 3.10a. This condition is automatically satisfied when both the
matrix and the precipitates are isotropic but have different elastic constants.
However, in a real polycrystal a vector V in the material volume will intersect
grains (of a given phase) with different crystallographic orientations (figure
3.10b). In this case a volume average of the elastic constants Cij is required to
determine the average stresses from the average strains and vice versa.

3.9 The Voigt Average

63

a
b
Fig. 3.10. Distribution of elastic constants in the volume of a piecewise homogeneous body (a),
and an inhomogeneous body (b)

The evaluation of the average elastic constants of polycrystals and multi-phase


materials is one of the important problems in micromechanics. Various approaches to the problem have been suggested over the years. In all of these, various
models have been used to account for the partitioning of stress and/or strain
between the constituent grains of the multi-phase body. In the earliest work on the
subject, Voigt [10] assumes a constant strain distribution in all the grains of the
material. Reuss [11] treats the case where it was assumed that all grains have a
constant stress. Hill [12] shows that the Voigt and Reuss limits are, respectively,
upper and lower bounds for the average elastic constants 1 and suggests the use of
the mean of these bounds as an average elastic constant, which, indeed yields
values close to the measured elastic moduli. More recent analyses with various,
more sophisticated, models (Eshelby [3,5]; Kroner [13,14] have also been
proposed. These, in general, yield values close to the mean of the Voigt and Reuss
limits. In the following sections these approaches will be briefly reviewed.

3.9 The Voigt Average


In the Voigt approximation it is assumed that in a polycrystalline or composite
body sUbjected to a stress at the boundary, all composite elements or grains are
subjected to the same uniform strain E~. Assume, for example, that a shear stress
0" 12 is applied at the boundary of the material. The relation between the average
stress cr 12 (which contains contributions from the elastic incompatibility between
grains) and the strain E~2 at any point is then given by

(3.38)
Here y is the engineering shear strain [Eq. (2.36)] and D-v is the average shear
modulus in the Voigt limit.
For a piecewise isotropic multi-phase material, the average shear modulus is
defined as
D

D-v=

L filli
i=O

(3.39)

1 The methods of Z. Hashin and S. Shtrikman (J. Mech. Phys. Solids 10, 343 (1962); 11,
127 ( 1963) ) which are based on a variational principle, generally yield better upper and lower
bounds for the elastic moduli than the Voigt and Reuss limits. These calculations are
complicated however, and are beyond the scope of this book. The reader is referred to M ura [1]
for a review of these methods and their extensions.

3 Analysis of Residual Stress Fields Using Linear Elasticity Theory

64

Here Il; is the shear modulus of the ith phase which has a volume fraction f; in the
material.
For a material composed of anisotropic single crystals with elastic moduli
Cijkl' the average elastic moduli C;jkl can be calculated in a similar manner. In this
case the average stress is related to the uniform strain through the generalized
Hooke's law. For example, the average stress for the shear strain S~2 is
( 3.40)
Here D is the volume of the material and the average elastic constant
the Voigt limit is given by
n

(C 1212 )y=

;=0

fdC 1212 );

J C 12dD in
D

(3.41 )

Here (C 1212 );, the elastic modulus of the ith crystal, is referred to the same set of
axes in which S~2 is measured.
Calculation of the bulk modulus is similar: Consider the case where a
hydrostatic stress O'H is applied to the boundary of a material consisting of i
discrete phases and the matrix. In this case the Voigt limit assumes that all
elements of the solid have the same dilatation sa. The average hydrostatic stress
within the body is then evaluated from
n

crH = ;=0
L r;K;"sa

(3.42 )

Thus, the bulk modulus, by definition, is


n

Kv= L

;=0

f;"K;.

(3.43 )

If the orientation distribution of the constitutive single crystals is random, all


crystallographic orientations can be found with equal probability at all points
defined within the sample volume. In this case the average elastic moduli of the
material will be isotropic since the same average is obtained along any direction.
For example, it a hydrostatic stress O'H is applied at the boundary of a
polycrystalline material with a random distribution of crystallites, the normal
stress, O'jj, at any point within the material will be
(3.44)
The average hydrostatic stress in the total body is found by integrating Eq. (3.44)
over the total volume:
-

0'

s~; JCjjjjdD.
= 3"1 JD O'jjdD = 9"

(3.45 )

Cjjjj is an invariant tensor quantity within the volume (Nye [15] ,p 146). Thus
Eq. (3.45) becomes
_

0' =

9" CjjjjSjj

(3.46 )

3.10 The Reuss Average

65

from which the average bulk modulus of the polycrystal in the Voigt limit is found
as
(3.47 )
Expanding Eq. (3.47):
9Kv= {Cllll +C 2222 +C 3333 } +2{C 1212 +C 2233 +C 3311 }

(3.48 )

which, for cubic crystals, is equal to


(3.49 )

9Kv=3C ll 11 +6C 1212

since C llll =C2222=C3333, and C1212=C2233=C3311. Equation (3.49) is


normally written in the matrix notation as
(3.50)
The average shea~: modulus in the Voigt limit of a polycrystal with a random grain
distribution is also isotropic. The derivation for this case is similar to that of the
bulk modulus, but more complicated. For this case, for cubic systems Eq. (3.41 )
becomes
(3.51 )
or, in terms of elastic compliances (in the matrix notation)
(3.52 )
since
1
C ll -C 12 = S11'-s 12 '

1
C44 =-s
44

(3.53 )

Poisson's ratio v and Young's modulus E can be evaluated from Eq. (3.50) and
(3.51) using the formulae

K=

3 (1-2v)

(3.54 )

3.10 The Reuss, Average


In the Reuss approximation it is assumed that the stress in the elements of a
composite material (or the grains of a polycrystal) is equal to the average stress
applied to the material. For piecewise isotropic materials this model assumes that

66

3 Analysis of Residual Stress Fields Using Linear Elasticity Theory

the average shear strain is related to this uniform shear stress through an equation
of the form
(3.55 )
Thus, the average shear modulus in the Reuss limit is

ilR =

.L

,=0

fJlli

)-1

(3.56 )

Similarly, the average bulk modulus in the Reuss limit is given by


(3.57 )
For a random polycrystalline sample, the average elastic moduli in the Reuss limit
is obtained from the average of the single crystal compliances Sijkl over all possible
directions. The average bulk and shear moduli for the general case are
,.
1
KR = (Sl1 +S22+ S33) +2(S12+ S23+ 8 3d
(3.58a)
15

IlR

=4(Sl1 +S22+S33) -4(S12+S23+83d +3(S44+S55+866)

(3.58b)

which, for cubic crystals become


KR =
5

ilR

~ {(8

11

:28 12 ) } =

=4(Sl1- S 12) +3S 44

~ (C

ll

+2C12 ) =Kv

( 3.59a)
(3.59b)

It must be noted that the fundamental assumption in the Reuss model, that
stress is constant over the total volume of the polycrystalline body, does not satisfy
compatibility at the grain boundaries, in that two points on either side of a grain
boundary between grains of different orientation would suffer different displacements due to the different respective strains given by Hooke's law.

3.11 Other Approaches to Elastic Constant Determination


In actual composite and polycrystalline materials subjected to a homogeneous

stress on their surfaces, both the stress and strain vary from point to point in the
material volume due to the inhomogeneous distribution of elastic constants along
any given direction, as discussed in Sect. 3.8. The actual stress at a point in such
materials will be affected by the shape of the precipitates (through the
components of Eshelby's tensor Sijkl), and the actual distribution of elastic

3.11 Other Approaches to Elastic Constant Determination

67

constants in the composite elements around the point. Thus, the average elastic
moduli linking stress to strain at a point for such a material contain components
from these effects in addition to the elastic moduli Cjjkl,Sjjkl. The effect of such
parameters was determined by Eshelby [3,4,5] through the use of the equivalent
inclusion method (Sect. 3.6) . For the case of a piecewise homogeneous composite
material, with n elliptical inclusions, each with elastic constants (C jjkl ) and
volume fraction fj, where fj is very small compared to the volume fraction of the
matrix fo, the average shear modulus linking the shear stress a jj to shear strain Ejj is
( 3.60)
where Sjjjj is the appropriate Eshelby tensor component and Jlo is the shear
modulus of the matrix. Similarly, the average bulk modulus is given by

KE = Ko

I[ 1 + Jl fj (Ko - Kd I{Ko + i Sjjjj (K j- Ko) } 1

(3.61 )

When the total volume fraction of inclusions becomes comparable to the volume
fraction of the matrix, or when a polycrystalline material is under analysis (in
which case the terms matrix and inclusions become vague) the interaction effects
among particles become more important and formulae (3.60), ( 3.61) can no
longer be used. For such cases the "self-consistent analysis", proposed first by
Kroner [13] may be utilized. Consider a given grain embedded in a random
polycrystalline aggregate subjected to a load on the material surface. The average
stress and strain in the material are crij and Eij respectively. The self-consistent
method assumes that the elastic response of the material surrounding the
crystallite can be represented by a homogeneous material with average elastic
constants (Cjjkl ) B, ( Sjjkl) B, equal to the average constants of the polycrystal. The
stress in the embedded crystal is directly proportional to the average strain
through an equation of the form
aij ( n) = (Cjjkl+ rjjkl ( n) ) Ekl .

(3.62 )

Similarly:
Eij (n) = (Sjjkl + tijkl (n) ) crkl

(3.63 )

Here CjjkbSjjkl are the elastic constants of the single crystal of volume n. The terms
rjjkl,tjjkl are tensors of the fourth rank which describe the interaction of the crystal
having a particular shape and orientation with the matrix. They are related to the
Eshelby tensor Sjjkb and have the same symmetry as C jjkl. If Eqs. (3.62), (3.63)
are integrated ove"r all possible orientations, the average values for the bulk should
be obtained:
crij = (Cjjkl ) BEkb Ejj = (Sjjkl) Bcrkl .

(3.64 )

Therefore:

JD rjjk1dD = JD tjjk1dD = 0 .

(3.65 )

68

3 Analysis of Residual Stress Fields Using Linear Elasticity Theory

For an anisotropic spherical crystal with cubic symmetry and elastic constants
C jjkl (0) in an isotropic medium, the following relations are given between
rjjkbtjjkl (Kroner [13]):
rjjkl (0) = (Cjjkl)B +Cjjmn (0) +umnkl (0)
tjjkl (0) =ujjmn (0) (Smnkl)B

(3.66)

Ujjkl(O) = - Vjj~n(O) [Cmnkl(O) - (Cmnk1h]


Vjjkl(O) =Cjjkl(O) - (Cjjklh+CjjmnWmnkl,
where Wmnkl is a fourth rank tensor dependent only on (Cjjkl ) B:
_ Wllll-W1122
W
1212 2

5 K+3~K

4 K+6~K'

(3.67)

where K and ~K are the bulk and shear moduli of the matrix. Since for cubic
crystals K is equal to the single crystal bulk modulus [see the derivation for Kv
given in the V<Jigt assumption, Eq. (3.50)], it is known exactly. The average shear
modulus ~K can be obtained from the integration of Eq. (3.62) or (3.63).
In later work Kroner [14] derived the average elastic moduli of a perfectly
random composite material from the local elastic moduli using correlation
functions. For an aggregate of cubic crystals the K value is equal to that for the
single crystal, as discussed above. The shear and bulk moduli are given by
( 3.68a)

a
x

3K+6~v
3K+4~v

5S 1111 (0) -S1122 (0) -2S 1212 (0)


3S 1111 (0) -3S 1122 (0) +4S 1212 (0)

- 1(

KK=3

(3.68b)

1 )- -

Sl1l1 +2S 1122

=KR=Kv

This approach also yields values very close to the self-consistent method [for
example for Cu, the elastic constants calculated from formula (3.68a,b) are
within 1.1 % of the values calculated from the self-consistent method].
In Table 3.2 the bulk averages in the Voigt, Reuss approximations and mean of
Voigt and Reuss values, as well as experimentally determined elastic moduli of
various materia~s, are shown.
Based on these and other similar results, (the fact that the measured values of
elastic moduli usually fall between the values calculated by the Voigt and Reuss
assumptions was first pointed out by Neerfeld [16] in 1942), the arithmetic mean
of the Reuss and Voigt values are frequently used as bulk elastic constants in
practical stress analysis when experimentally determined values, or means of

3.12 Average Diffraction Elastic Constants

69

Table 3.2. Average bulk elastic constants in the Voigt and Reuss limits calculated from
experimentally determined single crystal stiffness coefficients. Experimentally determined bulk
elastic constants and Hill averages
Material

AI

Cu
Au
Fe
Material

Al
Cu
Au
Fe

Single crystal constants (x 10 Pa)


Cll

C l2

C44

'J.lR

fJ.v

10.8
10.7
18.6
23.7

6.22
12.3
15.7
14.1

2.84
7.5
42
11.6

26

26
54
31
89

40

24
74

Experimental values

Calculated average bulk values


(GPa)

'K

.&t

.liy

'va

Vv

78
139
167
173

71
109
69
193

71
144
87
229

0.349
0.369
0.431
0.313

0.348
0.328
0.413
0.280

GPa

Hill Averages

il

ilH

KH

vH

26.5

74
133
166
159

71
118
79
207

0.34
0.35
0.42
0.285

26
47
27.5
81.5

78
139
167
173

71
127
78
211

0.349
0.349
0.422
0.297

43.6~

27.8
80.8

Data adapted from Hill [1]

complicated modeling, are not available. (Measured x-ray values are given in
Appendix F.)
All of the analyses treated above have assumed a random distribution of
elements in the composite and polycrystalline bodies. This assumption is not
always justified in practice, in textured materials for example. Further extensions
of these procedures as well as other methods have been proposed for such cases.
However, a general theory applicable to all cases is still lacking.

3.12 Average Diffraction Elastic Constants


Since in polycrystalline materials, diffraction methods obtain information from a
large number of grains, the strains measured by x-rays are average values. In order
to determine the average stress from these strain values, average elastic constants
are required. However, since diffraction can occur only when the normal of these
grains bisect the incident and diffracted beams (see Sect. 4.9) when monochromatic radiation is used, the averaging is not over all of the grains in the material
volume, but over the particular set of diffracting grains, all of which have a specific
form of lattice directions (hkl) in a particular orientation. Thus, in addition to
the assumptions required as to the stress and/or strain distribution, an added
constraint is required for the theor.etical calculation of the averages from the single
crystal elastic constants. The pioneering work in this field has been done by
Glocker [17] and Moller and Martin [18] in the Voigt and Reuss models. The

3 Analysis of Residual Stress Fields Using Linear Elasticity Theory

70

extension of the self-consistent method was given by Bollenrath et al. [19J. In the
following these results are reviewed.
Since the Voigt method assumes constant strain in all grains, regardless of
orientation, the directional restraint imposed by the diffraction condition does not
affect this model. Thus, the average values obtained from the treatment in
Sect. 3.9, [Eqs. (3.50), (3.51 ) J are used without modification. Expressed in terms
of (1 +v) IE, -viE, which are termed "x-ray elastic constants", since they appear
in the fundamental x-ray stress equation (given in Chap. 4) , the Voigt values are
( 1 +v) = 10S 1212 (Sl1l1 -S1212)
v
3S 1111 -3S 1122 +4S 1212
E
So' (Sl1l1 + 2S 1122 ) + 10S1122,S1212
3S 1111 - 3S 1122 + 4S 1212

(3.69a)

(3.69b)

where
(3.69c)
The Reuss model, on the other hand, assumes constant stress. This means that
grains with different crystallographic directions along the stress axis suffer
different strains, since the elastic modulus along the stress direction is not the same
due to elastic anisotropy. Similarly, the average elastic constants obtained from
grains of a particular orientation will not be equal to those obtained from grains
with a different orientation. This effect necessitates modification of the Reuss
treatment given by Eqs. (3.58a,b) such that the averages are taken over only
certain directions.
For a flat specimen in the normal position on a diffractometer only those
grains which have (hkl) planes with normals in the direction of the surface
normal S3 can satisfy the diffraction condition (Fig. 3.11 ). Thus the average
strain is obtained from
( 3.70)
where do is the unstressed lattice spacing, and d, the lattice spacing of planes
parallel to the surface, is along the S3 direction. This strain is related to the stress
O'jj in Sj by HObke's law:
(3.71 )
where compliances S~3kl are referred to Sj'
The relationship between the true elastic compliances, which are referred to the
axes Cj of the unit cube of the crystal, and the elastic compliances S~3kl is given by
the fourth rank tensor transformation equation
(3.72 )

where aij are the direction cosines between the two sets of axes. The relative
orientations of the two sets are defined by the one direction which is given in both

3.12 Average Diffraction Elastic Constants

71

[hkt]

Fig. 3.11. Definition of the measurement axis for a nat specimen in the
normal focus ing position on a
diffractometer

5,

1 ---

[hk ll

531

I
I

I
I

c,

Fig. 3.12. Relative orientations of the measurement and crystal axes in an x-ray
experiment

sets. This is the surface normal Sj, which is also the normal to the diffracting planes
(hkl) , and is thus a crystallographic direction (Fig. 3.12). It must be noted that
the orientation of C j with respect to Sj is not uniquely determined by the
coincidence of S3 with a lattice direction. The entire set of axes C j can be rotated
around S3 to any arbitrary orientation without violating this condition. Thus, in
order to obtain the average value of S;jkl along Sj, the average of all possible
rotations must be considered.
F or cubic crystals, where the normal to the plane (hkl) is the direction [hkl] ,
the direction cosines between Sj and C j can be written in terms of the indices h,k,l,
since h,k,l are the coordinates of a point on the vector rhkl> connecting this point to
the origin. Thus, writing all components of aij in terms ofh,k,l and then taking the
average of Eq. (3.72) for all possible rotations of C j around <hkl), the Reuss
averages for the elastic constants can be shown to be
( - ; ) R =Sl122 =sor

(3.73a)

(El+V)

(3.73b)

R =SIIII

-S1122 -3S or.

3 Analysis of Residual Stress Fields Using Linear Elasticity Theory

72

The orientation term r is a function of h,k,l only, and is given by


h 2 k 2 + k 2P + h 212
r= (h2+k2+P)2 .
The term So is given by Eq. (3.69c)
A similar treatment is also used in the average diffraction constants evaluated
according to the Kroner model. However, in this case transformation and
averaging of the interaction coefficients [Eq. (3.62), ( 3.63 )] as well as the elastic
moduli, Cjjkl,Sjjkl is necessary. Thus, from Eq. (3.63), one obtains
( - ; ) K = (S3311 h+t3311 +tor
(1

~v) K= (S3333 h-S3311 +t3333 -t3311 -3tor

(3.74a)

to=t3333-t3311-2t3131
Here (SjjkdB':are the average bulk values and are given by
1)
1 (1
(S331dB= 3 3KK - 2JlK
1

(S3333h- (S3311 h= ---= .


2JlK

(3.74b)
(3.74c)

The average shear modulus ilK and average bulk modulus KK in the Kroner limit
are given by Eqs. (3.68a,b) respectively. For single-phase polycrystalline
materials with random grain orientations, the interaction tensor components
t 3333 , t 3311 , t 3131 are given by
(3.75)

1
where t3311 =t 3131 = - 2t3333' and Sjjkl(Q) are the single-crystal elastic
constants.
Of the three approaches discussed above, the average diffraction constants
calculated in the Kroner limit are closest to experimentally determined values. The
median of the Voigt and Reuss values (the Neerfeld-Hilllimit) is also close to the
Kroner limit, and can also be used in practice. For example, for the 211,310
reflections usually employed in residual stress determination in iron or steels, the
mean values are within 5 % of the Kroner values. All of the approaches discussed
above break down however, when non-random orientation distributions
(texture) exist within a material. X-ray elastic constants also exhibit variations
with plastic deformation in certain cases; thus it is usually safer to use
experimentally determined elastic constants. The experimental determination of
elastic constants with diffraction methods will be discussed in detail in Chap. 5.
Some typical values are given in Appendix F.

References

73

Summary
Inhomogeneous partitioning of inelastic strains (such as plastic or transformation strains) over macroscopic volumes of a given specimen causes the formation
of residual macrostresses. These stresses arise due to the mutual constraint of such
regions. Macrostresses are homogeneous over macroscopic dimensions that are
larger than the grain size of a common engineering material.
Residual microstresses may be caused by inhomogeneous partitioning of
plastic deformation on a micro scale (microstresses due to differential plastic
deformation) or, in the presence of an applied load, by inhomogeneous
partitioning of elastic constants on a micro scale (microstresses due to elastic
incompatibility). These stresses are distributed inhomogeneously on a macro
scale.
Pseudo-macro residual stresses are the average microstresses where the
average is taken over a statistically representative volume. As long as external
parameters (such as the deformation distribution and/or the applied load) are
constant in the ~aterial volume, the PM stress is independent of the position of
the averaging volume.
Macro- and micro residual stresses must obey the equations of equilibrium at a
point (3.7). PM stresses on the other hand obey the average equation of
equilibrium (3.14) when the stress values are determined from mutually-exclusive
subsets of the same volume such as matrix and precipitates.
Calculation of micro- and PM residual stresses are possible for elliptical
precipitates through the equivalent inclusion method of Eshelby. Such formulae
are generally too restrictive for real specimens with random precipitate shapes.
However, analysis of the solutions yields valuable information on the possible
contributions to the total residual stress tensor.
The macroscopic average elastic constants of a polycrystalline solid may be
calculated from the elastic constants of the constitutive single crystals using
various assumptions such as the Voigt limit (constant strain in all grains) , or the
Reuss limit (constant stress in all grains) . Hill and Neerfeld have shown that the
measured values ate very close to the mean of the Voigt and Reuss limits. The
Kroner calculations for anisotropic precipitates coupled to an isotropic matrix
also yield values close to this mean. The same approximation is also valid for the
average constants applicable to diffraction, which take into account the added
constraint imposed by the diffraction conditions. If there is preferred orientation,
none of these approaches may be valid.

References

2
3

Toshio Mura, "Micromechanics of Defects in Solids", Martinus Nijhoff Publishers, The


Hague, Netherlands, 1982
S.P. Timoshenko and J.N. Goodier, "Theory of Elasticity", McGraw-Hill, New York, 1970
J.D. Eshelby, Proc. Roy. Soc. A241, 376-396 (1957)

74

3 Analysis of Residual Stress Fields Using Linear Elasticity Theory

4
5

J.D. Eshelby, Proc. Roy. Soc. A252, 561-569 (1959)


J.D. Eshelby, Progress in Solid Mechanics, 2, eds. LN. Sneddon and R. Hill, North Holland,
Amsterdam, 89 - 140 (1961)
K. Tanaka and T. Mori, Acta. Met. 18,931-941 (1970)
T. Mori and K. Tanaka, Acta. Met. 21, 571-574 (1973)
L.M. Brown and W.M. Stobbs, Phil. Mag. 23,1185-1233 (1971)
L.M. Brown, Acta. Met. 21 879 - 885 (1973)
W. Voigt, Lehrbuch der Kristallphysik, Teubner, Leipzig/Berlin (1928)
A. Reuss, Z. Angew. Math. Mech. 949-58 (1929)
R. Hill, Proc. Phy. Soc. A65, 349-354 (1952)
E. Kroner, Z. Physik 151, 504-518 (1958)
E. Kroner, J. Mech. Phys. Solids 15, 319-329 (1967)
J.F. Nye, "Physical Properties of Crystals", Clarendon Press, Oxford, 1976
H.Neerfeld, Mitt. K.W.L Eisen. Dusseldorf, 24, 61-70 (1942)
R. Glocker, Z. Techn. Phys. 19, 289-293 (1938)
H. Moller and G. Martin, Mitt. K.W.I. Eisen. Dusseldorf, 21, 261-269 (1939)
F. Bollenrath, V. Hauk, E.H. Muller, Z. Metallkde. 58, 76-82 (1967)

6
7
8
9
10
11
12
13
14
15
16
17
18
19

4 Fundamental Concepts in X-ray Diffraction

4.1 Introduction
Diffraction methods of residual stress determination basically measure the angles
at which the maximum diffracted intensity occur when a crystalline sample is
irradiated with x-rays or neutrons. From these angles one then obtains the spacing
of the diffracting lattice planes by using Bragg's law. If the material is under load,
these values will be different than the unstressed plane spacing, and the difference
will be proportional to the stress acting on the planes. At this point one can use
elasticity theory, lis discussed in Chap. 2 and 3, to determine the stress (residual or
applied) acting on these planes. Thus, no matter how sophisticated the elasticity
analysis, the final stress results are only as good as the data supplied by the
diffraction methods.
In this chapter the theory and instrumentation for x-ray diffraction required in
residual stress determination will be discussed in detail. These concepts, along with
those discussed in Chaps. 2 and 3 will then be used in Chap. 5 to derive the
equations of residual stress analysis. In addition, the pertinent concepts for
neutron diffraction, which is also used in residual stress analysis, will be discussed
briefly. It must be noted, however, that even though the pertinent topics are
reviewed as completely as possible, the treatments given are by no means
exhaustive, and the reader is urged to consult the standard textbooks on the
subject, such as Cullity[l], Schwartz and Cohen [2] , Azaroff[3], Barrett and
Massalski [4], Klug and Alexander [5] and Bacon [6], when necessary.

4.2 Fundamentals of X-rays


In 1895 German physicist W. C. Roentgen discovered that when a high enough
potential was applied across two metal plates placed in an evacuated tube,
invisible rays of an unknown nature (x-rays) were produced. Experiments
showed that these rays travelled in straight lines and could expose photographic
film, similar to visible light. In addition, x-rays penetrated materials opaque to
visible light; photographic film could be exposed even when its protective cover
was still in place. The amount of exposure however was inversely proportional to
the thickness of the protective material. This property of x-rays is utilized in
radiographic analysis of materials where x-rays are allowed to pass through a
specimen before impinging on photographic film. Any local variation in density or
thickness, such as cracks, voids, embedded foreign material etc., can then be seen

76

4 Fundamental Concepts in X-ray Diffraction

on the fIlm due to the contrast produced by these features in the transmitted x-ray
intensity. Generally, features larger than a millimeter can be resolved, and the
technique is mainly used in (relatively) macroscopic defect studies, as well as in
medicine.
In 1912 von Laue established the electromagnetic wave nature of x-rays by
showing that they are diffracted by crystalline materials. These experiments
proved that the diffracting medium is, in fact, the periodic atomic structure of the
material, and that the wavelength of x-rays is of the same order as the lattice
spacing of materials. Further study by W. L. Bragg, who derived a simple, elegant
set of laws relating diffracting phenomena to crystal structure, facilitated the use
of x-ray diffraction in the study of the structure and behaviour of materials on a
very fine scale, on the order of the dimensions of the unit cell (10 - 7 mm).

4.3 Short-wavelength Limit and the Continuous Spectrum


X-rays may be produced by two fundamentally different methods. The first
method utilizes the fact that all electrically charged particles emit radiation during
rapid deceleration. This is discussed here. The second method involves ionization
and is covered in Sect. 4.4. In the case of interest to us, electrons are used as the
charged particles. These electrons are usually produced by heating a filament
(usually tungsten) in vacuum. They are then accelerated to high velocities by
placing them in an electrical field, for example between two metallic electrodes
across which a high potential has been applied, and then decelerated rapidly by
directing them into the positive electrode (the target). At the time of impact a
small part of the kinetic energy of the electron is given off as x-ray photons. The
main bulk of the kinetic energy however is converted into heat. This usually
necessitates some sort of cooling of the anode.
If the potential difference between the plates is Y (in esu ) , the kinetic energy of
the electrons at the time of impact is
1

lmv =eY,

(4.1 )

where m is the mass of the electron and v is its speed just before impact. For a given
voltage, Eq. (4.1) yields the maximum kinetic energy of the x-ray photons
produced during an impact in which the incoming electron loses all its kinetic
energy in a single collision. Using wave-particle duality, the energy of the x-ray
photon may be expressed as
(4.2)

eY =hc/A,

where A is the wavelength, h is Planck's constant and c is the velocity oflight. It can
be seen that a given acceleration voltage cannot produce x-ray radiation with
wavelengths shorter than a limiting value, the short-wavelength limit, \WL' From
Eqs. (4.1) and (4.2):
Amin = ASWL =

12.400
-Y-

(4.3 )

77

4.4 Characteristic Radiation Lines

Intensity

Intensify

28
a

or

28-or

--?-

A~

A---

Fig. 4.1. Variation of intensity vs. acceleration voltage (a), or vs. filament current (b)

Here Y, the applied potential difference, is in practical units (volts). ASWL is calIed
the short-wavelength limit and represents the most energetic photons possible at a
given acceleration voltage. On the other hand, an incoming electron may not give
up alI of its energy with a single impact. In such a case only a portion of the total
energy is available for x-ray photons. These impacts cause producton of lower
energy (longer wavelength) x-rays. For a given voltage this effect causes the
presence of a continuous spectrum of wavelengths in the x-ray beam produced by
deceleration. This type of x-ray radiation is calIed polychromatic, white or
bremstrahlung (slowing down) radiation. In Fig. 4.1 a,b the intensity vs. wavelength plots are shown for an x-ray tube operated at various accelerating voltages
( a) and filament currents (b). Increasing the filament current simply increases
the total intensity at alI energy levels since the number of impinging electrons
increases. However, the energy distribution itself is a function of acceleration
voltage only and does not change with filament current.
In addition to these parameters the total x-ray intensity (the area under a
given curve in Fig. 4.1) is also dependent on the atomic number Z of the target
material:

(Icont.spect.dA)

cdZy n .

(4.4 )

A.SWL

Here i is the filament current, Y is the accelerating voltage and n varies between 1.5
and 2 for different targets.

4.4 Characteristic Radiation Lines


If the accelerating voltage applied to an x-ray tube is increased above a certain
voltage, sharp intensity maxima superimposed on the continuous spectrum
appear (Fig. 4.2). These maximum intensity lines are called characteristic
radiation lines. Both the minimum voltage required for their appearance (the
excitation voltage) and the wavelengths at which these maxima appear are
characteristic of the target (anode) bombarded by the electrons. The origin of the

4 Fundamental Concepts in X-ray Diffraction

78

: 4 1---+-II---lo-od-+--+--+----1
Q)

...=.
III

c:

Q)

C 2 I---++--++---."..~~-'k---+---I
:>.

~ I l---tH'--I----b-T--<.po~~~-I
I

SWL

O~~~~~~~~~-~-~

Fig.4.2. Variation of the total output of a


sealed x~ray tube with accelerating potential
(V)

Wavelength (angstroms)

characteristic spectrum can be traced to the energy of various electronic shells


around the nucleus. Consider Fig. 4.3. If the incoming electron is energetic enough
it can eject an electron from a K, or 1s, shell, and excite the atom to a high energy
state. The atom then attains a lower energy state by filling this vacancy by an
electron from a higher energy shell (L or M ) . The difference between the electron
energies of die K shell and the higher energy shell is emitted in the form of
characteristic x-rays during this process. Since the electron shell energies are
quantized, the energy, and thus the wavelength associated with the emitted
radiation for a given transition is unique. Thus characteristic lines appear at
definite wavelengths. Furthermore, since the electronic bonding energies are
dependent on the atomic number, Z, the wavelength for a given transition is a
characteristic of the target material. This fact was discovered by Moseley in 1914
who showed that there is a linear relation between the square-root of line
frequency (v) and the atomic number:
Vv=C'(Z-o') ,

(4.5 )

where C and 0' are constants for a given transition and material. It can be seen
from Eq. (4.5) that characteristic radiation lines occur at wavelengths independent of the acceleration voltage. For these lines the x-ray intensity is proportional to
the filamentcuqent, i, and acceleration voltage, V, through an equation of the form
ICh. Rad = B+ (V - VChar. ) n ,

( 4.6)

where B is a constant and VChar. is the excitation voltage of the characteristic line.
The characteristic lines are named according to the primary excited shell and
the shell that supplies the electron required for the decay to the lower energy state.
For example, a Kline is produced when an electron is ejected from the K shell and
the L shell supplies the electron to it. The L shell however has two electrons with
different spins and depending on which electron is supplied one can have a Kcx 1 or
Kcx2 line. If the electron is supplied by the M shell the line is termed a KII line. In
Fig. 4.4 the relative position and intensities of these lines are shown schematically.

79

4.4 Characteristic Radiation Lines

..!.. Imll

n
4

Ionization
Level

4s
3d
K -excltahon

L -excitation

3p

Ll[

Lm

2
2

Lx

3s

I
0

2p }

2s

is

L Shell

K Shell

K state (K electron removed)

II AK
u

.c
'-

c:

.9

E
.,~

EllA
<lIlA

'"

L"

'0 IIX- ~
Lm

'"'
c

II

II

Kil
emiSSion

L state (Lelectronremoved)

La

-'

AM

II AN

Q)

""
"''''

1M IMa
~

M state (M electron removed)


N state (N electron removed)
Valence electron removed
Neutral atom

Fig. 4.3 a, b.
Electronic
transitions causing emission of characteristic radiation. a Electron energy
levels of Cu metal with associated
quantum numbers n,l,m. the arrows
indicate the direction of the electron
transition that causes the excitation
to a K state, L state etc., indicated
in b

It must be noted that if the electron beam can excite K characteristic radiation
it will also excite all the other, less energetic, lines. However usually the K lines are
used in x-ray diffraction because of their high penetration through air. For targets
with moderate atomic numbers, like CU,Fe,Cr, lower energy lines are strongly
attenuated in air through scattering and are usually not observed. There are,
I

................ .

Ka ,

Intensity
0.5 ................ .

0.16-0.25

Kp

28-

Fig. 4.4. Relative wavelengths and intensities of


characteristic lines

80

4 Fundamental Concepts in X-ray Diffraction

however, important L lines from targets like Pt,Au,W, etc. In fact, for these
elements they are the only lines produced by the usual voltages (20 - 60 kV)
encountered in standard diffraction units.

4.5 X-ray Sources


Currently there are three widely used sources of x-rays: sealed tubes, rotating
anode generators and synchrotron radiation sources. These sources are listed in
the chronological order of development, x-ray intensity and cost. Since the higher
the incident beam intensity, the faster a given experiment can be performed, the
latter two sources are attractive and sometimes essential. On the other hand,
constraints such as portability and cost can also restrict the selection of an x-ray
source, in which case a sealed tube is the obvious choice.
Sealed tubes and rotating anodes both produce x-rays by accelerating
electrons (produced by heating a tungsten filament in vacuum) through a high
potential field: and then directing them into a target, which then emits x-rays. The
emitted radiation contains both the continuous and characteristic spectrum, as
discussed above. This process also generates a considerable amount of heat in the
target, and it must be continuously cooled during operation. In a sealed beam tube
the target is stationary, and is coupled to a heat sink which, in turn, is usually water
cooled (Fig. 4.5), although there are portable tubes that are air cooled. For these
focusing. plates
gloss

target

current
for
filament

gloss
to metal
seal
W filament

Fig. 4.5. a Schematic of a sealed xray tube. b An actual tube

4.5 X-ray Sources

81

Energy (keV)
1016 . - . -_ _-----.1o
___----,-1_ _ _ _,.1---.,
6 pole
supercon. wiggler
_ - - __

/I.e

---- ----

'}/

/'

Arc source

X-roy ring

I
i

0,1

Fig. 4.6. Intensity vs. wavelength plot of


the x-ray emission from the Brookhaven
synchrotron x-ray ring
1

10

100

Wavelength (A)

tubes the maximum power is limited by the amount of heat that can be dissipated,
and is usually of the order of 1.5 - 3 kW. The direct beam intensity at the sample
for such tubes is around 109 counts per second. In rotating anode generators the
target is cylindrical in shape and is hollow. The coolant is circulated through this
interior cavity. During operation the target is rotated at high speed, such that the
beam impinges on new material all the time, facilitating much greater heat
dissipation. With these generators, beam currents of up to 2000 rnA, (compared to
40 - 60 rnA for sealed tubes) can be used, resulting in beam intensities around
10 10 cps at the sample. The targets for rotating anodes are demountable, so that
different anodes can be used to produce characteristic radiation at desired
wavelengths.
In synchrotron., radiation sources a packet of high energy electrons is
accelerated to speeds close to the speed of light in a circular "ring". The
centripetal acceleration due to the curvature of the path causes the electrons to
emit electromagnetic radiation. Depending on the energy of the electrons, and the
local radius of curv~ture at any point in the ring, the emitted radiation will have a
range of wavelengths, and thus will be essentially white radiation. It is also
possible to insert electromagnetic devices (undulators or wigglers) locally into
the ring, and thus enhance the production of a certain range of wavelengths. In
Fig. 4.6 the wavelengths and the intensity distribution that make up the white
radiation spectrum at the Brookhaven Synchrotron X-ray ring is shown. This
particular ring is 170 meters in circumference and operates at 2.5 Ge V stored
energy and 200 rnA. It can be seen that the x-ray intensities are much larger than
conventional sources. Also, in contrast to the tubes which produce high intensities
only at the characteristic wavelengths, synchrotron radiation has much higher
intensities over a large wavelength range, facilitating the selection and use of a
number of wavelengths during a given experiment.

82

4 Fundamental Concepts in X-ray Diffraction

4.6 Absorption of X-rays


X-rays are attenuated as they pass though matter, so that the transmitted beam is
weaker than the incident beam. Various processes contribute to this decrease in
direct beam intensity. Part of the beam may be lost through coherent scattering,
incoherent scattering, activation of fluorescent x-rays, production of heat or
excitation of photoelectrons, etc. The total loss in intensity due to the sum of all
these processes is termed absorption. The total attenuation is related to the
thickness dz of an infinitesimal slab by the relation

dl

T = -~dz.

( 4.7)

Here I is the transmitted beam intensity, and (~), the proportionality constant, is
called the linear absorption coefficient. It is proportional to the density (Q) of the
absorber and is ll;sually listed in the tables as (~/Q). This ratio is called the mass
absorption coefficient. The term (~/Q) is a material property and is independent
of the phase Of the material (solid, liquid or gas). Mass absorption coefficients of
some elements are given in Table 4.1. It can be seen from this table and Eq. (4.7)
that the incident beam is attenuated typically to 1/1001h of its value in the first 20
microns of a solid sample. Tabulated values of ~/Q for various elements and
radiations are listed in the International Tables for X-ray Crystallography
[7,vo1.3].
For a homogeneous slab of finite thickness z Eq. (4.7) can be integrated to yield
( 4.8)

where 10 is the incident beam intensity.


The mass absorption coefficient of a mixture, consisting of n substances, each
with weight fractions Wi can be calculated from
n

(~/Q)I=

L Wd~/Q)i'
i= 1

( 4.9)

With the wavelengths of interest in diffraction, the largest contribution to


absorption is the activation of fluorescent x-rays, which is also termed "true"
Table 4.1. Mass absorption coefficients and densities of selected metals for commonly used
radiations
Metal

Density
(gmjcm 3 )

Mo Kex
(1..=0.711

Aluminum
Iron
Nickel
Copper
Titanium
Chromium

2.70
7.87
8.90
8.96
4.54
7.19

5.30
38.3
47.7
49.7
23.7
30.4

A)

Cu Kex
(1..= 1.542
48.7
324
49.3
52.7
204
259

A)

Co Kex
(1..= 1.790
73.4
59.5
75.1
79.8
304
392

A)

Cr Kex
(1..=2.290
149
115
145
154
603
89.9

A)

4.6 Absorption of X-rays

Kedge

t:. edges

83

Fig. 4.7. Variation of absorption coefficient with the


wavelength of incident x-ray radiation

A-

absorption. In this process the incoming x-ray photons are active enough to eject
electrons in the electron shells around the nucleus. This causes the atom to be
excited to an activated state, and it decays to the ground state by supplying an
electron from one of the higher energy electron shells, emitting characteristic xrays for the particular transition in the process. These characteristic x-rays are
identical to the x-rays produced by accelerated electrons as discussed in section
4-4. However, here they are caused by the incident x-ray beam and are called
fluorescent radiation. Similar to the accelerated electrons, the incoming x-ray
photons require ~ minimum energy before they can activate fluorescence.
For lower energies (longer wavelengths) attenuation is due, in large part, to
absorption processes involving scattering. As the energy increases to the minimum
energy for the excitation of K electrons for example, there is a sudden increase in
the beam attenuation. At this point true absorption, in addition to scattering
processes, is taking place, with the irradiated material producing K fluorescent
radiation, and there is a large increase in the absorption coefficient. This point is
called the K absorption edge. As the incident beam energy further increases, the
absorption coefficient gradually decreases. This is so since, even though scattering
and true absorption mechanisms are still operating, high energy x-ray photons
have a higher probability of passing unaffected through the material. The
variation of mass absorption coefficient with wavelength ( or energy) of incoming
x-rays is shown schematically in Fig. 4.7. Between the absorption edges, absorption depends on the atomic number Z of the material:
( 4.10)

Here A is the wavelength of the incident x-rays and K is a proportionality constant


which is different for each branch of the curve in Fig.4.7.
Since flourescent radiation is characteristic radiation, its wavelength depends
only on the type of atom producing it, and one can identify the element by
determining this wavelength. This is the basis ofx-ray fluorescence analysis, where
a substance is irradiated with high energy x-rays, and the resulting fluorescent
radiation analyzed for the different wavelengths, and hence the different elements,
contributing to it. In normal diffraction experiments however, fluorescent
radiation drastically increases the background radiation level and can, at times,
totally swamp the signal from the diffracted beam. In such cases one may change
to a less energetic incident beam (from a lower Z target) that cannot excite
fluorescence from the sample, or use a special x-ray detector and associated
electronics (which will be discussed later in this chapter), that can reject the
unwanted wavelengths, thus minimizing the background.

4 Fundamental Concepts in X-ray Diffraction

84

4.7 Filtering of X-rays


In Sect. 4.4 it was shown that the x-ray spectrum contains the white spectrum and
the characteristic lines, especially K CX" K cx2 and K~. In most x-ray experiments it is
desirable to have monochromatic Kcx radiation. This may be achieved by placing a
ftlter material, which has its K absorption edge between the Kcx and K~
wavelengths of the x-rays, in the direct beam path. Thus most of the energy ofK~
photons will be spent exciting fluorescent radiation from the ftlter, while the Kcx
lines will be attenuated much less by the scattering processes. This is shown
schematically in Fig. 4.8.

28...,,-

28-

Fig. 4.8 a,b. Unfiltered and filtered output of an x-ray tube


Table 4.2. Kp filters for commonly used radiations
Target
Molybdenum
Zinc
Copper
Nickel
Cobalt
Iron
Manganese
Chromium

(Atomic no.)
( 42)
(30)
(29 )
(28 )
(27 )
(26 )
(25 )
(24 )

Filter
Niobium
Copper
Nickel
Cobalt
Iron
Manganese
Chromium
Vanadium

(Atomic no.)
(41 )
( 29)
(28 )
(27 )
(26 )
(25 )
(24 )
(23 )

In Table 4.2 ftlters for some commonly used materials are shown. It may be
noted that for all cases, the atomic number of the filter material is 1 or 2 less than
that of the target metal.
Another important parameter in x-ray filters is the thickness of the filter foil.
Too thick a filter foil will cause unnecessary intensity loss. Usually foil thicknesses
that decrease the K~ to 1/5001h of its original intensity (which will reduce Kcx
intensity by 50 % ) are used in practice.

4.8 Scattering of X-rays


When an x-ray beam, containing x-ray photons of wavelength A, impinges on a
specimen, the photons collide with the fundamental particles (electrons and
nuclei) making up the specimen, and bounce off (scatter) in all directions.
Practically all of the collisions occur with electrons, since the nuclei are, in effect,

4.8 Scattering of X-rays

85

shielded by the electrons surrounding them. It is possible to have two types of


collisions with the electrons: If the electron is in one of the inner orbitals and is very
tightly bound to the nucleus, the collision will be elastic; that is, there will be no
momentum transfer between the electron and the photon, and hence, the scattered
photon will have the same energy and wavelength after the collision as before.
Such scattering is called coherent scattering since, if the wave nature of the x-ray
beam is used to analyze the scattering process, it can be shown that there is a
definite relation between the phases of the incident and scattered x-rays. If the xray photon, on the other hand, collides with a weakly bound electron in the outer
orbitals of an atom, there will be some momentum transferred from the photon to
the electron (i.e., the electron will move under the imp;lct) , with a corresponding
decrease in the energy of the scattered photon. Thus, the scattered photon will
have a longer wavelength than the incident photon. This type of scattering is called
Compton modified scattering. Another important property of Compton scattering is that there is no definite phase relationship between the incident and scattered
beams, since the amount of momentum transferred, and thus the wavelength of
scattered radiation depends on the binding energy of the particular electron
involved in a giveil collision. For this reason, Compton scattering is also called
incoherent scattering.
In both of the above processes electrons scatter the photons in all directions.
However, the intensity of the scattered beam depends on the scattering angle. If an
unpolarized x-ray beam of total intensity 10 impinges on a single immovable
electron, the total coherent scattering at a point P is given by the Thomson
equation. In cgs units:
I = I . ~ . ( 1 + cos 2 29 )
o r 2 m 2 c4
2
.

( 4.11 )

Here m is the electron mass, c is the speed of light and e is the electron charge. The
term r is the length of the position vector to P, and 29 is the angle between r and the
incident beam direction (Fig. 4.9). The term (1 +cos 2 29), which describes the
variation of scattered intensity with scattering angle is called the polarization
factor. It can be seen that, due to this term, the scattered intensity in the forward
(29 = 0 0 ) and backward (29 = 1800 ) directions will be stronger than directions
90 to the incident beam 1 .
Equation (4.11) gives the scattered intensity from a single electron. When, on
the other hand, the x-ray beam impinges on an atom, with atomic number Z, the
total scattered intensity to a point P in space will be slightly different. In this case
there will be Z electrons, and, therefore Z (scattered) beam sources at different
positions around the atom. Therefore Z rays, all with wavelength A. will arrive at a
point P away from the atom. However, the total scattered intensity at P, It, will
depend on the relative phases of the individual rays arriving at the point. If all the
beams are in phase [i.e., 29 = 0 in Eq. (4.11 ) for all the electrons in question] , the
It must be noted that this polarization factor is applicable only when the incident beam is

unpolarized. If the incident beam is polarized, as in the case of synchrotron radiation, there is
no polarization correction term to the scattered intensity. The beam may also be partially
polarized, if a single wavelength is chosen by, say, diffraction from a "monochromator"
( crystal) prior to impinging on the sample.

86

4 Fundamental Concepts in X-ray Diffraction

E
X1

Fig. 4.9. Geometry for the coherent scattering by a single electron to a point P

total intensity is Z2 '(le), where the intensity Ie is given by Eq. (4.11). For any
other direction, the beams from individual electrons will be out of phase by
various amounts and a certain amount of destructive interference will take place,
causing a reduction in the intensity scattered into non-zero 29 angles. It can be seen
that only coherent scattering, where all the beams have the same wavelength and
definite phase relationships with each other, can take part in this process.
Compton modified scattering, due to its random wavelength distribution and
arbitrary phase relationships, cannot take part in constructive or destructive
interference and contributes only to the overall background.
The term used to describe the total coherent scattering from an atom to a point
P is "r', the atomic scattering factor, which is defined as
f = amplitude of wave scattered by an atom
amplitude of wave scattered by an electron

( 4.12)

As indicated above, fis equal to Z at 29=0, and decreases with increasing 29. If
the total diffracted intensity at a point P is calculated, it is seen that f is, in fact, a
function of sin 9/1.., where 29 is the angle of scattering and Ais the wavelength of
the incident radiation. Further information, as well as tabulated values on the
atomic scattering factors of common ions can be obtained from the International
Tables for X-ray Crystallography [7, volume IV, pp 71-146].

4.9 Scattering from Planes of Atoms


In real materials of finite volume there are, of course, a great many atoms. In single
cry.stals_ these atoms are arranged in periodic planes, as discussed in Sect. 2.8.
Wiien~an x-ray beam impinges on the material, all of the atoms in the irradiated
volume will scatter x-rays in all directions. However, the periodic distribution of
atoms on planes may cause constructive, or destructive, interference of the
coherent component of the scattered radiation from the individual atoms,
depending on the spacing between the planes, and the wavelength of the incident
radiation.
Consider what happens when a beam of x-rays, where all rays are parallel,
impinges on a crystal where the (hkl) planes are parallel to the surface
(Fig. 4.10). The waves ABC and DEF will be in phase, and so mutually reinforce

87

4.10 The Structure Factor of a Unit Cell


F

Fig. 4.10. Diffraction of x-rays by a crystal

each other if the path difference GEH between them is an integral multiple (n) of
the radiation wavelength A.. Since
GE=EH=dsin9.
Constructive interference, that is, diffraction will occur when
nA.=2d sin 9.

(4.13 )

Equation (4.13) was first derived by British scientist W. L. Bragg in 1912, and is
called Bragg's laW:
In the case of an actual three-dimensional crystal of finite dimensions, there
are a very large number of atomic planes. Thus there are, in fact, a large number of
mutually reinforcing beams causing appreciable (diffracted) x-ray intensity at
the correct angles. Along directions that do not satisfy Bragg's law, however, the
scattered waves are out of phase, resulting in destructive interference and no
diffracted intensity. In addition there will, of course, be background radiation at
all points, caused by inelastic (Compton) scattering and, if the incident beam is
sufficiently energetic, fluorescent radiation.

4.10 The Structure Factor of a Unit Cell


In the derivation of Bragg's law, the three-dimensional symmetry of the unit cell
was not taken into account. This causes the angular relationship established by
Bragg's law to be a necessary but not sufficient condition for diffraction. If Bragg's
law is not satisfied no diffraction can occur. However, in certain cases there is no
diffracted intensity at angles predicted by Bragg's law. This effect is due to the
partiCUlar positions of the atoms in the unit cell. Consider, for example, diffraction
from the (100) planes of a body-centered crystal. The unit cell in this case is
shown in Fig. 4.11 a. It can be seen that between the two (100) planes there is
another plane of atoms, which corresponds to the atoms at the body center;
1/2,1/2,1/2 (Fig. 4.11 b). In this configuration, when the phase difference between
rays 1 and 3, ABC, is one wavelength, the rays from the atoms on planes 1 and 3
can reinforce each other. However, the phase difference between rays 1 and 2,
DEF is half a wavelength. Thus, ray 2 will interfere destructively with ray 1,
resulting in zero net diffracted intensity, even though the rays are impinging on the
surface at the correct Bragg's angle for diffraction.
As illustrated by the example above, the particular arrangement of atoms on
planes can cause dramatic changes in the diffracted intensity, even though Bragg's

88

4 Fundamental Concepts in X-ray Diffraction

3'

,,

a
b
Fig. 4.11. a The unit cell of a body-centered crystal. b diffraction from the (100) planes of a body
centered cell

law is satisfied. Of course, total destructive interference is not always the outcome.
However, in general, the diffracted intensities vary considerably, depending on the
particular unit cell under consideration.
The total intensity diffracted by a given unit cell can be determined from the
summation of the waves from the individual atoms. If the scattering amplitude
atom, with fractional dimensionless coordinates Uj,Vj,Wj' is fj' the
from the
relative intensity scattered by the unit cell for a given reflection hkl is proportional
to

(4.14 )
Here the summation is over all the atoms present in the unit cell. The term Fhkl is
called the structure factor and represents the sum of the amplitudes and phase
differences of the waves scattered by the individual atoms. In general F is a
complex number, containing real and imaginary parts, and the magnitude of the
total scattered intensity is proportional to the square of the absolute value of F,
1F12. In Table 4.3, the structure factors for some simple Bravais lattice types are
summarized. The reader can derive these from Eq. (4.14).
From the discussion given above, it can be seen that when a perfectly
monochromatic, parallel beam of x-rays impinges on an infinite single crystal that
is perfect throughout its volume, diffraction can occur if and only if:
a) Bragg law 'nA. = 2d sin e is satisfied for one or more sets of crystallographic
planes (hkl),
Table 4.3. Structure factors and allowable reflections for selected lattices
Bravais
lattice

Structure factor

Reflections
present

Reflections
absent

Simple
Body-centered
Face centered

F=f
F=f( 1 +e'i(h+k+l
F= f( 1 +e'i(h+k) +e'i(h+l) +e'i(k+I

all
(h+k+l) even
h,k,l unmixed

none
(h+k+l) odd
h,k,l mixed

89

4.11 Experimental Utilization of Bragg's Law

69
29

28--

Fig.4.12. a Variation of intensity


with diffraction angle predicted by
the Bragg law. b Line broadening
caused by deviation from the theory

b) The structure factors F for these hkl reflections are non-zero for the unit cell
of the particular crystal.
Under these conditions Bragg's law predicts that the variation of diffracted
intensity I with incidence angle 29 can be described by a step function (Fig. 4.12a ) .
However, the ideal conditions assumed in the derivation do not exist in reality. The
beams produced by x-ray tubes are polychromatic and even if a flltered
characteristic line is used, there is still some white radiation left in the beam and the
characteristic line itself has a range of wavelengths (about 0.001 A). Furthermore,
the beam contains parallel, divergent and convergent rays. The crystal itself is
rarely "perfect", but contains mosaic structure, that is, small regions within the
volume which are slightly misoriented with respect to each other. In addition the
size of the crystal is finite, and in some cases, very small. In certain cases, there can
be elastic strains and stresses distributed throughout the volume of the crystal. All
these effects cause "broadening" of the diffracted beam. That is, the diffracted
intensity is observed over a range of 29 (Fig. 4.12b ), with the maximum intensity
occuring at the exact Bragg angle. The width of the peak at half the maximum
intensity is proportional to the deviation of the experimental set-up and specimen
from ideality. Such broadening can be analyzed to yield information about the
mosaic size and strain distribution within the irradiated volume, and will be
treated in detail in Chap. 8.

4.11 Experimental Utilization of Bragg's Law


For planes with non-zero structure factors, Bragg's law, nA = 2d sin 9 can be used
in two modes: By using a monochromatic beam of x-rays of known wavelength,
and measuring the angle of diffraction 29, one can determine the spacing of the
diffracting planes (dhkl ). This technique is the basis of structure analysis and of
residual stress measurement with x-rays. Alternatively, one can determine the
wavelength of the incident radiation, if the Bragg angle 29 is measured, and d hkl is
known. This technique, known as x-ray spectroscopy, is used in analyzing
flourescent radiation, and can also be used to obtain a monochromatic x-ray beam
to be used in structure analysis. Crystals used for this purpose are called crystal
monochromators.

90

4 Fundamental Concepts in X-ray Diffraction

Structure analysis methods, including residual stress/strain analysis, are based


on Bragg's law of diffraction, and thus contain a number of implicit assumptions.
These can be summarized as follows:
1) The x-rays are treated as traveling waves.
2) The path difference between the waves arriving at a point is a linear function of
atomic plane spacing.
3) The scattering process is elastic and energy is wholly conserved during
scattering, resulting in diffracted waves with the same wavelength as the
incident waves.
4) The scattered waves within the material, including the diffracted beam, do not
rescatter.
These assumptions form the basis of the "kinematic" theory of diffraction,
which is adequate for normal diffraction experiments, including the measurement
of residual stresses by x-rays.
In the following sections the methods and equipment required for the precise
determinatio;n of d hkl , required for residual stress measurement by x-rays, will be
discussed. From the discussions given above, it can be seen that, in order to
determine lattice spacings, one needs a well defined monochromatic beam, a
means of setting the crystal at a given angle with respect to the incident beam, and
a means of measuring the diffraction angle 2e. There are various ways of achieving
these objectives and these are treated below.

4.12 Monochromators
The beam exiting from the traditional x-ray sources contain white and characteristic spectra, and synchrotron sources produce white radiation. In many
experiments, however, monochromatic radiation is required. This may be
achieved by passing the incident beam through a fIlter with the appropriate
absorbing edge, as discussed in Sect. 4.7. After fIltration KII intensities usually drop
to 1/500 of their initial value, while K", drops by 50 %.
Another way of obtaining a monochromatic beam is by scattering the incident
polychromatic beam from a set of lattice planes within a single crystal. Only one
wavelength will diffract. Such devices are called crystal monochromators. Since in
I

./

Dr/~3\!

"-

"

"-

"\
Fig. 4.13. Focusing monochromators in
a diffractometer in (a) back -reflection
and (b) transmission modes. In both
cases S is the x-ray source, PD is the
powder specimen and D is the detector

91

4.13 Collimators and Slits


Diffractometer plane

11

Fig. 4.14. Schematic description of diffraction angles in an experiment utilizing a


monochromated incident beam

the monochromator the x-ray beams are attenuated through absorption and
scattering, the exit beam is substantially weaker than the incident beam. In some
applications monochromators are bent to focus x-rays (Fig.4.13), as this
increases the incident x-ray intensity on the specimen. Such set-ups are termed
"focusing monochromators". The diffracted beam from a crystal monochromator contains not only the primary diffracted wavelength ')., but also its
harmonics at ').,/2;:').,/4, etc. Various strategies can be employed to get rid of the
harmonics, and to increase the intensity of the diffracted beam. The reader is
referred to the standard textbooks on the subject for a detailed treatment.
One other effec;t arising from the use of a monochromator to filter the incident
beam is the change in the polarization correction. If the diffraction angle from the
monochromator is 29m (Fig. 4.14) the polarization factor for the beam diffracted
from the sample at an angle 29 is given by
( P F) = 1 + K cos 2 29

m
1 +cos 2 29 '
where K is equal to cos 2 29m for a perfect monochromator crystal, and Icos 29m I for
an imperfect one.

4.13 Collimators and Slits


X-rays are radiated in all possible directions at all points of the target which is hit
by an electron. This causes the production of a highly divergent beam of x-rays
with a conventional sources. Collimators and slits are devices that are used to limit
and define a portion of this divergent beam.
Collimators are primarily used in x-ray diffraction to obtain a direct beam as
close to parallel as possible. In its simplest form a collimator consists of two
apertures of radius r, separated by a distance d. With a point source (Fig.4.15),
such a collimator will limit the maximum divergence to
R=

...

2r
d'

(4.15 )

The beam exiting from the target of an x-ray tube will be composed of rays from a
very large number of point sources. This beam will thus contain divergent,

92

4 Fundamental Concepts in X-ray Diffraction

------il

~~lL---

- - d - - - ------t

Fig. 4.15. Geometry of a pinhole collimator with a point source


Spmn.

S ~=d~====----~~-

!-

- -- d

~w

Fig.4.16. Geometry of a pinhole


coll imator with a source of finite
size

convergent and parallel rays. The effect of the collimator discussed above on the
beam from such a source is shown in Fig. 4.16. The maximum angle of divergence
in this case is
A=

,...

4r
d'

(4.16 )

The beam will also contain convergent rays, with a maximum convergence
2r
cx= d+w .

( 4.17)

Line source

Soller plates

Tncident: beom sl its

Tube

Fig. 4.17. a Soller slit geometry and its divergence limiting effect. b Use
of slits and soller slits to limit divergence of x-rays in two mutually
perpendicular planes

4.14 Diffraction Patterns from Single Crystals

93

The pinhole collimators discussed above utilize very small apertures, causing a
large loss of incident beam intensity. In certain applications soller slits are used to
form a parallel beam. These devices are simply a number of thin foils placed
parallel to each other. The exiting beam is in the form of a line and contains mainly
rays parallel to the metal foils (Fig. 4.17). Thus, the divergence perpendicular to
the plane of the plane of the foils is, in large part, eliminated. The divergence
parallel to the plane of the soller foils may be limited by the use of rectangular
apertures which have their narrow dimension in the plane of the foils (Fig. 4.17b).
These apertures are called slits.

4.14 Diffraction Patterns from Single Crystals


Consider the case where a beam of x-rays impinges on a thin single crystal which is
placed at a distance D away from a flat holder containing photograhic fIlm
sensitive to x-rays.J Fig. 4.18). If the beam of x-rays is monochromatic, diffraction
can occur if any set of crystallographic planes with a non-zero structure factor is
oriented at the correct angle, such that Bragg's law is satisfied. In this case there
will be two spots on the fIlm, one from the transmitted beam at the center of the
fIlm and the other due to the diffracted intensity from this particular set of planes.
The Bragg angle corresponding to this spot can be obtained from
tan9=

d'

(4.18 )

where r is the distance between the transmitted beam spot and the diffracted spot.
The plane spacing can then be determined from Bragg's law. If the crystal is now
rotated an angle <I> around the incident beam, while keeping the Bragg angle
constant, the diffracted spot on the fIlm will move on the circumference of a circle
of radius r, to a new position on the fIlm.
If, however, a polychromatic beam ofx-rays impinges on the crystal, there are
a number of wavelengths (with the shortest at AsWL ) available for diffraction. In
this case diffraction will occur for all the sets of planes which can satisfy Bragg's
law with one of the available wavelengths, resulting in a number of diffracted spots
on the fIlm. The Bragg angle for any of these spots is also given by Eq. (4.18). In
this case the wavelength of the diffracted beam is not known. However the pattern
Film

Fig. 4.18. Transmission Laue camera geometry

94

4 Fundamental Concepts in X-ray Diffraction


Film

d---l
Fig. 4.19. Back-reflection Laue camera geometry

of spots on the fllm is related to the point group symmetry, and can be used in
some point group symmetry determinations. These patterns"are also valuable in
determining the orientation of a given face of the crystal. X-ray patterns of this
type are called Laue patterns, after von Laue who obtained the first pattern of this
kind in the first diffraction experiment ever performed. The particular geometry in
Fig. 4.18 is a iransmission Laue pattern. Laue patterns can also be taken in the
back-reflection mode. The geometry for this case is shown in Fig. 4.19. In this case,
the Bragg angle associated with any spot is given by
.

tan ( 1800 - 28)

D '

( 4.19)

where the terms are the same as in Eq. ( 4.18 ) .

4.15 Diffraction Patterns from Polycrystalline Specimens


Consider the case when a coarse grained polycrystalline specimen, instead of a
single crystal, is placed in the transmission Laue camera shown in Fig. 4.18. Since
there are a number of grains (single crystals) in the specimen, with different
orientations to the incident beam, planes of different spacings may be available for
diffraction from each grain. Thus, a monochromatic beam will diffract at different
angles from each grain, causing a number of spots on the film. (In this case, as
opposed to the .polychromatic Laue pattern, each spot is from a different grain. )
The Bragg angle for each spot is given by Eq. (4.18). As the grain size gets smaller,
the number of the crystallites in the irradiated volume increases, and a larger
number of crystallites may diffract at a given 28 angle. These crystallites may be at
various orientations with respect to the incident beam. Thus the diffraction
pattern from this particular set of crystallites will look like a number of spots
arranged in a oircle of radius r given by equation (4.18). The total diffraction
proflle, from all the grains in the material diffracting from various (hkl) planes,
will be a series of these concentric spotty rings. For solids with very small grain sizes
(less than 100 microns), and random grain orientations, the spotty rings will
become continuous, since, for any (hkl) plane, there will be a statistically equal
number of diffracting grains for any rotation around the incident beam. In this

4.16 Basic Diffractometer Geometry


D

~----a

95

Fig.4.20a,b. Formation of the


diffracted cone from a small grained
random polycrystalline specimen
irradiated with monochromatic
radiation. Statistically there can be
an equivalent number of grains for
all rotations of the diffracting plane
about the incident beam

case the diffraction pattern emerging from the material for each (hkl) plane is in
the shape of a cone, with the specimen at the apex (Fig. 4.20). The intersection of
this intensity cone with the film plane causes a ring for a given family of planes,
{hkl}. (This case is equivalent to rotating a single crystal, diffracting at an angle
29 to a point P which is at a distance r from the transmitted beam, 3600 around the
axis defined by the incident beam while keeping the 29 constant. As noted before,
the diffraction spot will describe a continuous circle of radius r on the film.) These
patterns may also be obtained in the back-reflection geometry. In this case the
diffraction angle is related to the ring radius through Eq. (4.18). In both cases the
circles in such "pinhole" patterns are called Debye rings. From the radius of
Debye rings one can obtain the spacing of the diffraction planes through
Eqs. (4.18) or (4.19) and (4.13). The appearance of the rings also yields
information about the grain size in the material.

4.16 Basic Diffractometer Geometry


The methods, discussed above, for obtaining the diffraction pattern from a
sample, utilized photographic film sensitive to x-rays. The most common
apparatus used for examining diffraction patterns is the diffractometer, which
utilizes electronic counters sensitive to x-ray photons, mechanical rotation
devices to rotate the specimen, and a detector to facilitate real time measurements of x-ray intensities. In Fig.4.21 a commercial diffractometer that can be

Fig. 4.21. A Powder diffractometer commonly used in residual stress measurements

96

4 Fundamental Concepts in X-ray Diffraction

-diffractom ete r circle

Source

28
28 + 268

normollo specimen's surfoce

+
focuSing
Fig, 4.22. Schematic of an x-ray
diffractometer

circle

Fig. 4.23. Focusing in the diffractometer geometry.


During a 0 - 20 scan the ratio of the specimen
rotation to the detector rotation is 1:2

used for residual stress measurements, as well as powder work is shown. The
geometry of such a diffractometer is shown schematically in Fig. 4.22. X-rays from
a tube (or any other source) impinge on the specimen which is on a holder
that can rotate around an axis (X) perpendicular to the plane of the diffractometer. The diffracted beam is then detected by a suitable detector, which can be
rotated around X, along the circumference of the diffractometer. The rotation of
the specimen holder and detector are through stepping motors and gear drives and
can be determined very accurately (0.005 to 0.00052e). Slits are mounted
in front of the tube and the detector, respectively, to collimate the beam
and define the tp.aximum "horizontal" divergent angles (i.e., in the diffractometer
plane) permissible in the incident and diffracted beams. Soller slits may also be
used in the incident and diffracted beam paths for better "vertical" collimation. In
this case the foils in the soller slits are parallel to the diffractometer plane.
During a normal run with a flat polycrystalline sample the detector is moved along
the diffractometer circle to detect the diffracted beam from the specimen. At the
same time the specimen holder is also rotated around X half the rotation angle of
the detector in order to maintain focusing, as is shown schematically in Fig. 4.23.
In this way, the divergence in the incident beam is made to converge or focus at the
receiving slit, because both sets of rays section equal arcs on the focusing circle.
The detection surface of a diffractometer, i.e., the surface the receiving slit
describes as it is rotated around the diffractometer center, is a cylindrical segment
of radius R (the diffractometer radius) and height I (the length of the receiving
slit). The detector in a diffractometer sees only the intersection of the diffracted
rays with this ~urface. The width of the receiving slit determines the accuracy with
which the diffracted intensity at any 2e position along the circumference of this

97

4.17 Intensity of Diffracted Lines for Polycrystals

cylindrical segment is measured, while the area of the slit controls the total
intensity admitted into the detector.
Thus optimization of angular resolution with intensity usually results in a
rectangular slit (with its small dimension in the plane of the diffractometer circle)
which has a rather high aspect ratio (1/15-1/5).
The particular strategies for peak detection, slit arrangements and errors
caused by these will be discussed at length in Chap. 6.

4.17 Intensity of Diffracted Lines for Polycrystais


The most common kind of data one obtains from a diffractometer is a chart
(Fig. 4.24) or a table that shows the variation of diffracted intensity as a function
of 2a, from which the apparent peak positions can be easily determined. From
Fig. 4.24 it can be seen that the diffracted intensity is not equal for each peak. The
relative diffracted intensities of diffraction peaks for a given diffraction pattern
are dependent on six parameters:
a) structure factor
b) multiplicit'y,
c) polarization factor,

d) Lorentz factor,
e) absorption factor,
f) temperature factor.

Of these, the structure factor and the polarization factor have been previously
discussed, and are given by equations (4.11) and (4.14) respectively. It may be

.'j
--

...

~
-

Fig. 4.24. Typical output from a diffractometer

,--

98

4 Fundamental Concepts in X-ray Diffraction

recalled that the polarization factor is a trigonometric factor that is a function of


29, while the structure factor is a material property, dependent on the symmetry of
the particular crystal. These factors may cause changes in the apparent peak
positions determined from the raw intensity vs. 29 obtained from the diffractometer electronics, and such data must be corrected for these factors before the 29
position of the maximum intensity is determined. In the following sections the
remaining factors affecting the variation of intensity with 29 will be discussed.

4.18 Multiplicity
Multiplicity (p) is also an intrinsic property of the unit cell and the indices h,k,l of
the diffracting planes. This term describes the number of equivalent planes that
can diffract at a given Bragg angle, i.e., the members of a given form of planes
{hkl}. For example, for a cubic crystal all eight members of the {111} form have
the same d spacing (thus, Pill = 8) and any grains in which one of these planes
satisfy the Bragg conditions diffract at the same angle. By comparison there are
only six members of the {1 OO} form (p I 00 = 6 ). Thus, all other things being
equal, the diffracted intensity from the {1 oo} will be 3/4 of the diffracted intensity
from {111} planes.
Multiplicity is also a function of the unit cell of the crystal. For example, in a
tetragonal lattice (001) and (100) planes do not have the same spacing, and thus
can not diffract at the same Bragg angle, and do not really belong to the same
form (since they are not, in effect, equivalent planes). For this lattice PIOO =4,
and POOl =2.

4.19 Lorentz Factor


The Lorentz factor is a collection of trigonometric terms that describes the
dependence of the diffracted intensity (maximum or integrated) on the diffraction angle. There are three factors that contribute to the Lorentz factor:
1) The number of grains which are oriented such that they can diffract at a given
angle 29.
2) The diffracted intensity per unit length of the diffraction cone.
3) The dependence of diffracted intensity from anyone crystallite on 29.
The first factor is dependent on the volume fraction of grains oriented such
that they can diffract at or very near the particular Bragg angle. Even if a
completely random distri~ution is assumed, this volume fraction is not the same
for all orientations. Assume that a sphere of radius r is drawn around a
powder sample (Fig. 4.25). If the angular range around the Bragg angle into
which appreciable intensity is diffracted is 09, then the normals to the planes
that can diffract will intersect the sphere within a band of width r' 09,

4.19 Lorentz Factor

99

Fig. 4.25. The distribution of plane normals in space


for a particular cone of reflected rays

with an area ofr082m [sin (90 -8)]. The fraction of the crystallites is the ratio
of this area to the total area of the sphere, 4m 2 :
AN

N =

r08 2m sin ( 90 - 8 ) A8cos 8


41tr2
~
2

( 4.20)

Thus the fraction of diffracting particles is proportional to cos 8 and is small for
large 8 (the back-reflection region).
The second factor, intensity per unit length of the diffraction cone, is
important because, as discussed before, in normal x-ray work only a small
segment of the intersection of the diffraction cone with the detection plane is used
in measuring intensities. However, since the radius of each cone is different, the
diffracted intensity into each unit length of the cone circumference is different. If
the radius of the diffractometer is R, the radius of the circumference of any
diffraction cone is 2nR sin 28. Thus, the diffracted intensity per unit length is
proportional to l/sin 28.
The third factor describes the variation of intensity with 28, when all other
factors are constant. It was previously shown that, when deviations from ideality
occur in the wavelength distribution of the x-ray beam, or in the degree of perfection
of the crystal, appreciable x-ray intensity is observed at angles 08 away from the
Bragg angle 8B, resillting in the broadening of the x-ray peak. Maximum intensity
on the other hand, occurs at 8B Thus, diffracted intensity is a function of 8.

9,

. \.
a

A. a

- Na- - -

Fig. 4.26 a,b. Scattering from planes that are inclined at angles 28 - 88. a Specimen geometry;
b geometry of the diffracting planes

4 Fundamental Concepts in X-ray Diffraction

100

Consider a plane containing N atoms which is rotated an angle 09 from 9


( Fig. 4.26) . In this case the angles the incident and diffracted beams make with the
plane are 91 and 92. The path difference between the rays scattered by the two
atoms on the plane is
~=

a cos 92- a cos 9 1

(4.21 )

which can be expressed as


~=

2a09 sin 9B ,

4.22 )

where, since 09 is small, sin 09 ~ 09. The phase difference between the rays
diffracted from the first and the Nth atom on the plane is
N09=N2a09sin 9B

(4.23)

If this difference is equal to an integral multiple of the wavelength, the net


diffracted intensity is zero. Thus, the maximum rotation of any crystallite from the
Bragg position is

':
A.
(09ho=o= 2Nasm
. 9B .

(4.24 )

At 09 less than this limiting value there will be some diffracted intensity, with the
maximum intensity at 09 = O. Thus the diffracted intensity varies as a function of
1/sin 9.
All the trigonometric terms discussed above can be combined to form a single
factor which describes the variation of intensity with angle 9. This factor is called
the Lorentz factor, and is given by
1
sm \J

L.F=K.~,

( 4.25)

where K is a constant. The Lorentz factor and the polarization factor are usually
combined together to form the L - P (Lorentz-polarization) factor:
(L.P.) = 1 ~:2S:29

( 4.26)

The constants from both terms have been dropped in this case since they are not
dependent on 9, and will not affect relative intensities. It may be noted that this
equation is valid only for an unpolarized incident beam. If an incident beam from
a monochromator is used in a particular experiment, the polarization factor for
this case, given in Sect. 4.12, should be used.

4.20 Absorption Factor


Relative diffracted intensities may also be affected by absorption, which is directly
related to the path length the incident and diffracted beams traverse in the
specimen for a given geometry. Here the case for absorption in a flat plate on a

4.20 Absorption Factor

101

Fig. 4.27. The geometry of the specimen and the x-ray beams
for calculation of the absorption factor from a flat-plate
specimen in a diffractometer (a-goniometer)

diffractometer, (as given by Cullity [1 J , and Koistinen and Marburger [8J ), will
be treated, since this is the most common situation encountered in practice.
Consider the case where a beam of intensity 10 , of unit area in cross-section, is
incident on a flat plate at an angle (Fig. 4.27) . The total energy reaching a layer of
length 1and thickness dx, located at depth x below the surface, is proportional to al o
'e - IIAB because of absorption (here I.l. is the linear absorption coefficient of the
sample). The total energy diffracted by the layer is ablIoe - II AB dx, where a is the
volume fraction of crystallites that can diffract at this angle, and b is the fraction of
incident energy diffracted by unit volume. This diffracted intensity is also
attenuated by absorption along BC by a factor e- IIBC until it exits from the
material, thus the total diffracted intensity outside the specimen is given by
dID = albloe - II(AB + BC)dx .

(4.27)

From Fig. 4.27 it can be seen that


1
1=-.-;

smex

AB=-.-;
smex

thus,
dl D= (Ioab/sinex).ex

BC=~,

sm

P{-l.l. [-._l- + ~J}dX.


x

smex

(4.28 )

sm p

In an experiment the angle 'l' can be measured conveniently. Thus, making the
substitutions
ex = e+ 'l';

= e- \jJ ,

Eq. (4.28) becomes


loab 'ex - {x[
1
dl =
D sin ( e+ \jJ )
p
I.l.
sin ( e+ \jJ )

+ sin ( e1- \jJ )

J} 'dx

( 4.29)

4 Fundamental Concepts in X-ray Diffraction

102

The total diffracted intensity is obtained by integrating for an infinitely thick


specimen:
00

ID=

x=o

dID,

which can be shown to be equal to


ID =

;Jl (1 - tan 'V cot 9) .

I ab

(4.30)

Thus the angular dependence of absorption for this particular geometry is given
by the term (1 - tan 'V cot 9) which is called the absorption factor. We note that
for negative tilts (\jJ < 0) the factor becomes 1 - tan ( -ltVl) cot 9 = 1 + tanl'Vlcot 9.
Some examples of absorption calculations are given in Appendix E.
It can be seen from Eq. (4.30) that if the flat plate makes equal angles
with the incident and diffracted beams, i.e., 'V = 0, then the absorption factor is
equal to 1 for ~ll 9. This is the correct focusing configuration for diffractometers as
discussed before. Thus, there is no absorption correction for data obtained from a
29-9 run, where focusing is maintained. However, if 'V is non-zero, as during the
various tilts for a residual stress measurement, where the tilts are around the
diffractometer axis, a correction is necessary. The tilts required for residual
strain/stress determination can also be achieved by the geometry shown in
Fig. 4.28. This particular geometry is referred to as the 'V-goniometer. In this
geometry the tilts are around an axis parallel to the plane of the diffractometer and
the incident and diffracted beams have equal path lengths independent of the 'V-tilt
(i.e., in this case, the specimen behaves at all tilts as a flat plate in the correct
focusing condition since the 1:2 relationship between 9 and 29 is not affected by the
tilts). Thus, for this case, there is no absorption correction.
It may be noted that all the derivations so far have been based on flat
specimens. If the specimen surface has curvature, these formulae have to be re-

O-axis

~/-axis

(-

Incidence angle D

x- ray

source

Fig. 4.28. Psi-goniometer geometry for residual stress measurements

4.22 X-ray Detectors

103

evaluated in accordance with the specific geometry. Such curvature may also cause
other errors. These errors will be discussed in Chaps. 5 and 6.

4.21 Temperature Factor


In all the discussions so far, the crystal was treated as a collection of atoms at fixed
points in space. However, in reality atoms vibrate around their mean positions,
with an average displacement u that is proportional to the temperature. This
displacement is approximately 5 -10 % of the atomic spacing at room temperature and increases with increasing temperature. This factor causes certain effects:
1) The tabulated values of atomic scattering factor f do not include a vibration
term, and must be corrected for the temperature effect. In this case the
structure factor is written as f = fae - B sin 28fl.2, where B is a constant related to u,
which is dependent on the temperature of the experiment and the Debye
temperature of the particular atom. The corrected atomic scattering factors are
then used in the calculation of the structure factor [Eq. (4.14)]. The constants
required for the determination ofB are listed in the International Tables for xray Crystallography [7, vol. 3, p. 241]. (For further discussion also see
Schwartz and Cohen [2], p. 152f.) This decrease in intensity occurs because the
atomic planes can no longer be treated as planes but, in reality are slabs of
approximate thickness 2u. This affects constructive interference, since the path
(or phase) differences are changed and the total diffracted intensity
decreases. This decreases is especially noticeable at high 29 angles, since the
planes that diffract at these angles have small d spacings, and thus for a given
displacement u, the path differences are affected more.
2) There is a general increase of coherent radiation scattered in all directions, with
a corresponding increase in background intensity. This type of background is
called temperature diffuse scattering, and is more intense at high 29 (low d)
angles.

4.22 X-ray Detectors


X-ray detectors are used to determine the presence and the intensity of an x-ray
beam. Films or counters may be used for these purposes. X-ray beams cause
blackening in films. The degree of "blackness" is proportional to the x-ray
intensity, and can be measured with film densitometers. The position of the spot or
ring on the film can also be used to determine the Bragg angle of diffraction for the
particular spot, as discussed previously.
There are a number of counters that are commonly used in the detection of xrays. All ofthese counters translate the x-ray beam into voltage or current pulses
which are proportional to the x-ray beam intensity. The output from the detector
is then processed by electronics and displayed by an appropriate device. The usual

104

4 Fundamental Concepts in X-ray Diffraction

Fig. 4.29. Typical electronics rack for an x-ray detector. a, High voltage supply for detector; b, preamplifier; c, SCA; d, timer/scaler; e, ratemeter; f,
chart-recorder; g, micro-processor for motor control
and data acquisition

set of electronics (Fig. 4.29) consists of a matched set of amplifiers (preamplifieramplifier) which amplify the output from the detector, a single channel analyzer
( SCA) 1 which discriminates on the basis of energy to reject unwanted radiation
(thus decreasing background), a timer/scaler which keeps count of the number of
incident photons (counts) per unit time, and a ratemeter that is an analog display
of the number of counts per second. The output from a ratemeter is usually fed
into an x-y recorder to obtain the variation of counts per second, cps, y with time
(x) . By synchronizing the x axis with the rotation of the detector, one can obtain
the cps vs. 29 plot which was previously shown in Fig. 4.24. Alternately the digital
output from the SCA can be interfaced with a computer for on-line processing of
the data. The detectors commonly used in x-ray diffraction work are discussed
below.
,
A proportional detector consists of a gas filled tubular metallic cathode with a
conducting wire (anode) running along the long axis in the center of the cylinder
(Fig. 4.30). During operation a high potential (800 -1200 V) is applied between
the anode and the cathode. X-rays photons entering the tube ionize the gas, with
the electrons moving quickly to the central wire and the positive gas ions moving
slowly to the cathode. During this motion, these ions collide with more gas atoms,.
In normal x-ray work a filter and an SCA are used together to achieve the maximum rejection of
unwanted energies. The filter is most effective in minimizing intensities from rays with energies
close to that of the primary wavelength used in the experiment (such as preferential absorption
ofK~ with respect to K.), while the SCA is used to minimize the background from wavelengths
further away in the energy spectrum.

4.22 X-ray Detectors

105

level of detection

volts

time

Fig. 4.30. Schematic of a proportional


counter

Fig. 4.31. Schematic response of a proportional


counter to multiple pulses

causing further ionization. Thus, the amount of electrons (and ions) produced by
a single photon is amplified considerably (10 3 -10 5 ). This discharge produces a
current pulse, and momentary change in wire voltage, whose time duration can be
controlled by the composition of the gas in the counter. The quicker the pulse (i.e.
the discharge of the ions and electrons at the respective electrodes) , the smaller the
recovery time of the detector and the better the ability of the detector to detect high
count rates. In Fig. 4.31 the output of the detector in response to multiple pulses is
shown schematically.
Another feature of proportional counters is that the amplitude of the voltage
change in the wire is proportional to the energy of the incident photons. This
permits the use of pulse-height analyzers (PHA) which compare the amplitude of
the detector pulses to preset high and low voltage limits and record an event if, and
only if, the incoming pulse is in the interval. Thus only pulses within a given
wavelength band are recorded. This permits the rejection of unwanted radiation,
such as flourescence and white radiation, and also of the higher harmonics from a
crystal monochromator.
Depending on the type and resistivity of the wire used in a proportional
counter, it is possible to locate the position (along the wire) at which the photon
( or rather the ionization event caused by the photon) is detected. This is done by
measuring the relative time for the voltage pulse to reach the two ends of the
detector. A proportional counter used in this mode is called a position sensitive
detector (PSD). By curving the detector to fit the circle of the diffractometer, or
by using a short detector along a large diffractometer circle, it is possible to detect
the intensities over a range of two theta angles simultaneously, shortening
measurement times considerably. Such PSDs can achieve a spatial resolution
between 50-180 11m. Thus for a typical diffractometer radius of '" 15 cm, the
angular resolution is approximately 0.02 - 0.07 degrees two theta. (Another
interesting variation of this detector is the placement of PSD wires in a two
dimensional array. This set-up permits detection of diffraction spots referred to
two axes and, for e{(ample, may be used in real-time detection of Laue spots in a
camera.) Since the PSD substantially reduces the time necessary to record 20 vs.
diffracted intensity over a range of 20 (by a factor of ten in some cases), its
operation will be described in detail, following the summary given by James [9]:
In a PSD, the pulse on the anode wire can be used to determine the position of
the initial ionization (or the incidence position of the x-ray photon). There are

106

4 Fundamental Concepts in X-ray Diffraction

two methods that may be used to determine this position, utilizing the amplitude
or the rise time of the pulses occuring at two ends of the wire.
In the amplitude or current ratio method, a high resistance electrode is used. In
this mode the ends of the detector wire are effectively grounded and the charge
induced in the electrode leaks off to both ends. The sharing of the current i,
between the left (L) and right (R) flow directions is inversely proportional to the
resistive paths and, thus, directly related to the positions XR, XL:

here iL + iR = i and XR+ XL = Lo, which is the wire length. The position of incidence
is then determined by the use of a ratio circuit.
The rise time method utilizes the time difference between rise times exhibited at
each end ofth,.e anode in determining the position of the incident photon. The rise
time depends on the time constant RC, where R is the resistance along the anode
wire and C is the effective capacitance seen by the charge. For matched load
capacitance at each end of the wire, the time difference is directly related to the
resistance, and thus to the incidence position, if the distributed capacitance along
the wire is negligible. However, it has been shown that for such a design, a nonlinear region of position sensitivity exists within 20 % of the wire length from the
ends, which shortens the effective detector length by approximately 40 %. On the
other hand, rise time method allows the use of simpler (and cheaper) electronic
circuitry than that used for the current ratio method, with better angular
resolution. Thus, this method enjoys wider usage.
The basic circuit for a PSD utilizing the rise time technique for location of the
ionization event is shown in Fig. 4.32. The voltage sensitive preamplifiers are
located right next to the anode-cathode assembly (usually within the same box)
to reduce noise., The voltage induced at each preamplifier, Vo, is amplified further
in the main amplifiers. Shaping is accomplished by a double RC differentiating
and integrating circuit in the main amplifiers, producing a bipolar pulse. The rise
time of Vo can be referenced to the crossover point of the bipolar pulse by using a
crossover pickoff discriminator. When the amplitude of the input signal to the
discriminator exceeds a biased threshold level (a variable amplitude sensitive
register used to detect the leading edge of the signal) the pickoff is armed. When
the input reaches zero volts (the crossover point), an output pulse is generated.
The discriminator is reset when the negative amplitude from the bipolar pulse is
detected. The output signal at this point is a rectangular pulse of constant
amplitude, whi<;;h is used in timing.
Due to the different resistance of the wire segments, the crossover point from
the left and right sides will occur at different times as shown in Fig. 4.33. Thus each
event is characterized by two pulses which can then be used to operate a time to
amplitude converter (T AC). This circuit simply initiates the charging of a
capacitor at a linear rate when the start pulse from one end is received, and stops

4.22 X-ray Detectors

107

LEFT

RIG HT
LEFT

RIG HT

Fig.4.32. Block diagram of the electronics employing risetime difference detection principle

-----v.:::C2
Dl-~

SUM

T SeA

Fig. 4.33. Signal processing diagram for the electronic shown


in Fig. 4.32

Cl

T AC

charging when the stop pulse from the other end of the wire is sensed. Thus, the
charge in the capacitor is proportional to the time difference between the start and
stop pulses. If a second start pulse is issued before the stop pulse, it is ignored. This
effect contributes to the deadtime, i.e., the resolving time of the detector.
In order to make sure that the start and stop pulses are always in the correct
sequence, independent of where the ions hit the wire, delay circuitry is used. For
example, in Fig. 4.33, if the event takes place closer to the left end, a stop pulse is
issued before the start pulse. However, it is delayed a fixed time, adjustable at the
crossover pickoff analyzer, such that the start pulse reaches the TAC first.
The energy analysis in this case is performed in a timing single channel
analyzer (TSCA) which compares the sums of the amplitudes from both
preamplifiers (which have been summed concurrently by a summer circuit) , and
compares it to the user-set voltage limits. If the pulse is within this window, a
rectangular pulse is produced by the TSCA which can be synchronized with the
output of the T AC storage capacitor. If this pulse is absent, the output of the T AC
is grounded. Ifit is. present, the signal generated by the TAC, which indicates the
incidence of an x-ray photon of a given energy on the detector, and which contains
the necessary information about its incidence position, is sent to the storage device
which may be a multi-channel analyzer (MCA) 1, or a computer. A detector used
in this mode is terIJled to operate in the coincidence mode. A logical true pulse,
signifying correct energy from the TSCA must be present to validate the recorded
events, resulting in effective energy discrimination, which yields lower background intensities.
Another type of counter currently in use is the scintillation counter which
utilizes materials that produce visible light upon encountering an x-ray photon.
The visible light flashes are then detected by a photomultiplier tube and then
processed with various electronics. Generally these set-ups produce amplifications
In the MCA the pulses from the TAC are stored in memory locations (channels) according to
their amplitude. Every time a pulse of a given amplitude is received, it is processed by an analog
to digital converter and the memory location corresponding to that amplitude is incremented
by one.

108

4 Fundamental Concepts in X-ray Diffraction

of the order of 107 , resulting in voltage pulses of several volts. These pulses are also
proportional to the energy of the x-ray photons. However, the amplitude-energy
distribution is about twice as broad as for gas fllled proportional counters, so
energy resolution with a scintillation counter is poor. These detectors, coupled
with the appropriate slits, are widely used in step or continuous scanning through
peaks in diffractometers.
Solid state detectors (SSD) utilize a crystal of pure Ge or Li doped Ge or Si,
and are operated at liquid nitrogen temperature. The incident x-ray photons excite
electrons from the valence band or an impurity level into the conduction band,
producing electron-hole pairs. Under an applied voltage these pairs will produce a
current proportional to the number of incident photons and minute voltage
variations. These signals are amplified and anaiysed by electronics. SSD's are
operated at liquid nitrogen temperature to eliminate thermal noise. They have
very high energy resolution, typically better than AE=200 eV. With standard
electronics these detectors can exclude radiation outside a band of 0.02 keY,
which far exceeds the 20% resolution typical of gas proportional counters and
results in excellent background tp peak ratios for a given wavelength. However,
these detectors are cumbersome and expensive and thus enjoy limited use in
residual stress determination by x:-rays, although they are sometimes very useful
for this purpose.

4.23 Deadtime Correction for Detection Systems


In the course of a given diffraction experiment a detection system must be able to
process a broad range of intensities, ranging from very low (20 -1 00 cps) to high
( 8 -10 Kcps ). If the total (detector and electronics) is not linear over this range,
bias is introduced into the measurement, and the data, (cps vs. 29 for example) ,
must be corrected for such non-linearity before being processed for further use.
This correction has been described in Schwartz and Cohen [2]:
If in a given detection system No is the observed number of events, and NT is
the true number, these quantities are related to each other for small deviations
from linearity by the equation

(4.32 )
where 't is the time constant (dead time) of the system. Let the subscripts 1 and 2
refer to a foil in and out of the beam. Then

(4.33)
Defining the terms RT , Ro:

(4.34)

4.24 Total Diffracted Intensity at a Given Angle 28

109

0.49

0.48

0.47

Cu radialion
35 kV -23mA
T=

(6.08

:t

0.20) X 10- 6 sec

1,000 2,000 3,000 4,000 5,000 6POO 7POO 8,000

Fig.4.34. A typical plot for the


determination of dead-time

N 10 (cps)

substituting (4.34) into (4.33), and rearranging:


R o =RT +t(1-RT )(No )t

(4.35 )

As (No) t goes to zero, Ro approaches RT . This equation can be used


experimentally to determine the dead time t. One chooses a strong diffraction
peak from a single crystal, and then obtains the counts with and without a foil at
maximum intensity. This procedure is repeated at different 29 angles, i.e., over a
range of intensities. By plotting the ratio of the counts with and without the foil in
the beam path (Ro) vs. the number of counts without the foil (No)t, one can
obtain RT from the intercept of this plot and the deadtime t from the slope. A
typical result is shown in Fig. 4.34. Once t is obtained, measured intensity values
can be corrected for this effect using equation (4.32). It may be noted that, since
the correction is dependent on the number of counts recorded per unit time, and in
a Bragg peak intensity varies with 29, the correction is a function of 29.

4.24 Total Diffracted Intensity at a Given Angle 29


In previous sections, the factors contributing to the total diffracted intensity at a
point 29 were treated. These terms can be combined to obtain an expression
describing the total intensity at any 29:
I-K f2 e -2B" 29/')..2
1
1
. a
sm
. ABS(9,\jJ) DT(29)

(4.36 )

Here K is a constant that contains all the terms not dependent on 29, DT ( 29) is
the dead time correction and ABS (9,\jJ) is the appropriate absorption term

4 Fundamental Concepts in X-ray Diffraction

110

(Sect. 4.20) . The terms in Eq. (4.36) describe all the terms that contribute to the
angular variation of the observed intensity. These terms may cause distortion of
the x-ray peak, such that the apparent maximum is shifted from the position
dictated by Bragg's law. Thus, for high accuracy, the observed intensity at each
point on the x-ray peak must be corrected for these factors before the 29 position
of the maximum intensity is used in Bragg's law [Eq. (4.13)] for the determination of the plane spacing (d).

4.25 Depth of Penetration of X-rays


In previous sections absorption and its effect on the peak position of the diffracted
rays were discussed. The strong attenuation due to absorption processes also
limits the depth of penetration of x-rays, such that diffraction occurs only from a
very thin surface layer of the material. This depth, which is dependent on the
absorption coefficient of the material for a given beam, and the beam dimensions
on the specimen surface, which are defined by the slits and the geometry of the
experiment, aefine the effective irradiated volume in a given experiment.
. Another effect of absorption is the variation of diffracted intensity with depth.
Since attenuation of the incident beam is proportional to the thickness of the
material it passes through, the contribution to the diffracted beam from layers
deeper down in the material within the irradiated volume is less. Furthermore, the
diffracted beam has to traverse still more material before it can exit from the
surface, suffering even more attenuation. The intensity diffracted from an
infinitesimally thin layer located at a depth x below the surface was given by
Eq. (4.29):

It~a.b
) .exp{-llx[ . (~
) + . (~
) ]}.dX.
+tp
sm +tp
sm-tp

dID= .
sm

The total intensity diffracted by the slab of material between this layer and the
surface, expressed as a fraction of the total diffracted intensity is thus given by
x
G =
x

xLJ dID
dID = 1-exP {-llx[
1
+ _ 1_ ] }
sin ( 9 + tp ) sin ( 9 - tp )

( 4.37)

x=o

which, for a diffractometer at the focusing position (tp = 0) becomes


Gx={1_e-Zf.lx/Sin6} .

(4.38)

Equation (4.38) describes the fractional contribution of progressively deeper


layers to the total (observed) diffracted intensity. The effective depth of
penetration is usually defined as the thickness that contributes 99 % of the
diffracted intensity, and may be calculated from this equation. For example, the
effective depth of penetration of Cr radiation into steel for the 211 peak
(29 = 1540 ) is approximately 5.4 microns. Examples of G x are given in Appendix
E.

4.26 Fundamental Concepts in Neutron Diffraction

111

It may be observed from Eq. (4.37) that the effective depth of penetration is
also a function of the 'I'-tilt as well as 29, and decreases with increasing '1'. Thus, if
the "d" spacing of the diffracting planes varies with depth over the maximum
(effective) penetration depth of x-rays, different parts of the gradient will be
sampled during successive "'I'''-tilts. This effect causes bias in the information
obtaned by the x-ray beam and may cause errors in the measured residual stress
values if not taken into account. Various stragegies for minimizing such errors and
correcting for them will be given in Chaps. 5 and 6.

4.26 Fundamental Concepts in Neutron Diffraction


Neutrons are fundamental subatomic particles with the mass of the proton, and
carry no charge. These particles are usually produced at nuclear reactors during
the fission process, after which they are allowed to scatter from the light atoms in a
moderator of heavy water or graphite. During these scattering collisions the
neutrons lose energy, coming to thermal eqUilibrium with the moderator. The
kinetic energy of the majority of neutrons is given by
1 2 3
2"mv =2"kTm ,

(4.39 )

where T m is the moderator temperature. For T m;:;:: 600o K, the moderated neutron
energies are in the millielectronvolt (meV) range. The wavelength of a neutron
beam can thus be determined from the equation of wave-particle duality (Sec. 4.3 ) :
h2
A.m= ( 3mkTm) 1/2

'

(4.40)

As the mass of the neutron is app-roximately 10- 27 kg, the wavelength of the
moderated neutrons is around 1 A, which is appropriate for diffraction from
crystals. The neutron sources, like synchrotron sources, provide a white beam
without any sharp peaks of characteristic radiation ( Fig. 4.35a). This, as
discussed before, necessitates the use of a monochromator to obtain a
(monochromatic) beam suitable for use in structure or residual stress analysis.
Single crystal monqchromators are used for this purpose. Elimination of second
order wavelength (1../2) in such a beam can be achieved through the use of a
crystal whose structure factor is zero for the second order radiation, or through
filters.
A typical arrangement for neutron diffraction is shown in Fig.4.35b. The
hydrogen in the paraffin or masonite shield slows down any stray fast neutrons,
which are then abserbed by the boron. The lead shielding is to stop any x or yradiation from the reactor from reaching the diffraction area, thus decreasing the
stray radiation that can contribute to the background. Typical neutron fluxes at
the reactor core are about 10 14_10 16 neutrons!cm 2 !sec, and in order to attain
adequate intensity, the neutron beam at the reactor face has a rather large crosssection of 2 to 8 cm 2 This beam is collimated by the use of soller slits that are

4 Fundamental Concepts in X-ray Diffraction

112

wOler

--z.- COllima tor

Fig. 4.35. Distribution of intensity vs. wavelength in a steady-state neutron source. b Schematic
of a steady state neutron source and associated diffractometer

approximately 1 to 2 meters long, and have a plate spacing of 0.5 -1 cm apart.


This results in 'it horizontal divergence of the order of 0.25 to 0.5, and a vertical
divergence of about 1. A similar collimator is also used for the monochromatized
beam. The beam intensity at the specimen is of the order of 106 -10 8 counts per
second. However, neutron absorption in matter is generally much less than
absorption of x-rays. Consequently, large samples may be used to increase the
diffraction volume, and thus compensate for the low incident beam intensity. Even
then, the data acquisition time for neutron diffraction is longer than that for xrays.
In the set-up described above, only a small fraction (1 %) of the total
spectrum is selected by the monochromator, which then is utilized in a standard
diffractometer. If the neutron beam is pulsed, on the other hand (Fig. 4.36), all of
the available wavelengths may be utilized. In this apparatus, the beam from the
source is collimated and then fed into a chopper, which then emits a burst of

m---

NEUTRON SOURCE

COLLIMATOR #1
DISK CHOPPER
!!>----SAMPLE

28'1

~~COLLIMATOR #21
DETECTOR

Fig. 4.36. Schematic of the apparatus used in the production of


a pulsed neutron beam

113

4.26 Fundamental Concepts in Neutron Diffraction

."
40K

a:
I

~ 30K

(f)

420
331

311
II I

220

"

l-

200

>-

l-

v;

"

z 20K
UJ

I-

\..........
i\
CHANNEL NUMBER

Fig. 4.37. Typical data obtained from a pulsed neutron beam

neutrons at regularly spaced intervals (for example at 30 Hz). Alternately, one


may use a pulsed neutron source, in which a target (uranium for example) is
bombarded periodically with particles. These pulses cause a nuclear reaction on
impact, during which pulses of neutrons are emitted. The pulsed "packets" of
neutrons produced by either method then hit the sample and are scattered. At any
Bragg angle 29, a detector "sees" a pattern consisting of many Bragg peaks
( reinforced elastic scattering maxima from correctly oriented crystallites) , and a
white radiation background from incoherent scattering. These neutrons arrive at
the counter at different times t, because each plane chooses a different wavelength
or neutron velocity. Their time-of-flight ti over the fixed chopper-sample-detector
distance L depends inversely on their velocity Vi. From wave-particle duality we
can express the flight time in terms of wavelength Ai:
( 4.41)

Thus, if the incoming pulses are detected, processed electronically, and inputted
into a multi-channel analyzer which is re-started as each burst of neutrons leaves

4 Fundamental Concepts in X-ray Diffraction

114

Monitor

I
IHelium

chamber
atmosphere

Fig. 4.38. Schematic of the detector banks at the Intense Pulsed Neutron Source (IPNS) at
Argonne National Laboratories

the chopper, each channel in the MCA corresponds to a given wavelength. Typical
data. obtained in this manner, are shown in Fig. 4.37. This mode of analysis is
called time-of-flight (TOF) analysis, and has a higher resolution than neutron
diffractometers for d spacings larger than 1 A. For a detailed treatment of this
method the reader is referred to Schwartz and Cohen [2, pp. 250 - 252].
The time of data acquisition can be decreased even further if, instead of one
detector at an angle 29, as shown in Fig. 4.36, a bank of detectors is used, arranged
in a circle around the specimen. One such arrangement is shown in Fig. 4.38.

4.27 Scattering and Absorption of Neutrons


Neutrons, similar to the scattering of x-ray photons discussed earlier, also scatter
in elastic or inelastic modes. The elastic scattering component has the same
wavelength as the incoming neutron beam, and is also termed coherent scattering,
while the inelastic scattering causes the presence of an incoherent background.

4.27 Scattering and Absorption of Neutrons

115

Table 4.4. Scattering factors of selected elements for neutrons (b) and x-rays (f)
f-10 14m

b10 14m
sin a/A.
Hydrogen
Copper
Tungsten

0.1
-0.378
0.67
0.466

0.5
-0.378
0.67
0.466

0.1
0.23
7.65
19.4

0.5
0.02
3.85
12.0

Table 4.5. True absorption coefficients for neutrons and total absorption coefficients for x-rays
for selected elements
Element

Neutrons, m 2 jkg, 1.08

Be
Al
Cu

0.00003
0.0003
0.0021
0.0036
0.24

X-rays m 2 jkg, 1.54

0.150
4.86
5.29
1.72
0.24

Similar to x-rays, the coherently scattered neutrons interfere constructively or


destructively within the crystal, and form diffracted beams at the appropriate
angles predicted by Bragg's law. From the position of these maxima, the spacing of
the diffracting planes can be obtained.
On the other hand, as opposed to x-ray scattering, in which the fundamental
scattering body is the electron, neutrons are scattered mainly by the nuclei for
most materials (in magnetic materials, however, there is some scattering from the
electron cloud as well). Due to this effect, the amplitude of neutrons scattered by a
given atom does not depend on the atomic number, as it does for x-rays
( Sect. 4.8), but varies randomly within the periodic table. The scattering factors
of all atoms for neutrons are within a factor of two or three of each other (Table
4.4 ). The amplitude scattered also does not depend on the scattering angle, 29.
This (isotropic) distribution of scattering in space is due to the fact that the
dimensions of the nucleus, as opposed to that of the electron cloud, is very small
with respect to the wavelength of the incident radiation, and thus cannot cause
significant path (or phase) differences. Thus there is no polarization factor, or a
variation of neutron scattering factor with sin 9/A, as for x-rays.
In the absorption of neutrons in matter, the scattering contribution to absorption is higher than for x-rays. This is so, since, due to much lower scattering factors,
the penetration of the neutron beam, and hence the amount of matter it sees, (which
is proportional to the scattered intensity) is much larger than for x-rays. True
absorption coefficients for neutrons in some elements, as well as the data for xrays are shown in Table 4.5.
The large penetration depths of neutrons ( '" 20 mm in steel) , may be used to
advantage in residual stress determination. The measurement of "d" spacings
from deep within the material, which, when coupled with elasticity theory, yields
the average stress distribution at that particular depth, is an important extension
to the diffraction methods of residual stress determination and will be discussed

4 Fundamental Concepts in X-ray Diffraction

116

further in Chap. 5. For further information about neutron diffraction and neutron
sources the reader is referred to Bacon [6].

Problems
4.1. A Cu target x-ray tube is run at 40 kV and 20 rnA. The power input is given by
T/2

p=

V-i-sin 21t ( tjT) 2t


T/2
.

In standard tubes 1 % of this power is converted into x-rays.


a) If there is no dissipation of heat by water cooling, conduction, radiation, etc.,
estimate the time at which a 0.1 kg. Cu target starts to melt.
b) If the target is cooled by a water flow of 2 liters/min., determine the
temperature rise in the cooling water at steady state.
4.2. Monochromatic Cu K.. radiation is used to examine a polycrystalline Fe
specimen in a back-reflection Laue camera (D = 5.3 ems).
a) Determine the radii of the 200 and 400 diffraction circles.
b) Determine the same radii assuming Mo K.. radiation is used.
c) None of the radiations assumed above, would, ordinarily, be used in
diffraction from Fe. Why? What is a "good" radiation for Fe?
4.3. Calculate the infinite thickness of Fe for Cr,Co,Cu and Mo K.. radiations.
Infinite thickness can be taken to be the depth at which the incident beam is
attenuated to 1/1000th of its intensity at the surface. (Assume a 211 reflection).
4.4. The width of an x-ray peak from background to background is 2.6 29. Calculate the percent change in intensity due to the Lorentz-polarization- absorption
correction at the peak position and at the leading and trailing edges of the peak if
a) the peak is at 680 29 and b) at 1540 29.
0

Bibliography and References


1
2

B.D. Cullity, "Elements of X-Ray Diffraction", 2nd ed. Addison Wesley, Massachusetts, 1978
L.H. Schwartz and J.B. Cohen, "Diffraction from Materials", Academic Press, New York,
1977
3 L.V. Azaroff, "Elements of X-ray Crystallography", MacGraw Hill, New York, 1968
4 C.S. Barrett and T.B. Massalski, "Structure of Metals", 3rd. ed., MacGraw Hill, New York,
1966
5 H.P. Klug and L.E. Alexander, "X-ray Diffraction Procedures", Wiley, New York, 1967
6 G.E. Bacon, "Neutron Diffraction", 3rd. ed., Clarendon Press, Oxford, England, 1975
7 "International Tables for X-ray Crstallography I - IV", 3rd. ed., Buerger et al. eds, Kynoch
Press, Birmingham, England, 1976
8 D.P. Koistinen, R.E. Marburger, Trans. ASM, 51, 537 (1959)
9 M.R. James, "An Examination of Experimental Techniques in X-ray Residual Stress
Analysis", Ph.D. Thesis, Northwestern University, Evanston, II., 1977

5 Determination of Strain and Stress Fields by


Diffraction Methods

5.1 Introduction
Up to this point the mechanical and micromechanical behavior of solids and basic
concepts of x-ray and neutron scattering from crystalline solids have been
presented. In this chapter these concepts are combined in the derivation of the
basic equations of residual stress determinatipn with diffraction. The fundamental
assumptions inherent in these derivations and the limits they impose on the
applicability ofthe,.stress measurement will also be discussed. Various problems in
an actual stress measurement, such as the effect of stress gradients, the separation
of micro and macrostresses, determination of stresses in thin films and single
crystals, etc., are also considered, with special emphasis on the interpretation of
the data within the limitations of the theory.
In this chapter, however, it will be assumed that a perfectly aligned x-ray unit is
used in the measurement of exact data from a flat sample. These ideal conditions,
of course, are not always available to the experimenter and in chapter 6 the errors
caused by deviations from ideality will be treated in detail.

5.2 Fundamental Equations of X-ray Strain Determination


The orthogonal coordinate systems used in the following discussion are shown in
Fig. 5.1. The axes SI define the surface of the specimen, with Sl and S2 in this
surface. The laboratory system Li is defined such that L3 is in the direction of the
normal to the family of planes (hkl) whose spacing is measured by x-rays. L2 is in
the plane defined by Sl and S2 and makes an angle <I> with S2. In what follows,
primed tensor quantities refer to the laboratory system Li and unprimed tensor
quantities refer to the sample coordinate system Si> following the convention
established by Dolle[l]. Once the lattice spacing, d",,,,, is obtained from the
position of the diffraction peak for a given reflection hkl, the strain along L3 may
be obtained from the formula
(5.1 )
where do is the unstressed lattice spacing. This strain may be expressed in terms of
the strains t ij in the sample coordinate system by the tensor transformation
(Sect. 2.5)
( 5.2)

5 Determination of Strain and Stress Fields by Diffraction Methods

118

Fig. 5.1. Definition of the laboratory coordinate system Lj , simple coordinate system S;,
and the angles <l>,1J'

where a 3k ,a 31 are the direction cosines between L3 and Sk> Sl respectively. The
direction cosine matrix for this case is
aik =

cos 4> cos IP


- sin <I>
cos</> sinlP

sin <I> cos IP


cos 4>
sin 4> simp

-sinlP

(5.3 )

cOSIP

Substituting for a 3k> a 31 in (5.2):


')
doj>tp - do
( E33
oj>tp =
do

2,J,. . 2

. 2,J,. . 2

= Ell cos 'f'sm IP + E12 sm 'f'sm IP

+ E22 sin 24> sin 21P + E33COS 21P


+ E13COS 4> sin 21P + E23 sin <I> sin 21P,

(5.4 )

which is the fundamental equation of x-ray strain determination.


In polycrystalline materials, where it is possible to obtain a diffracted beam,
and thus a "d" spacing at alllP-tilts, three basic types of "doj>tp" vs. sin 21P behavior
are observed. These are shown in Fig.5.2a,b,c respectively. Fig.5.2a,b depict
"regular" d vs. sin 21P behavior which can be predicted by Eq. (5.4). When E13 ,E 23
are zero, Eq. (5.4) predicts a linear d vs. sin 21P behavior (Fig. (5.2a). When either

,
~

"tl
"tl

>/1>0

"tl

el/r<O

sin2",

sin2",

sin2",

Fig. 5.2. Types of "d" vs. sin21J' plots commonly encountered in residual stress analysis from
polycrystalline materials. The curves a,b exhibit regular behavior, the data of which follow
Eq. (5.4). The curve c exhibits oscillatory behavior which cannot be explained by this equation

5.2 Fundamental Equation of X-ray Strain Determination

119

or both of these components are non-zero, d measured at positive and negative 'I'
will be different due to the argument "sin 2'1''' associated with these terms, causing
a "split" in the d vs. sin2tp data (Fig. 5.2b). This effect is termed "tp-splitting"
[1,2,3]. Data exhibiting "regular" behavior can thus be analyzed by methods
based on Eq. (5.4). These methods are treated below. On the other hand, the
oscillatory d vs. sin2tp behavior shown in Fig. 5.2c cannot be predicted by
Eq. (5.4) without further modification. Analysis techniques for oscillatory data
will be treated in Sect. 5.13.

5.3 Analysis of Regular "d" vs. sin 2 1p Data


Equation (5.4) is a linear equation in six unknowns, E11,E12,E22,E33,E13,E23, and
may be solved exactly if d~1P is measured along six independent directions,
(L33) ~1P' In practice, however, more points are measured to improve accuracy.
For example, if the data exhibits "tp-splitting", the solution given by Dolle and
Hauk [2J, may be used. These procedures are described below.
Defining the parameters a 1,a 2:

(5.5a)
(5.5b)
where 'I' _ = ( -1 ) ''1' + and sin2tp + - sin2tp _ = 2 sinl2tpl,
Equation (5.5a) predicts a linear variation of a 1 with sin 2tp, with the slope and
intercept given by
(m~)al =E 11 coS 2<1> +E 12 sin2<1> +E22sin2<1>-E33 ,

( 5.6a)

and
(5.6b)
respectively. Similarly, a 2 varies linearly with sinl2tpl. The slope in this case is,
( 5.7)
Thus, if d~1P data is obtained over a range 'I' at three <I> tilts (0,45,90),and a 1
vs. sin 2tp and a 2 vs. sinl2tpl are plotted for all '1', the quantities Ell - E33 ,
1/2(E11 +2E12+E22'-2E33)' and E22 -E 33 , will be obtained from Eq. (5.6a) (at
<I> =0,45,90respectively). The intercept of a 1 vs. sin1tp is equal to E33 at all 'I' tilts
[Eq. (5.6b) J 1. Similarly, the slope of a 2 vs. sinl2tpl, [Eq. (5.7) for <I> = 00,900J
yields the quantities E13 ,E23 .
1 This condition may be used to check the compliance of experimental data with theory.

5 Determination of Strain and Stress Fields by Diffraction Methods

120

A simpler procedure, requiring less data points, may be used for linear d vs.
sin 2tp plots (Fig. 5.2a) that exhibit no tp-splitting. In this case, the strain tensor in
the Si coordinate system is of the form
Eij=

Ell E12 0
0 E22 0
o 0 E33

and d vs. sin2tp data is required at positive (or negative) 'I' tilts only. For such a
strain tensor, equation (5.4) becomes
. 2'"
. 2",
. 2tp+E
d",..,-do
do
= {Ell COS 2",
",+E 12sm
",+E22 sm
",-E'33}sm
33

( 5.8)

The right hand side of Eq. (5.8) is equivalent to that of Eq. (5.6a) and a similar
analysis may be used to determine Ell,E12,E22,E33.

5.4 Determination of Stresses from Diffraction Data


Once the strains are obtained, the stresses in the Si coordinate system may be
calculated from the general form of Hooke's law (Sect. 2.9):
(5.9)
where the elastic stiffness coefficients, C ijkl , are referred to the Si coordinate
system. The stresses in any other coordinate system may be determined from the
transformation rule for second rank tensors:

(5.10)
where ami are the appropriate direction cosines.
The procedlJre discussed above is the basis of all types of diffraction techniques
for stress determination. These techniques, however, usually express Eq. (5.4) in
terms of stresses, and simplify or modify the resultant equation according to the
properties of the material under investigation and according to the stress state
expected in the ~rradiated volume. In the following, these procedures are examined
in detail.
In the most general case, where an anisotropic material contains a general
triaxial stress tensor
O"ij=

0"11 0"12 0"13


0 P"22 0"23
o 0 0"33

( 5.11)

the strains in the sample coordinate system can be expressed in terms of stresses by
the inverse of Eq. (5.9):
(5.12 )

S.4 Determination of Stresses from Diffraction Data

121

1_-

_\

~
I

C2

/1

Fig. 5.3. Definition of the crystal axes C i


and their orientation with respect to the
laboratory axis L3 and surface axes Si

where Sijkl are the elastic compliances. The equation linking stresses to measured
diffraction data may be obtained by substituting Eq. (5.12) into Eq. (5.4) for all
Eij.1t must be remembered, however, that the elastic compliances are also referred
to the Si system of axes, and must be obtained from the elastic constants referred to
the unit cell axes (Fig. 5.3 ), by the tensor transformation rule for fourth rank
tensors [Eq. (2.50)]. For example, for Ell:
(5.13 )
Here akO are the direction cosines between the crystal axes and the surface
coordinate system and Smnop is defined in the crystal axes. Equation (5.13) and
similar equations, written for the other components of the strain tensor, can be
substituted into Eq. (5.4) to obtain the general equation linking measured d
values to the stresses existing in the sample coordinate system, Si.
If the material under investigation is isotropic, Eq. (5.12) becomes (Sect. 2.7)

1+v

Eij

=E

(Jij -

oij

(Jkk ,

(5.14 )

which upon substitution into Eq. (5.4) yields

( 5.15 )

5 Determination of Strain and Stress Fields by Diffraction Methods

122

It can be seen from Eq. (5.15) that, for isotropic materials, if the stress tensor
existing in the specimen coordinate axes, S;, is one of the following forms:

(T n (T
(T cr~J (T
0
0"22
0

g)

(5.16a,b)

0"12
0"22
0
b

0
0"22
0
c

0"12
0"22
0
d

cr~,)

(5.16c,d)

the d vs. sin2'1' plot obtained from the surface layers will be linear (Fig. 5.2a) . A '1'spitting in the d vs. sin 2'1' plot (Fig. 5.2c) on the other hand, indicates the presence
of the shear stresses 0"13,0"23.

5.5 Biaxial Stress Analysis


a) The "sin 2 tp" Technique

If the stress tensor existing in the irradiated layers is biaxial, (5.16a,b), or


Eq. (5.15) becomes
d~1p-do

do

where

O"~,

1 +v
. 2
V
= EO"~sm '1'- E (0"11 +0"22) ,

the stress component along the


2,j,.

2,j,.

"'+O"22sm '"
2
O"~ = 0" 11 cos <j> + 0" 12sin2<j> + 0"22sin2<j> ,
O"~=O"l1COS

S~

(5.17 )

direction (Fig. 5.1 ), is given by


(5.18a)
(5.18b)

for the tensors (5.16a,b) respectively.


Equation (5.17) is a form of the traditional x-ray residual stress equation, and
has been in use for over 60 years [3]. It predicts,as noted before, a linear variation
of d vs. sin 2 '1'. The stress in the S~ direction may be obtained directly from the slope
of a least-squares line fitted to experimental data, measured at various '1', if the
elastic constants E, v and the unstressed plane spacing, do, are known. This
procedure is known as the "sin 2'1''' technique [4], since it utilizes multiple 'I'-tilts.
Even though various methods are available for its evaluation (Sect. 5.7), do
may not be readily available in practice and the lattice spacing measured at 'I' = 0 is
substituted for do in the procedure described above. Such substitution is based on
the fact that, for most materials, elastic strains may introduce at most, 0.1 %
difference between the true do and d at any '1'. Since do is a multiplier to the slope,
the total error introduced by this assumption in the final stress value is less than 0.1
%, which is negligible compared to the error introduced by other sources.
Similarly, for textured materials, the x-ray elastic constants SI ,S2/2 are used instead

123

5.5 Biaxial Stress Analysis

ofE and v. Sl,S2/2 may be obtained from the literature for a given material and
reflection combination. If experimental values are not available, they may be
calculated from single crystal elastic constants using the various approximations
discussed in Sects. 3.1 0 - 3.12. They may also be measured experimentally, and the
procedures used for such measurements will be discussed in Sect. 5.14.

b) The Two-Tilt Method


This method (often used for rapid testing) assumes, a priori, that the
variation of measured d with sin 2tp is linear. Thus, only two tp tilts, tp = O,tp are
used in the measurement to defme d vs. sin2tp. Other modifications, designed for
ease of calculation, are also introduced. Eq. (5.17) is rewritten as
Ad 1 +v
. 2
V
( f = EO'.sm tp- E (0'11 +0'22) .

(5.19a)

The term Ad/d may be written in terms of the shift in 28 of the diffracted intensity
maximum between these two tp-tilts. Differentiating Bragg's law, nA,=2d sin 8:
Ad
cot8A28
(f=2

(5.19b)

Thus, from Eqs. (5.19a,b),


(5.19c)

0'.=KA28,

where A28 is the peak-shift. K is called the stress constant, and is given by
K=

E'cot8
2( 1 +v) sin2tp .

(520)
.

This procedure is accurate for small A28, but substitution of Eq. (5.19b) into
Eq. (5.19a) may not be valid for large A28. In such cases, one may use Eq. (5.19a)
with two tilts without this approximation.

c) Single Exposure Method

This method, as the name implies, acquires all the data required for a biaxial stress
analysis at a single tp'tilt, where tp O. The Debye ring from a stressed material at a
non-zero tp-tilt is non-circular as shown in Fig. 5.4a. This is so since the normals
N 1 ,N2 of the planes that diffract to points 1,2 on the diffraction cone are at
different inclinations (n 1 , n2) to the surface normal. Thus, the resolved strain on
these planes are different, and the plane spacing varies along the Debye ring
causing the deviation from circularity. Determination of the difference, A8,
between the Bragg' angles corresponding to the points PloP 2 from a film
( Fig. 5.3a), or from PSDs placed at the appropriate locations around the
incident beam (Fig. 5.4b), enables the calculation of the stress along S. from
}
E
{
cot [1/2(8 1 +82)]
0'.= (l+v)' sin 2 (p+n)-sin 2 (p-n) ,

.(5.21a)

124

5 Determination of Strain and Stress Fields by Diffraction Methods


Debye ring from on
unstressed standard

--~ "

ring from stressed


0(ecimen

(~'I-------l-------- \I
\

1
1

:
~
:
I

;I1

"'-

P2

1
1

:
--.!.-

..,/

/i

Bi
:
-----t--...,...--+--T"'""'--+..:..--

film

~'"\
\

a
~s

:------~

X-Roy S

Position Sensitive 1

Detecto~~
1
I
I

Specimen
SUrface"

--~-7~------------------

~\\~I
b

Fig.5.4. a Geometry of the single-exposure technique for residual stress measurement with a
film at specimen tilt ~ and effective ~ -tilts n 1, n 2 b The same technique with position sensitive
detectors

5.6 Triaxial Stress Analysis

125

where n1=90-9 1, n 2 =90-9 2 and it is assumed that n 1~ n2 ~ n. This equation


is usually expressed in the form
(5.21b)

crlj>= KL\9 ,
where K, the stress constant, is given by
K= ~.~.
1
9'" 9 1+9 2
2
180 1 + v 2 sin 2~ sin 29 '

It may be noted that the procedure is similar to the two-tilt method in that plane
spacings at two effective ",-angles, n 1 and n 2, are used to determine the stress. If
(9) Pl' (9) P2 are determined at multiple inclinations o,'" 1,... ,"'" one can use the
data thus acquired in the sin 2", analysis discussed above.

5.6 Triaxial Stress Analysis


The assumption in the traditional methods discussed above, that stress components in the direction of the surface normal (S3) are negligible in the volume
sampled by the x-ray beam, is not necessarily true; ",-splitting, which indicates the
presence of strains &13,&23' has been observed in untextured polycrystalline
materials [2,5,6J. Since such materials may be treated as isotropic solids [lJ, the
presence of these strains indicates the presence of the shear stresses cr 13 ,cr23
[Eq. (5.14)].
Analysis of "",-split" d vs. sin 2", plots in terms of stresses is similar to the
analysis for strains. The terms a 1,a2 are expressed in terms of stresses:

a 1 -_ {dlj>'P+ +dlj>'P_
2do

1}

1 + v {crdcos 2,l,.
2,l,.
2,l,.
2
E
..,+cr 12sm
..,+cr22 sm
..,-cr 33 } sm
'"
( 5.22a)
(5.22b)
Thus, the stresses cr11,cr12,cr22,cr33 may be obtained from the slope and intercept
of a 1 vs. sin 2", for <1>=0,45,90. The slope of a 2 vs. sinl2",1 for <1>=0,90,
respectively, yields the shear stresses cr 13'cr23. After the stress tensor in the sample
system Si is determined, the stress tensor in any other coordinate system, such as
the principal stresses and their directions, may be determined using the procedures
discussed in Sect. 2.6.
It has been noted in Sect. 5.4 that, if a triaxial stress tensor that does not
contain cr 13'cr23 exists in the irradiated volume, then d vs. sin 2", will be linear. In

126

5 Determination of Strain and Stress Fields by Diffraction Methods

this case, Eq. (5.15) becomes


')
dl>",-do 1 +v {
2
.
. 2
} . 2
( c33
1>",=
do
=
0" 11 cos <\>+0"12sm2<\>+0"22sm <\>-0"33 sm \jJ
1 +v
v
+0"33-E (0" 11 +0"22 +0"33).

(5.23 )

It may be seen from a comparison of Eqs. (5.17) and (5.23) that, if 0"33 is finite
within the penetration volume, the stress 0"1> determined by bi-axial methods from
the slope of the d vs. sin 2\jJ line will contain an error equal in magnitude to 0"33. In
such a case, the following method way be used [7]. If d vs. sin 2\jJ data is obtained
for two <\> tilts, <\>=<\>A' <\>=<\>A+90, where <\>A is the angle between Si and the
principal axis Pi (which is generally not known) , one can obtain from the slopes
of the (C~3) 1>", vs. sin 2\jJ plots,
l+v{
2,1,.
.
2
}
ml>A =
O"llCOS 'f'A + 0"12 sm2 <\>A + 0"22 sm <\>A - 0"33

1 +'v

(O"I>A -0"33)

(5.24a)

l+v{
2
.
ml>A+90 =
O"llCOS (<\>A +90) + 0" 12sm2 ( <\>A +90)
+ 0"22 sin2 (<\>A + 90) - 0"33}
l+v
=
(O"I>A+90-0"33)

(5.24b)

The sum of the slopes (5.24a,b) is equal to


l+v
ml>A + ml>A+ 90 =
{O"ll +0"22- 2 0"33}

(5.25 )

The intercept, I, of (C~3) 1>", vs. sin 2\jJ does not depend on <\>:
1=

C~V)0"33-; (0"11 +0"22 +0"33).

(5.26)

Equations (5.25), (5.26) may be solved together for 0"33 after which the stresses
O"I>A'O"I>A + 90 along the surface directions ~I>A'~I>A +90, may be determined from
Eqs. (5.24).

5.7 Determination of the Unstressed Lattice Spacing


In some of the procedures discussed above, it was assumed that do, the unstressed
plane spacing, was known. As noted above, for methods based on the biaxial
assumption, one may substitute the plane spacing at \jJ = 0 for do without
introducing a large error into the analysis. The triaxial methods, on the other

5.7 Determintation of the Unstressed Lattice Spacing

127

hand, utilize the difference, ~d = dcl>'P - do, in the calculation [Eqs. (5.5), ( 5.22) J.
The term ~dis very small for elastic strains. Consequently, even a 0.1 % error in do
can cause a large error in ~dl, and thus in the subsequent analysis. It may be seen
from these considerations, that in order to obtain stress results with acceptable
error from a triaxial solution, do must be within less than 0.01 % of its true value.
Determination of do to such accuracy may not be trivial and in the following,
various problems that may be associated with this measurement are discussed.
One method uses ftlings of the material under investigation. The powder
particles of such a sample cannot sustain a macrostress (the powder particles are
not rigidly bonded together and can not sustain macrostresses across their
boundaries) and, with small randomly oriented particles, the average microstress
in a representative volume is zero. Thus, the plane spacing determined from such a
sample will be that of an "unstressed" sample. In another variation of this method,
the ftlings are annealed to relieve all residual stresses and the plane spacing do is
determined from the stress-relieved powder. Alternately, one may anneal a solid
sample, such that all the residual stresses are relieved, and determine do from the
stress-relieved sample. These procedures, however, may also introduce other
factors that can affect the plane spacing. For FCC materials, a general formula
relating the change in 29 between an annealed sample and a cold worked sample,
to various parameters was given by Wagner2 [8J:
~29hkI = 29deformed - 29annealed
.

.~a

( 5.27)

=JtCx tan u+hCXBttan u+hO",tan u+J4-tan u.


a
Here jj are constants that are given in Table 5.1 for various reflections,

a
~a

a
9

=probability of an intrinsic stacking fault on (111) planes,


=the first stress invariant (0",=0"11 +0"22+0"33, [Eq. (2.29)].
=the fractional change in the (111) spacing associated with a stacking
fault,
=the lattice parameter,
= the change in "a" due to segregation to stacking faults,
= Bragg angle.

Thus, if filing or annealing for stress relief causes changes in ex, Ct, ~a , the do
a
values obtained from such procedures may be erroneous for stress measurement
purposes (Table 5.2). Such errors can be especially large in materials susceptible
to stacking faults. In such cases, the do measured from the annealed sample may be
corrected if, without annealing, one obtains the terms ex, Cl'

~a ,0", by measuring
a

1 For example, consider data from an iron sample. In this case, do = 2.8665 A. Assume that
d+",=2.8675 A. In this case, L\d=O.oo1. If do changes by 0.1 % due to an error, L\d becomes
0.00187, which is almost a 90 % change.
2 For ordered FCC structures, there are other terms that contribute to L\29hkl These terms are
discussed in [31]

128

5 Determination of Strain and Stress Fields by Diffraction Methods

Table 5.1. The constants j\,j2j3,j4 in Eq. (5.28)


Peak

j\

j2

Peak

j\

j2

100
110
111
200
210
211
220

-5.265
2.632
3.949
-7.897
-0.526
0
3.949

- 6.37
- 6.37
-23.87
19.10
10.82
- 6.37
19.10

310
311
222
320
321
400
331

-0.526
-1.436
-1.974
0.810
-0.188
3.949
0.831

- 6.37
- 4.34
-23.87
- 6.37
5.91
19.10
-38.20

The term j3 is the bulk modulus [Eq. (3.50)] for the particular material, and j4 is - 360/1t for all
materials.
Table 5.2. Variation oflattice parameter, ao, with cold work and various heat treatments in two
FCC alpha-brass alloys'
Material
Cu-30 Zn

Condition

Treatment

L~ttice

Filings

10 Mins. 300C
+ Water Quench
Room Temperature
1032 Hrs. 120C

3.6825
3.6832
3.6803

Filings
Filings
Cu-35 Zn

Filings
Filings
Filings
Filings

10 Mins. 300C
+ Water Quench
10 Mins. 300C +
1072 Hrs. 120C
Cold Worked Room Temp.
Cold Worked Room Temp.
+ 1031.25 Hrs. 120C

(A)

parameter,

3.6946
3.6940
3.6955
3.6918

Data excerpted from Otte [9]

L\29 of at least four peaks, and solving Eq. (5.27) for these terms through linear

regression. Then, the 29 of the material that is equivalent to the cold worked
specimen in stacking fault density and related terms, but has no stresses, is given
by
(5.28 )
do can then be obtained through Bragg's law.
For Bee structures Eq. (5.27) becomes [10]

A2e'

e'
L\a
e.
hkl=jzcxE 1tan 9
+hO"ltan
+J4-tan

il

(5.29 )

Thus, similar considerations may apply to Bee materials also, but the effects are
smaller since CXE 1 is a small quantity; the term in cx present for fcc system is absent
for bcc (and hcp) systems.

5.7 Determination of the Unstressed Lattice Spacing

129

Another method of do determination, given by Hauk et al. [11], utilizes data


acquired during the stress measurement itself. If the stress state in the irradiated
volume is biaxial, this method may be used to determine an exact do value from the
d vs. sin 2\jJ plots obtained from the actual specimen during the stress measurement.
For <\>=0, Eq. (5.17) becomes
. 2
V
d4>",-d o 1 +v
do
= E O " 11 sm \jJ - E (0"11 + 0"22) .

(5.30)

If for \jJ=\jJ', ~.=do, Eq. (5.30) may be written as


. 2
viE
'{1 0"22}
viE
(1 m2)
sm \jJ = (1 +v)/E)
+ 0"11 = (1 +v)/E + m 1

'

(5.31 )

where m 1 ,m 2 are the slopes ofd vs. sin 2\jJ at <\>=0,90 respectively and may be
determined experimentally. Thus sin2\jJ' may be obtained from Eq. (5.31) after
which do is obtained from the d vs. sin 2\jJ at <\> = 0.
A similar procidure for a triaxial stress state has also been suggested [11]. For
Eqs. (5.22a) or (5.23), sin2\jJ' is given by

(5.32)

This equation may be simplified considerably if it is assumed that v = 0.5 in the


term (1 + v) IE in the numerator. Thus, (5.32) becomes

(i) (1+

C~v)

m2)

m
1

(5.33 )
'

where the terms m 1, m 2 may be obtained from the slopes of d vs. sin 2\jJ for <\> =0,
90. The assumption that v =.5 in one term in the numerator, though, causes some
error in do determined from equation Eq. (5.33) [12]. This error may be obtained
by substituting Eq. ( 5.33) into Eq. ( 5.23 ):
( do ) ",' - do ~do
( 1 - 2v )
do
= do =
E
0"33 .

(5.34)

It may be seen that the error in do determined from Eq. (5.33) is a linear function of
0"33' Due to this error in do, the stresses, O"ii' determined from the triaxial methods

130

5 Determination of Strain and Stress Fields by Diffraction Methods

described above contain errors comparable to the value of 0"33 existing in the
irradiated volume. Thus, this method of do determination should not be used in
cases where a large 0"33 component may be present.

5.S Effect of Homogeneity of the Strain Distribution and


Specimen Anisotropy
It may be noted that the derivation of the fundamental equation linking measured

values to the strain tensor in the surface coordinate system, Eq. (5.4), contains no
assumptions about the symmetry of the unit cell of the material under
investigation. The procedure is based only upon the transformation law for second
rank tensors [Eq. (2.24)], and is applicable to all unit cells with arbitrary
anisotropy as well as to isotropic materials. The effect of anisotropy is taken into
account only~ when one tries to compute stress values from the strain tensor
[Eqs. (5.9), (5.13 ) ,( 5.15), etc.]. Thus, as long as the appropriate form of
Hooke's law is used, the methods are applicable to any type of unit cell. There are,
however, other implicit assumptions in the derivation and these are treated below.
In the derivation of Eq. (5.3) the position of the specimen axes, Si' is not
specified. The origin of Si can be at any point within the irradiated volume from
which d"[Eq. (5.1)] is obtained, and thus, the strain tensor, defined in Si is
assumed to be the same at all points, i.e., homogeneous in the irradiated volume.
Residual stress distributions, on the other hand, must be, by definition,
inhomogeneous. The stresses in one part of the body are balanced, according to
equations of equilibrium, with stresses in another part of the body, such that the
integration of stress over the entire volume is zero (Sects. 3.3 - 3.5 ) .
(3.11a)

Thus, in a given specimen, the dimension of the irradiated volume along a given
direction must be much smaller than the characteristic length along this direction
over which the residual stresses vary appreciably. Strain (and therefore stress)
values obtained from irradiated volumes with larger dimensions will be the
average of the actual stress distribution within the volume. The particular
averaging function in such cases is dependent on the distribution of the crystalline
regions within the specimen since only those regions that are diffracting can
contribute to the information obtained by the x-ray beam. Furthermore, the
diffracted intensity contribution from layers deeper in the material must be
weighted differently because of absorption (Sect. 4.20) . Finally, the desired strain
or stress components are obtained by least-squares methods from the average data
measured by x-rays. This, in effect, re-averages averaged data. Thus, the final
results are complicated averages of the local values, and unless the actual strain
distribution within the irradiated volume is known, interpretation of the average
data in terms of local deformation may not be possible.

5.9 Average Strain Data from Single Crystal Specimens

131

5.9 Average Strain Data from Single Crystal Specimens


Determination of the stress state in single crystals is of considerable theoretical
and industrial importance [13 - 21]. For example, in some applications, such as
in certain semiconducting devices, stress measurements from single crystal
specimens may be required. Similarly, in the case of materials with large grain size,
the stress state within an individual grain might be of interest. In such cases, the
methods discussed above may be used for strain/stress determination with the
appropriate modifications that allow the symmetry of the single crystal to be taken
into account (Sect. 2.9). These modifications are discussed below.
In a single crystal, the angles </>,\jJ that define the laboratory axis L3 are
uniquely defined by the particular crystal under analysis, since,in order to obtain
diffraction from a set of planes (hkl), the normal to these planes must be
coincident with L3 (Fig. 5.5) and because of symmetry, there is only one such
lattice direction. Once this normal is brought into the diffraction plane (which is
defined by the incident and diffracting beams) and rotated such that it makes the
appropriate Bragg':angle with the incident beam, the peak maximum (and thus the
strain along this direction) is determined by rotating (rocking) the crystal
through the Bragg angle, 29, and recording the variation of diffracted intensity
with the rocking angle. Once the lattice parameter and, thus, the strain along this
normal is determined, then the normal of another set of lattice planes (h'k'l') is
brought into the diffraction condition by the appropriate rotations 9,\jJ,</> and the
rocking curve around the appropriate angle 9h 'k'l' is recorded. This procedure
must be repeated until the strains (E~3).p,p are obtained along a minimum of six
independent directions, required for the solution of Eq. (5.4). Of course, more
data should be used (if possible ) so that a least-squares solution can be employed
to minimize random errors [13,14,15].

,/~
,/

,/

,/
,/

d hkl

'/'

,, ' '

~,

,/

',

, ,
,,, "

CI

-...~

I diffracted

Fig. 5.5. Relationship of the crystal axes C; with the incident and diffracted beams and
the laboratory axis L3 = ~hkl when the Bragg condition of diffraction is satisfied for a
set of planes, (hkl)

132

5 Determination of Strain and Stress Fields by Diffraction Methods


SAMPLE

position
sensitive
detector

Fig. 5.6. Parallel beam diffractometry from a single crystal using a position sensitive detector to
resolve the diffracted intensities from points on the specimen surface

For relatively perfect crystals, all points within the irradiated volume
contribute to diffraction because of the long range symmetry through the entire
volume of the crystal.
The strain at any given tilt is acquired from a homogeneous, continuous,
anisotropic volume and represents the average of the strains in this volume. The
calculation of the average stresses in the diffraction volume from the measured
average strains utilizes the appropriate elastic constants referred to the specimen
coordinate system and the general expression for Hooke's law.
If the single crystal under analysis has large subgrains with relatively large
misorientations, or the strain profIle is inhomogeneous within the irradiated
volume, it may be necessary to determine the strain in regions within the crystal
where the strain is relatively homogeneous. This may be achieved by a procedure
described by Mayo et al. [15], where a parallel x-ray beam impinges on a single
crystal and the diffracted beams are detected by a position sensitive detector
whose wire is parallel to the a-rotation axis of the crystal (Fig. 5.6) . The diffracted
intensity registered by the PSD at any position will correspond to a definite
position on the specimen. Thus, within the linear resolution of the PSD
( Sect. 4.22), the diffracted intensity from different positions along the intersection of the beam with the specimen may be stored in a channel of a multichannel
analyzer. If the specimen is rocked through the Bragg angle, the variation with
rocking angle of intensity in each channel is obtained, which is, in effect, the Bragg
peak profIle for each point on the specimen surface. The intensity maxima can
then be used to determine the strain corresponding to each position on the
specimen (Fig. 5.7). Rotating the specimen to bring other reflections into the
diffraction condition and repeating the procedure, provides the data required for
the strain analysis at each point along the intersection of the beam with the
specimen. Stresses may then be obtained through Hooke's law. It must be noted,
however, that if there is misorientation between the subgrains, the direction
cosines between the specimen axes and the crystal axes (and thus the elastic
constants referred to the specimen axes) will be different for each subgrain.
Furthermore, if the deformation distribution is inhomogeneous, reflections that

5.9 Average Strain Data from Single Crystal Specimens

133

Uniform
Strain
a
~=O

--

"g

.J SI

QI>.

B
C

28

Nonuniform
Strain

Fig. 5.7. Variation of the intensity with rocking angle at the points along
the surface when the strain distribution along A,B,C is homogeneous
( a) and inhomogeneous (b)

involve <j>-rotations can not be used, since, for such rotations, the intersection of
the parallel beam with the crystal will move to a new surface direction Sci>'
(Fig.5.8) , and the points along this new direction may not have the same strains as
the points along Sci>' Depending on the particular experimental arrangement and
beam dimensions used in a given experiment, similar problems are possible with
other rotations.
Another point that must be considered in the analysis is the presence of strain
gradients with depth. In such cases, the d spacing, and thus the strain obtained by
the x-ray beam will be the average of this gradient over the effective penetration
distance of x-rays, t. This distance is limited by absorption and depends on 29,\1'.
In Sect. 4.20, it was shown that the diffracted intensity from a unit volume located
at a distance, Z, from the surface can be written as
dI '" e - "ldV ,

(5.35 )

134

5 Determination of Strain and Stress Fields by Diffraction Methods

tS3
I

Incident
Beam

Diffracted
Beam

Diffracted
--H- -

S2

1-

Scp'

~I

Fig. 5.8. Variation of the measurement direction and diffracting points in parallel beam singlecrystal diffractometry with a rotation (<l around ~h

where I is the total path length of x-rays within the specimen and ~ is the linear
absorption coefficient. For the a-goniometer (Fig. 4.27), the path length for any
\jJ-tilt is given by

{ 1

1=.

sm(9+\jJ)

1}

+ sm(9-\jJ)
.
'z,

(5.36 )

which can be written as

(5.37)
Similarly, for the \jJ-goniometer (Fig. 4.28):
2
z.
sm 9 cos \jJ

1= .

(5.38 )

The effective penetration depth is defined as [22]:

(5.39)
Thus, the effective penetration depth for the a goniometer is given by [22]
(5.40 )

135

5.9 Average Strain Data from Single Crystal Specimens


~

Goniometer
Goniometer

5
6.

4
6

u;

c:

e
0

'E
Eo<

(T 8 )""0

(micron)

2.

O.
O.

0.3

0.6

sin'v

20

60

100

140

Fig. 5.9. a Variation of the penetration depth with sin 2 ", for Q and", goniometers (29 = 1550 , Cr
radiation on Fe, 211 reflection). b Variation of the maximum penetration depth 1: ('" = 0) with
29. (Fe specimen, Cr radiation)

and by
sin e costp
211

't=---~

( 5.41 )

for the tp goniometer. In Fig. 5.9 a,b, the variation of penetration depth with tp
(at constant 2e), and with 2e (at constant tp) is shown.
Within this depth, the diffracted intensity from points deeper into the material
is smaller than the contribution from the points that are closer to the surface
because of absorption (Sect. 4.25). Consequently, the average strain obtained by
x-rays will contain a larger contribution from the strains in the shallower depths.
The average strain over the total penetration depth ('t1p) can be expressed as

JCij ( Z ) 'e - z/T"'dz

T",

<Cij) ci>1p = -O-'----::-T"'----Se-z/T"'dz

(5.42 )

where e - ziT"" the weighting function, describes the intensity diffracted from
depth z.
Since the path length of x-rays within the material, and hence the effective
penetration distance, is dependent on the angles e,tp, a different portion of the
gradient will be sampled for different reflections. This means that the average
strains referring to the sample coordinate system, <cij), are also dependent on e,tp,
and the final strain tensor, obtained from the least-squares regression is a
complicated average of all the <cij)o,1p' If, for all rotations, the penetration depth is
greater than the thickness of the specimen (for some thin-films for example) , the
total gradient is sampled at all times and this effect may not be important.

136

5 Determination of Strain and Stress Fields by Diffraction Methods

However, in relatively thick specimens that may contain steep gradients, OEij/OZ,
within the irradiated volume, this effect may cause deviations from regular
behavior in "d" vs. sin 2tp, and may cause errors in the results obtained from the
least-squares analysis. The effect of such gradients with depth will be discussed in
detail later in the text (Sect. 5.12) .

5.10 Interpretation of the Average X-ray Strain Data


Measured from Polycrystalline Specimens
A polycrystalline specimen, whose grain size is very small compared to the
irradiated volume, contains a number of grains which can diffract at a given tp-tilt.
The lattice spacing determined from the angular position of the maximum
diffracted intensity will be an average value of all these grains. When the specimen
is tilted to a new tp (Fig. 5.1 0) , diffraction occurs from the same family of planes,
but from different grains. This means that the average strain values calculated
from Eq. (5.1 ) are defined in mutually exclusive subsets of the irradiated volume.
The average strain at any tp-tilt is related to the actual strain at a point A (x,y,z) by
[23J

J f(z)dV

(E jj \,. = ----v.cN:=-------

L Vk

(5.43 )

k=l

where Ejj(X,y,z) is the strain at the point, A(x,y,z), in a diffracting grain in the
irradiated volume, Vk is the volume of this grain, f(z) is a weighting function
relating the variation of diffracted intensity to depth (z), and Nk is the total
number of grains diffracting at a tilt angle tpl. The summation is required since the
diffracting grains are not necessarily contiguous. Equation (5.36) describes the
average strain (Eij)q,. in the population of grains diffracting at tpl.
At any tp-tilt, the average strain (E~3)q,1p is related to the average strains in the
sample coordinate system through Eq. (5.4):
o
(E33, >cj>1p-_ (dcj>1p)-d
do

= {(Ell )1pcos 2<1> + (E 12 )1psin2<1> + (E 22 )1psin 2<1> - (E 33 )1p} sin 2 tp

+ (E 33 )1p + {(E 13 )1pcos <I> + (E 23 )1psin <I>} sin2tp,

(5.4b)

where the terms (Eij)1p are the average strains in one population only. If the
experimentally determined (Ej)1p values vary regularly with sin 2tp, (Figs. 5.2a,b),
this indicates that the (Ejj )1p is equal in all the populations sampled during the tptilts, i.e., the average strain tensor {(Ej)1p} is homogeneous in these populations.
Non-regular (oscillatory) behavior, on the other hand (Fig. 5.2c), indicates the

5.11 Interpretation of Average Stress States in Polycrystalline Specimens

137

28

::-L~ ----~
iii

1"1

I \

I \

28

Fig. 5.10. Diffracting grains in a polycrystalline specimen at (a) tp =0, (b) tp =tp, during a stress
measurement. The x-rays are diffracted from mutually exclusive subsets of the total irradiated
volume at each tilt

presence of an inhomogeneous average strain state; for each diffracting subset, the
strain <E;j)", along S; is different. Interpretation of such inhomogeneity in terms of
local strains is also complicated by the fact that the diffracting subsets are not
distinct homogeneous volumes in real space. Even though diffraction can easily
separate the grains' into the respective subsets, the grains for all subsets are
intimately mixed in the material. A grain that belongs to the population diffracting
at 'l'='l'A may be surrounded by others that diffract at another 'l'-tilt. Thus,
in going from one grain to another in a given subset, grains that belong to other
subsets must be traversed. Therefore, even though diffraction obtains data from a
given subset as if S; is continuous in that subset, the regions along S; that
contribute to the diffracted intensity are discontinuous in real space.

5.11 Interpretatjon of Average Stress States in


Polycrystalline Specimens
In the preceding section it was shown that experimentally determined regular d vs.
sin2'l' plots indicates the presence of a homogeneous strain state within the
diffracting volumes sampled during the experiment. In addition, if for all 'l'-tilts

138

5 Determination of Strain and Stress Fields by Diffraction Methods

the diffracting population consists of a sufficient number of randomly oriented


grains 1, the average elastic constants of these volumes will also be isotropic as
discussed in Sect. 3.9. In such a case, the material may be treated as a quasiisotropic material and depending on the shape of the d vs. sin 2tp plot, the methods
of stress determination applicable to isotropic materials [Eqs. (5.5), (5.18), etc.]
may be used in the stress analysis.
The average stresses obtained from these analyses may contain contributions
from macro and micro residual stresses, as well as contributions from any stress
that is being applied during the measurement, since all of these effects cause
changes in the plane spacing. The contributions from the macro and micro stresses
may be identified by considering the equations of equilibrium (any applied
stresses may be treated as macrostresses as described in Sect.3.2).
Consider a ground steel specimen that exhibits a tp-split "d" vs. sin 2 tp plot which
indicates that (in addition to the strains in the plane of the surface) E13 and/or E23
must be finite [1]. For isotropic materials, these strains indicate the presence of the
shear streses 0'13,0'23. However, as discussed by Van Baal [24] and Brakman
[25] , it can b.e shown from a consideration of the symmetry of deformation, that
0'13 and 0'23 cannot be macrostresses as this would violate the equations of
equilibrium (2.58), (2.59) (see Sect. 3.5 for an equivalent discussion On shotpeening). The presence of E13 ,E23 may also imply the presence of monoclinic or
lesser symmetry in the surface layers, since for such anisotropy, Hooke's law
[Eq. (5.9)] yields
(5.44 )
where Ejj'O'ij,Sjjkl refer to the sample coordinate system. Thus, E13 (and similarly
E23 ), may be caused by the stresses in the plane of the surface. Steel, on the other
hand, has a cubic or tetragonal unit cell, and in order to have monoclinic
symmetry in the surface, the grain distribution must be highly non-random (i.e.,
the surface must exhibit texture). Grinding, however, does not cause appreciable
texture [26], and this second explanation is not adequate in this case. Furthermore, Armco iron ground with the same grinding parameters does not exhibit
tp-splitting. These facts imply that in the ground steel specimens [which are
surface-deformed two phase materials (Sect. 3.7) ], the stresses measured from
the ferrite phase contain contributions from the microstresses created due to the
inhomogeneous distribution of plastic deformation between the ferrite and the
carbide phases: This explanation also accounts for the absence of E13 and e23 in
Armcoiron, since this is a single-phase material.
If residual stresses are measured from both phases of a two-phase materiaF, it
is possible to separate the average stresses measured by x-rays into the
1 This assumption may be checked experimentally during a stress measurement. One obtains the
integrated intensity under the peak (corrected for LP and absorption factors) for all ",-tilts.
The integrated intensity at any tilt is proportional to the total diffracting volume for that tilt
( or to the number of diffracting grains) and should not fluctate more than 20% over the total
",-range of the measurement. In many experiments, the peak shape does not change
appreciably with ",-tilt. Thus the maximum intensity corrected for absorption and LP factor
may also be used instead of the integrated intensity.
2 This implies that the d vs. sin2 ", plots from both phases are regular.

5.11 Interpretation of Average Stress States in Polycrystalline Specimens


100

139

MPa

0
-100
-0
U~ - D

"

(1:'

-500

30

60

90

z-

120

150

~m

180

200

MPa

100

[]

c
c

+ +

c
[]
~

+ ~

[]

---- -~----

Vi

-200

-L.OO
b

(T_ - )(

x
x x

-300
0

-100
III
III

Fig.5.11. a Variation of the total


residual stress crt with depth in (l and
~ phases in a shot-peened 60-40
brass specimen. (In this study the
deeper layers were exposed to the xray beam by electropolishing. )
b Data shown in a separated into
macro and average micros tress
components in the respective phases
as a function of depth. (From
Noyan [27])

30

w w

60

,
x

o'-

90

u.''''
~

z-

u:- -

120

150

- 0

~m

180

contributions due to the microstresses and macrostresses [27]. The average stress
obtained from any phase [Equation (5.46)] may be written as
(5.45 )
where O"~ is the macrostress and <O"ij>~m is the average microstress (PM stress) in
that phase. For symmetric surface deformation O"g3'0"?3,O"g3 will be zero as noted
above. It can be seen that for the stress components 0"11,0"22,0"12 x-ray
measurements yield'two equations in three unknowns. For example, for 0"11:

<0" 11 >~ = O"? 1+ <0" 11 >~.m


(5.46 )
<0" 11 >p = O"? 1+ <0" 11 >g.m .
A third equation linking <cr 11 >~.m,< cr 11 >g.m may be obtained from the equation of

equilibrium for mictostresses (Sect. 3.5) :

(5.47 )
where f is the volume fraction of the second phase. Thus, one may solve for
<cr11>~m,<cr11>gm, which are the contributions from the microstresses and cr?1
which is the macrostress due to the constraining effect of the bulk.

140

5 Determination of Strain and Stress Fields by Diffraction Methods

This procedure requires accurate values for the unstressed plane spacing "do"
of both phases. However, even when these values are not available, it is possible to
separate the macrostresses cr? 1,crg 2 from the contributions due to micros tresses
[27]. For example, the slopes of d vs. sin 2 tp from both phases of a shot-peened
alloy are given by [Eq. (5.24)] for <l> = 0:

,'

mel = (do ),,' ( 1 ~v) {cr? 1+ <cr11)~m- <cr33)~m}


1+
m ll -_ (d 0 ) II . ( E

v) .{ocr + <cr
II

ll

ll

)p.m
_ <cr33 )P.m}
II
II

The total microstess contributions cr 11 )~.m Eq.5.47):

(5.48a)
(5.48b)

<cr33)~m)

must also obey


(5.49 )

Thus, Eq. (5.48a,b) and (5.49) may be solved for the macrostress term cr? l ' and
the total micrestress terms in r:t and ~. In Fig. 5.11a,b the measured and separated
residual stress profiles from both phases of a shot-peened two-phase brass alloy
are shown. The second phase contains appreciable microstresses in addition to the
macros tresses imposed by the bulk.

5.12 Effect of Stress Gradients Normal to the Surface


on d vs. sin,., Data
The terms cr 3j used in the previous equations are stresses normal to the surface.
These stresses must, by definition, be exactly zero at the surface of particles of
either phase. Thus, cr 3j can exist only as gradients in the near surface layers.
Presence of such gradients in cr 3j' as well as the variation of cr 11 ,cr 22' or cr 12 with
depth causes curvature in the d vs. sin 2 tp plots. Such non-linearity occurs due to the
changing effective depth of penetration of x-rays with tp-tilt; the average strain
state being sampled by the x-ray beam is different for each tp-tilt. If the gradients in
the diffracted volume are single valued, monotonically varying functions of depth,
the progressive sampling of different depths will introduce bias in the dtp value
determined by x-rays. The effects of such gradients have been treated in detail in
Refs. [7, 26, 28].
Assume that the behavior of the stress crij with depth follows a power law 1:
(5.50)

where crij (t = 0) is the stress value at the surface, aij,n ij are constants over the
depth of penetration, and z is the distance coordinate along S3' measured into the
material.
1 A power law is assumed to represent the gradient in stress in the following discussion for
simplicity. For a more general discussion the gradient may be expressed in any functional form.
The general arguments, however, will be similar as long as the gradient is a monotonically
varying, single valued function of z.

5.12 Effect of Stress Gradients Normal to the Surface on d vs. sin 2 \jl Data

141

1-1

1-2

2.873

.. I-3

2.870

0
~
jl

c:t
52.867

i!

II

"0

2.864

iii

0.0

0:4

0.8

Fig. 5.12.- Curvature in "d" vs. sin 2 \jl plots


caused by 0"33 (z). (This plot is from a
computer simulation. It was assumed that
in a steel specimen 0"11 = 0"22 = -400
MPa, and <0"33>=100 MPa, with
n33=1,2,3 for 1-1, 1-2, 1-3
respectively.) (From Noyan [27])

sin2 lY

The average sttess determined by x-rays at any tp-tilt is, from Eqs. (5.42) and
(5.43) ,
(5.51 )

where Kij is a constant, nij is the exponent of the stress variation and 'tlP is given by
Eqs. (5.40, 5.41).
Thus, if the stresses are functions of depth, instead of constant valued
unknowns, expressions of the form given by Eq. (5.51) should be substituted for
the stresses in Eq. (5.15). For example, for a stress tensor of the form
(

0"11 0 0
o 0"22 0 "
o 0 0"33,

where 0"11 =0"22=Ff(z),

0"33=K33't~33,

n 33 =F0, Eq. (5.15) becomes

(dcj>lP)-d o 1 +v
. 2
2v
K33't~33 {
. 2
}
do
= E .0" 11 . SIn tp - E cr 11 +
E
( 1 + v )( 1 - SIn tp) - v .

(5.52 )

Here the first part of the equation is equivalent to the classical biaxial stress
Eq. (5.17) and exhibits linear dependence versus sin 2 tp. The second part (which is
the x-ray average of the 0"33 gradient for the particular tp-tilt) is not linear in sin 2tp.
In case of single valued, monotonically varying gradients, this term increases in
magnitude systematically with decreasing tp (increasing 't1), and thus causes a
progressively larger deviation (curvature) in the d vs. sin tp plot. If d vs. sin 2 tp
does not exhibit tp-splitting, such curvature can be easily observed (Fig. 5.12). In
such cases, the use of high tp-points for the linear regression analysis will minimize

5 Determination of Strain and Stress Fields by Diffraction Methods

142

18

2.8750

'10 4

40
14

30

12

.0/1> 0

20

10

,0/1<0

2.8700

6
4

-;..

-e-

1::J

70

60

-2

50

-4

40

-6

2.8630 0

0
b

8
'10 4

.. '" >45
'" ~45

6
5
4

3
2

0,2

0.4
0.6
0.8
sinI2if;1-

Fig. 5.13. a tp-split "d" vs. sin 2tp plot. The


curvature in both arms of the plot is due to
the term sin2tp in Eq. (5.4) and the effect
of gradients in the S3 direction. b a 1 vs.
sin 2tp plot, c a 2 vs. sin2tp plot calculated
from the data shown in a. In these plots the
curvature, or deviation from linearity, is
due only to gradients in z. (From Moyan
[27] )

any curvature errors associated with the data, since the absolute value of the '1'dependent terms will be less in shallower depths. It must also be noted that the
other terms ofthe stress tensor such as a ll ,a22 may also possess gradients along
the z-direction, and thus may cause curvature in the d vs. sin 2 tp plot.
Similar effects must also be taken into account when the d vs. sin 2 tp exhibits '1'splitting (Fig. 5.13a). In such cases the stress terms a 13,a 23 may be finite, both of
which must exist as gradients in the irradiated volume, and the a 1 vs. sin 2 tp, and
a 2 .vs.sinI2tpl must be calculated [Eqs. (5.22a,b)] and plotted (Fig. 5.13b,c) in
order to separate the effects of curvature due to gradients from the curvature
caused by the trigonometric term sinl2tpl in Eq. (5.15). If any of the normal
stresses (ajj ) or the shear stress (a 12) exhibit gradients with depth, the behavior
of the term a 1 vs. sin 2 tp will be similar to d vs. sin 2 tp (Fig. 5.12, 5.13b), and to
minimize errors due to curvature high tp-points must be used. However, the

5.12 Effect of Stress Gradients Normal to the Surface on d vs. sin 2\jJ Data
MPa

400 f-

200

143

~1'~2

.. ~3>t/f
(033)t/f

t-

oC=::=~~---J
'b

-200 -

-400t~----~--------------

I
4

z-

2 IJ.m

Fig. 5.14. The residual stress profiles with


depth used in the computer simulation of
the "d" vs. sin 2\jJ plot shown in Fig. 5.13

argument sin 12", 1in a 2 is a multivalued function of", over", = 00 ,60 0 (which is the
commonly available ",-range for experiments). This, coupled with the fact that
different regions of the gradient are sampled at each ",-tilt, causes two branches in
the a 2 vs. sinl2",1 data (Fig. 5.13c). In such cases, either the linear part of a 2 vs.
sinI2",1, or non-linear regression over the entire ",-range must be used to obtain the
stress values. In Fig. 5.14 the average stress profiles used in simulating the d vs.
sin 2", in Fig. (5.13a) are shown. In Table 5.3, the solutions obtained from the
traditional methods in various ",-ranges are summarized. From a comparison of
the input values (Fig. 5.14) and the analysis results (Table 5.3), it may be seen
that gradients, if not properly taken into account, may cause large errors in the
stress results obtained from x-ray analysis.
Table 5.3. Analysis of ~l vs. sin 2\jJ, a2 vs. sinI2\jJ1, (Fig. 5.13b,c) over various \jJ-ranges. (These
plots were calculated from the input profiles shown in Fig. 5.14.)
\jJ-Range
(degrees)
al

0.00 -60
0.00 -33.21
39.23 -60

0'22
(MPa)

-442
-562
-412

-442
-562
-412

O'll

(MPa)

Error in
0'1 dMPa)*

(al3>

(al3>

107
55
81

42
162
12

50
86
278

0
0
0

(MPa)

96
-182
183
is defined as

(a 33 >

(a23>

(MPa)

0.00 -26.57
33.21 -45
50.84 -60

The error in

(MPa)

<al3>

\jJ-Range
(degrees)
a2

O'll

0
0
0
(O'll

(MPa)

(MPa)

Corr. coeff. of
a2 vs. sinl2\jJ1
0.9952
-0.913
0.992

)anaIYSis-O'~h where O'~ 1 is shown in Fig. 5.14

144

5 Determination of Strain and Stress Fields by Diffraction Methods

In the discussion so far, the effects of gradients on d vs. sin 2 ", has been discussed
with particular emphasis on the errors they cause in x-ray stress determination.
Determination of the gradient in stress with depth is also an important
measurement, extensively used both in research and industrial testing [4.29].
Various methods are available for the determination of the stress profile with
depth. If the stress proflle within 10- 20 microns of the surface is required, one
may use various radiations, penetrating to successively different depths, to
determine the variation of the average stress state with depth. For example, in
steels, one can use Cr, Co, Mo radiations with maximum penetration depths
('" = 0) of 5.4 microns (211 reflection), 10.5 microns (310 reflection) , and 16.2
microns (732 - 651 reflections) respectively. On the other hand, interpretation of
data obtained by this method is not easy; the strains in the shallower depths are
also integrated by the x-ray beams penetrating to deeper layers. It is also possible
to utilize the curvature in d vs. sin 2 ", to determine the shape of the gradient by
curve fitting techniques. Several combinations of gradients in <1ij , however, may
cause the same curvature in d vs. sin 2 "" and the functional form (i.e. power series,
Fourier series,'cetc.) assumed to represent the variation of the components of the
strain/stress tensor with depth and the uniqueness of the solution obtained from
curve fitting must be carefully checked in order to obtain the true stress gradient.
The methods discussed above are limited in penetration depth by absorption.
Thus, they cannot be used to study the stress through the depth of the specimen.
This can be accomplished, however, by first measuring the stress at the surface of
the specimen and then exposing layers deeper into the material by removing the
surface through electrochemical or chemical polishing and repeating the stress
measurement. This is repeated until stresses through the depth of interest have
been examined. Mechanical methods, such as grinding, should not be used for
layer removal since they introduce residual stresses of their own and may cause
misleading results. If a mechanical method has been used to remove a large
quantity of material, the specimen should be electropolished to depths not affected
by the plastic deformation caused by the process. This permits access to the
original deformation distribution in the material.
Layer removal, whether by chemical or mechanical methods, causes rearrangement of the elastic strain/stress distribution within the material. For
example, if the plastically deformed surface layer in a shot-peened specimen is
removed by electropolishing, the residual stresses in the bulk will decay to zero
since the material responsible for the elastic constraint no longer exists (Sect. 3.2) .
Thus, x-ray stress values obtained from all polished surfaces must be corrected for
stress relaxation effects in order to obtain the correct strain/stress distribution in
the specimen. Such corrections may be obtained from elasticity theory, and will be
discussed for several geometries in Chap. 6. It must be noted here that electropolishing affects microstresses due to plastic deformation (Sect. 3.4) in a different
way; for example, if the grains of two phases mutually constraining each other are
polished away at the same rate, the volume fraction of both phases at the surface
will not change, and thus no relaxation of the micros tress in either phase will
occur. If, however, preferential etching removes one phase faster than the other,
relaxation of microstresses can also take place.

5.13 Experimental Determination of X-ray Elastic Constants

145

5.13 Experimental Determination of X-ray Elastic Constants


The elastic constants used in the calculation of average stresses from average
strains also should be average values, where all the averages are taken over the
same volume. These average values may be calculated from single crystal elastic
constants by procedures discussed in Sects. 2.9, 2.10 when the material is
completely random. These theories are not adequate in the presence of texture,
however, and x-ray elastic constants S 1 ,S2/2 measured from a material identical to
the one undergoing stress measurements must be used for better accuracy. (A list
of some values from the literature is given in Appendix F. If high accuracy is not
needed, these may be useful.)
In the experimental determination of elastic constants a specimen is stressed on
a diffractometer by the application of known loads. In Fig. 5.15 a microprocessor
controlled tensile device capable of automatically determining elastic constants to
a user specified error is shown. Uniaxial tensile loading or four-point bending is
usually employed in x-ray elastic constant determination since the average stress
distribution caused by these processes can be determined exactly from elasticity
theory.
Consider the case where a polycrystalline specimen is loaded in uniaxial
tension. The strain at any point in the irradiated volume may be expressed as the
sum of two components:
(5.53 )

Fig. 5.15. A microcomputer controlled tensile device used in experimental determination of x-ray
elastic constants. A, tensile specimen; B, grips; C, diffractometer track; D, stepping motor;
E. load cell; F, gears; H, micrometer adjustment used in placing the diffracting volume over the
diffractometer surface. (From Perry et al. [371)

146

5 Determination of Strain and Stress Fields by Diffraction Methods

Here E~ is the homogeneous elastic strain that would be observed if the stress was
applied to a homogeneous isotropic material. The reaction strain component, E;j,
arises because of the variation of elastic constants along a given direction in the
surface plane of the sample (Sect. 3.4 ). The reaction strains depend on the
difference between the displacements, U b that have to be made compatible at a
point, and vary from point to point since the effect of surrounding grains at each
point is different. However, as long as the total stress at a point is lower than the
microscopic elastic limit at any point, the magnitude of the reaction stresses and
strains at any point will be directly proportional to the applied stress.
Thus, if a homogeneous stress CJ~l in the Sl direction is applied at the
boundary of the material, the components of the strain tensor at any point A in the
material may be expressed as

{1.

Ell (x,y,z) =CJoll ' E +Kl (x,y,z)

( 5.54a)
E13 (X,y,Z) =CJ~lK~(x,y,z)
E23 (x,y,z) = CJ~l'K~ (x,y,z)
Ell (x,y,z) =CJ~l'K~(x,y,z) .
Here x,y,z are the coordinates of the point A with respect to a coordinate system
describing the surface, and K; (x,y,z) are the proportionality constants at A
between the applied load and the resulting reaction strains. The constants
K~,K~,K~ may be finite depending on the symmetry of the constitutive grains of
the material or the symmetry of texture in the surface layers. Also, in certain cases
the local symmetry changes as one approaches a grain boundary. In the following
discussion it is assumed, for simplicity, that K~=K~=K~=Ol.
If the material also has an initial residual stress distribution caused by the
previous deformation history, the total elastic strain at a point, El j, will also have a
residual strain component in addition to the strains caused by the applied stress.
Thus, for this case, Eqs. (5.54a) become
(5.54b)
where Efj(x,y,z) is the residual strain at the point A(x,y,z).
Now assume that this material is placed (applied stress and all) on a
diffractometer and a beam of x-rays (whose edges extend from Xl to X2 along
Sl,yl to Y2 along::h and which penetrates a distance 't along S3 into the material)
1 If these terms are finite, the discussion follows along the same lines but the equations must be
modified to take into account the shear strains &13'&23'&12'

5.13 Experimental Determination of X-ray Elastic Constants

147

is used to measure the average strain along (L3 ) cjJ1jJ' From Eqs. (5.54b) and (5.4)
one obtains (<l> = 0),
(5.55 )
where the average strains, <Ei)~' in Eq. (5.55) are related to the strains at a point
through equations of the form of Eq. (5.43). [It must also be noted that because
x-rays penetrate to a different depth (z) for each tp-tilt due to absorption and, for
finite beam sizes, the intersection of the tilting specimen with the beam changes the
dimensions Xl -X 2'YI -Y2' the total irradiated volume.is also a function of the tilt
angle "tp". This means that parameters that are a function of volume such as K can
also be expressed as a function of tp.]
The evaluation of the average x-ray strains in terms of the strains at a point is
extremely complicated since the exact coupling terms in Eq. (5.54a) are not
known. However, by inspection of Eqs. (5.43), (5.54a,b), it can be seen that the
average strain may be written in terms of the average strain components
contributing to it:'
(5.56 )
or, in terms of stresses:

<E11)t=cr?I'{~ +KI(\jJ)} + <Ell)'


(5.57 )
<E33 )t=cr?I' { - ; +K3 (tp)} + <E33)"
Similar equations may be written for the other terms of the strain tensor. Here
Kj ( tp) is the average proportionality constant describing the average response of
the population of g~ains diffracting at tilt angle tp to an applied load, and <Ej)r is
the average residual strain in this population.
By substituting Eqs. (5.57) into Eq. (5.55) the relationship between the
applied stress and the measured strain is obtained:
< . ) _ <~)-'do
E33 1jJdo
= [cr?l'

{1 ~v

+KI (tp) -K 3(tp)} +<E;I)1jJ-<E33>~lsin2tp


(5.58 )

A linear least-squares line fitted to the data described by Eq. (5.53), (as is usually
done in practice), implicitly assumes that the function,
(5.59 )

5 Determination of Strain and Stress Fields by Diffraction Methods

148

describes the relationship between d and sin 2tp. Here e is the random error
component. The regression parameters ~O'~l are then obtained from the equations
[30]
( 5.60a)

(5.60b)
where n is the number of tp tilts and
n

L sin2tpj

-'-2-

sm

j=l

tpj=~--

d",.
=
TJ

L ~J

j= 1 ':

(5.61 )

~-=-----

It must be noted that, if d vs. sin 2tp is oscillatory, the parameters ~O'~l cannot be
called "slope" and "intercept" of the least-squares line in the traditional sense
(e.g. as in the conventional analysis of linear d vs. sin 2tp plots). Here they are
simply mathematical functions described by Eqs. (5.60a,b)].
Substituting Eq. (5.58) into Eq. (5.60a) one obtains

~1 =~'{0"?1
L cf

.t

Cj[{( 1

~v) +K1 (tpj) -K3 (tpj)} sin2tpj+K 3(tpj)

)-1

j= 1

+ J.L= 1 Cj[{<ell>~J-<e33>~Jsin2tpj+<0"33>~J+1
n

- .L

J=1

{<ell>~J - <e33>~J sin 2tpj + <e33>~J + 1]} ,

(5.62 )

(5.63 )
where the function 82 [ (1

~ v) ,Kj (tp) ]

is an average [defined by Eq. (5.62)] of

the terms 1 ~ v ,K 1( tp ) ,K3 ( tp ). The term


term.

F ( <ejj)")

is an average residual strain

5.13 Experimental Determination of X-ray Elastic Constants

149

A similar equation can be written for ~o. However, this case is more
complicated and will not be treated here.
Equation (5.63) predicts a linear variation of the parameter ~I with applied
load cr? l ' This linear variation is independent of the shape of d vs. sin 2\jJ
(regular or oscillatory behavior) and thus of the distribution of residual
strains/stresses (homogeneous or inhomogeneous) in the irradiated volume.
Thus either linear or oscillatory data (Figs. 5.16a - d) will yield linear ~1 vs cr? 1
(Fig. 5.17). On the other hand, the cause of oscillatory behavior (Kj (\jJ) or
<Eij>~) is important in the interpretation of the x-ray elastic constants SI,S2/2.
In the case of a homogeneous, isotropic material, where the elastic interaction
terms K j ( \jJ) are equal to zero by definition, S2/2 will be equal to ( 1 ~ v )
independent of the particular distribution of residual strains, <Eij>~' in the
irradiated volume since these terms affect only the intercept term, F ( <Eij>~), in
Eq. (5.63). In the case of quasi-homogeneous materials (where all the grains are
distributed randomly) , such that along any direction all possible crystal directions
are encountered" and the same average elastic constant is measured, two
possibilities exist. In the first case, the elastic constant, S2/2, may be equal to
1 ~ v + Kj ( \jJ ) , where Kj ( \jJ) is an average interaction term that is the same in all
directions. Alternatively, the interaction terms may integrate to zero such that
S2/2 = 1 ~ v, where 1 ~ v is the average of the single-crystal constants. In all
cases, however, the term S2 will be isotropic in the irradiated volume. For inhomogeneous materials, Kj ( \jJ) are finite and inhomogeneously distributed
in all subsets. For such materials, the x-ray elastic constant S2/2 is equal to

-S2 (1E'
+v K (\jJ) ) gIven
.
by Eq. (5.62).
j

The terms K j ( \jJ) that may appear in S2/2 for inhomogeneous or quasihomogeneous materials are configurational parameters that are dependent on
both shape distribution of the grains and the texture in the irradiated volume.
Thus, when these terms are finite, S2/2 is a measure of the average elastic response of
the inhomogeneous material to an applied load cr? I' In these cases, S2/2 is not an
elastic constant (like Cjjkl,E,v) in the strictest sense. Rather, it is an "average
effective elastic constant" of the diffracting volume of inhomogeneous material.
Comparison of such values with values obtained from quasi-isotropic materials
can yield information about the deviation from "randomness" that exists in the
diffracting volume. The average nature ofS 2/2 in this case must also be emphasized.
It can be seen from Eq. (5.62) that both KI (\jJ) and K3 (\jJ) contribute to S2/2.
Furthermore, these terms themselves may not be constant in \jJ and vary from one
diffracting subset to another (this is the case if the oscillations are caused by
plastic deformation rather than <Eij>~!)' Thus, the interpretation of such
constants in terms of local reactions and elasticity theory is (in most cases)
difficult.
It must be noted that the linear response predicted by Eq. (5.63) is possible if
and only if no plastic deformation occurs during the elastic testing; plastic
deformation may cause total or partial relaxation of the elastic reaction strains at

1.3067 l-

.:: 1.3070
a
~ 1.3069
:;; 1.3068

1.3072
1.3071

"01.3067

~
1.3068
N

1.3069

1.3070

"01.3067

~1.3068

1.3069

1.307,0

0.1

ll.

b.

b.

0.2

b.

b.

b.

sin 2

0.3
PSI

b.

b.

b.

,.

b.

b.

I-

0::::7.4 Mpa

0.4

0.5

0.6

0::::97.2 Mpa -

b.

b.

0::::49.7 Mpa

b.

b.

1.2041

~ 1.2043
:;; 1.2042

1.2045 l-

1.2046

1.2041

.:: 1.2044

"0

~1.2042

1.2044
~ 1.2043

1.2045

1.2046

;:: 1.2041.
~ 1.2043
"01.2042
1.2041

1.2045

1.2046 l-

0.1

b.

b.

b.

0.2

ll.

b.

b.

0.3
sin 2 PSI

b.

b.

b.

"T

b.

0::::7.2 Mpa

ll.

b.

0.4

0.5

0::::9S.2Mpa

b.

0::::47.6 Mpa

b.

b.

...J

-I

'"

s::

=:I

a.

'"

>Tj

'"'"

Cf.l

8-

~.

Cf.l

o
...,

i.

VI

VI

1.1711

1.171L.

l-

l-

0.1

I!.

I!.

I!.

0.2

I!.

I!.

I!.

0.3
sin 2 PSI

I!.

I!.

I!.

I!.

I!.

O.L.

0.5

a==209Mpa

I!.

0.6

a==S.4Mpa-

I!.

a==10SMpa

I!.

I!.

I
<!

1.1700

1. 1696

1.1697

1.1698

1.1699 I-

1.1702

1.1696

1.1697

1.1698

-0

1.1796 I-

1.1798

:: 1.1700

-0

:: 1.1699

-0

1.1700 l-

0.1

I!.

0.2

I!.

I!.

0.3
sin 2 PSI

I!.

I!.

I!.

I!.

O.L.

0.5

a==192.Mpa

I!.

I:J.

a==10S.Mpa

.g

Ul

;...
w
tIl

0.6

Ul

--

!il"

(J

~
o.
o

g,

~.

~.

1- ~
a==S.Mpa

I!.

I!.

Fig. 5.16. "d" vs. sin 2 '1' plots at various tensile applied loads from (a) Ct-brass, (b) l3-brass, (c) 1008 steel, (d) 1075 steel. The specimens
were loaded in situ on the diffractometer with the apparatus shown in Fig. 5.15. (From Noyan [27])

1.1708 I-

U 1.1710 l-

::: 1.1712

<!

1.1708

1.1709

U 1.1710

1. 1712

<!

1.1713

1.171L.

1.1709

U 1.1710

1.1712

1.1711

<!

1.1713 l-

1.171L.

152

5 Determination of Strain and Stress Fields by Diffraction Methods


I

1l.1l1l16

1l.1l1l12

,..
~

)( +

B.IlIlIlB

II

II

1l.1l1l1l4
0

If.

)(

II

C>

1l.1l1l1l1l

a-Br.
p-Br.
1008

+1075

1
Sil.

100.

1SIl.

RPPL. STRESS (MPR)

Fig. 5.17. Variation of the regression parameter


vs. applied loa&,

~1

21l1l.

(determined from analysis of Fig. 5.16a-d)

the particular load, destroying the linear relationship between the applied load
and the reactien strainsl. The effect of plastic deformation, however, is not limited
to relaxation: extensive plastic deformation may cause changes in grain shapes and
grain orientations (texture) such that the elastic interaction components, K; ( tp ) ,
before and after deformation are no longer identical. Thus, if the "x-ray elastic
constants" of a specimen are first measured (with loads below the elastic limit as
discussed), then the specimen plastically deformed and the measurement
repeated, the interaction constants K; ( tp) and, thus, the x-ray elastic constants,
may change as a result of the plastic deformation. In the literature, changes up to
40 % have been reported for steels deformed in uniaxial tension [31]. These
changes are due to the changes in K; ( tp ), since E, v are not affected by plastic
deformation as discussed in Sect. 2.9.
If the material in the measurement volume also contains residual strains,
<Eij>~, the effect of plastic deformation on these strains must also be considered.
For example, if the local plastic deformation is such that the inhomogeneous
residual strains, <E;j>~' change, but the coupling constants are not affected, then
no change in th,e "x-ray elastic constants" will be observed even though the shape
of the oscillatory d vs. sin 2 tp plot changes. Even though such a residual strain (or
stress) distribution causes oscillations in d versus sin 2 tp, the residual strain
distribution itself is not a function of the elastic loads applied (after plastic
deformation) and will affect only the intercept of the ~l vS. cr? 1 plot [Eq. (5.63)].
Thus, it is not possible to determine the extent of the interaction effects on the
elastic constants just by observing the magnitude of the oscillations.
One other conclusion from this line of reasoning is that even if the oscillations
in d vs. sin 2 tp plots from two samples of the same material look alike, unless the
shapes of any precipitates that might be present and the distribution of
1 This property may be used to determine the microscopic yield point since the load at which
vs. cr? 1 deviates from linearity indicates non-elastic strain effects.

~1

5.14 Determination of Stresses from Oscillatory Data

153

crystallographic directions in the constitutive grains of the matrix along surface


directions are the same for both specimens, and in both cases the oscillations are
due to the same effect (caused by Ki (tp) or <Eij>~)' they may not have the same
"x-ray elastic constants". In fact, for oscillatory d vs. sin 2tp behavior caused by
interaction effects, such effective elastic constants may be different along different
surface directions of a given specimen.

5.14 Determination of Stresses from Oscillatory Data


In the previous sections it was shown that oscillations in d vs. sin 2tp indicate the
presence of an inhomogeneous stress/strain state within the material, in which case
the total strain measured by x-rays may contain conributions from E~,<E:j'>,<Eij)'.
Expanding Eq. (5.58):

o}

V
1 +V 0 . 2
{T
a l l sm tp- E all

+ { a? 1 [Kd tp) -

+ {[ <Ell>~ -

K3 (tp)] sin 2tp + a? 1 K3 (tp)}

<E33>~]sin2tp + <E33>~}

(5.64) 1

In Sect. 5.13 it was shown that the first term in brackets in Eq. (5.64) is the
homogeneous strain that cause a linear variation of d with sin 2tp. The second and
third terms are inhomogeneous strain terms which cause the non-linearity in d vs.
sin 2tp. The second term is finite only when a? 1 ,Ki ( tp) are finite and describes the
average reaction strains to an applied or macro-stress in the population of grains
diffracting at the particular tp-tilt. The third term is the inhomogeneous residual
strain term in this population. This term may be finite, for example, if plastic
deformation is inhomogeneously distributed in the grains within the irradiated
volume such that various grains mutually constrain each other. The contribution
of all these terms must be taken into account for a general solution of the
strain/stress state in the irradiated layers.
Unfortunately, determination of the macrostress, a? 1, and the average
microstrains, <Eij>~' from Eq. (5.64) is not easy. If the interaction constants
Ki (tp) at each tp-tilt, as well as the functional dependence of <Ei)~ on tp are
known, d vs. sin 2tp plots may be analyzed by curve-fitting, or non-linear leastsquares analysis. However, in the general case, the equations describing
Ki ( tp ) ,< Eij>~ are not simple functions of tp and require the definition of a large
number of parameters 2 which can cause large errors in the results.
1 Here it is assumed that q, =0 and that only the applied and/or the macroresidual stresses cr? 1
are finite in the irradiated volume.
2 The interaction constants K j (II') are dependent on the texture function (which is usually
expressed as a spherical function [32]) and on the distribution of grain shapes in the given
population. The inhomogeneous residual strain distribution also depends on the texture
function (which determines the orientations of the grains) and on the type and direction of
loading during deformation (which determines the total load and consequently the plastic
deformation in grains of a definite orientation).

154

5 Determination of Strain and Stress Fields by Diffraction Methods

In special cases, analysis of oscillatory d vs. Sin2'l' may be easier. For example,
if the material under investigation has a large grain size such that a parallel-beam
geometry coupled with a PSD can be used to obtain the strain distributions in
discrete grains (similar to the apparatus described in Sect. 5.9) one can map the
average strains in discrete grains by treating each grain as a single crystal.
Similarly, if it is known, a-priori, that either of the components causing the
oscillations [terms II or III in Eq. (5.64)] is negligible, Eq. (5.64) may be
simplified accordingly.
Various other approaches have also been suggested to date for the treatment of
oscillatory d vs. sin2'l' data. These procedures are usually laborious and are based
on restrictive assumptions, however, and are of limited practical value. These
approaches are reviewed in Appendix B.

5.15 Stress Measurements with Neutron Diffraction


Residual stress measurement with neutron diffraction is based on the same general
principles as x-ray methods. Equation (5.4) is the fundamental equation in this
case as well and predicts regular d vs. sin2'l' plots when the strain distribution in the
volume probed by neutrons is homogeneous. Similarly, oscillations indicate an
inhomogeneous microstrain field. The differences between x-ray and neutron
methods lie mainly in the position and size of the diffracting volume in the
specimen. It was shown in Sect. 4.26 that because of the much lower absorption of
neutrons in most engineering materials, the distance in the material that a neutron
beam can traverse is longer (by a factor of 103 ) than that ofx-rays. For example,
a neutron beam is attenuated to half its incident intensity by 6 rom of steel or 70
mm of aluminium and to one tenth of its incident intensity by 20 mm of steel. Since
both the incident and the diffracted beams are attenuated, the effective depth that
can be stuQied is usually half of the material thickness that permits sufficient
diffracted intellsity for accurate lattice parameter determination. One other
important parameter than conrols the diffracted intensity is the diffraction
volume. Since, as discussed in Sect. 4.26, the scattering factors of neutrons are
much smaller than x-rays, and the incident intensities at the sample are low,
diffraction volumes of 10-30 mm3, which are larger than those of x-rays by a
factor of 2 or '3, are required to obtain sufficient diffracted intensities. The
strains/stresses measured are averages over this volume.
The dimensions and the shape of the diffracting volume must be selected such
that the stress/strain along any direction within this diffracting volume is
approximately homogeneous.
The shape a.nd dimensions of the diffracting volume, as well as its position
within the material is defined by the slits placed on the incident and diffracted
beams (Fig. 5.18 ). Since the position of the diffracting volume within the
specimen can be changed by translating the specimen over the center of the
diffractometer, it is possible to measure the residual stress profile with depth
without layer removal [33, 34]. The resolution along this direction is determined
by the length of the intersection of the specimen coordinate, Si' with the cross-

155

5.15 Stress Measurements with Neutron Diffraction


Neutron Source
.-----

Diffractometer
/ ' Circle

Fig. 5.IS. Definition ofthe position and shape ofthe probe volume in a neutron experiment by the
divergent and receiving slit/soller collimator combination (a), experiment geometry (b)
Incident Beam

Diffracted Beam

Fig. 5.19 a-c. Spatial resolution obtained by various divergent and receiving slit combinations.
The dimensions of the probe region must be chosen such that there is negligible variation of
residual stress within the probe volume

section of the diffracting volume. By using various shapes of slits, the size and
shape of this volume may be changed (Fig. 5.19) to match the characteristics of
the strain distribution. For example, in Fig. 5.19b,c, the long dimensions of the
volume would be along the specimen direction where strains are relatively
constant with length. The height of the diffracting volume perpendicular to the
diffractometer plane is defined by the height of the slits as shown in Fig. 5.19.
The cross-section of the irradiated volume, on the other hand, is a rhombus whose
dimensions depend on the slit width, W, and diffraction angle, 29. The intersection
length of the rhombus along the specimen coordinates Sl,S2 also depend on '1'.
These relationships are summarized in Fig. 5.20. An interesting feature of the
neutron stress measurement is that in contrast to the x-ray methods, where 'I'
between 0 and 60 can be used for most experiments, any 'I' between 0 and 90
may be used as long as the path length within the material is not longer than the
critical absorption length that will decrease diffracted intensity to unacceptable
levels. It must also be noted that since neutrons penetrate deep into the material,

5 Determination of Strain and Stress Fields by Diffraction Methods

156

.,,=0

Xmin
Xmax

W = SLIT WIDTH
K = W/sln 28
Xmax =W/sln8

Xmax = W/cos 8

Xmln = W/cos 8

Xmln =.W/sln 8

L~." = Xmln

L~." =X max
Definition of
Specimen axes

28

. d,

, ,
-'_

'L '

'cfAIt
I
I

~ KI
L~." = Xmln cos."

Fig. S.20. Probe region geometry


for 11'=0 and 11'=11' for a flat
sample

all terms of the stress tensor, including (113,(123,(133' may posses macro and micro
components. Thus, the biaxial methods discussed in the x-ray measurement
should be avoided unless there is independent evidence (for example, from
elasticity theory) that for the particular specimen, these terms are zero. Similarly,
in materials subjected to surface treatments, such as shot-peening, grinding,
hardening, etc., where the variation of residual stress in the surface layers is steep,
the neutron beam should not border on the surface. In such cases, as discussed in
the previous sections, both the macro and micro stresses will be a function of
depth.
One other important consideration is the homogeneity of deformation in the
irradiated volume. As discussed above, this may be obtained from the behavior of
d vs. sin 2 tp. If, because of low diffracted intensities and long counting times,
measurement of d for a large number of tp-tilts is impractical, independent
confirmation from elasticity theory, finite element analysis, or x-ray measurements must be used as a check. The best method, however, is to use a position

5.16 Effect of Composition Gradients with Depth

157

sensitive detector for neutrons [35], which considerably shortens measurement


time, and to measure as many points as feasible.
Another method utilized in neutron strain measurements is analysis of the
peak positions for various reflections obtained from time-of-flight measurements,
discussed in Sect. 4.26. This method yields lattice spacings, and thus strains for the
appropriate reflections. In this case, however, there are no slits and the entire
volume of the specimen intersecting the incident beam diffracts. Thus, macrostresses that are balanced between the surface layers and the interior of the
material average to zero for all reflections, since

Jz Jjj) macrodz = 0 ,

(3.5)

and the strains observed will be due to microstresses that are balanced between
various grains. These values will also be averages over the entire specimen volume.
Thus, materials that exhibit a deformation gradient from the surface into the bulk
should not be studied by this method. This method, on the other hand, has been
used successfully to analyze the average microstrains in the individual phases of
multi-phase materials where deformation is homogeneous through the sample
thickness [36]. Iftpeaverage strain in a given phase is inhomogeneous, however,
the results from this method are, as for all other methods discussed so far,
extremely hard to interpret.
Based on the discussion given above, one may conclude that x-ray and neutron
diffraction techniques of residual stress determination are complementary. By
using both these methods, which are sensitive to strain fields over very different
characteristic measurement lengths, it is possible to obtain detailed strain
information of the stress field in a given specimen. This has not yet been done.

5.16 Effect of Composition Gradients with Depth


The variation of stress with depth, or the presence of a residual stress field, are not
the only possible causesofthe variation of the plane spacing, dcjltp' with sin 2 tp. In
certain materials, such an effect may be due to the presence of a concentration
gradient. In such cases, a correction must be applied to the residual stress values
obtained from the methods described so far.
In alloys and solid solutions the equilibrium (unstressed) lattice parameter is
dependent on the particular composition and may change when the concentration
of any solute within the system is changed. The amount of change is dependent on
the relative atomic sizes, and the valency and electronegativity of the components
of the system in question. (The reader is referred to the texts by Pearson [37] and
by Barrett and Massalski [38] for a detailed treatment.) This variation of lattice
parameter with composition is usually linear for low solute concentrations. For
example, the size effect is described by Vegard's law [37]:
(5.65)

158

5 Determination of Strain and Stress Fields by Diffraction Methods

where (ao) A> (ao ) B are the equilibrium lattice parameters of the solvent and
solute respectively, (a) S.s is the lattice parameter of the solid solution and fB is the
atomic fraction of the solute. Similar relationships apply for the other effects.
Thus, in most cases, it is possible to obtain empirical relationships, describing the
variation of the plane spacing, do, of any set of planes with composition of the
form:
(5.66 )
Here A,B are constants and nx is the solute concentration. In the case of austenite
and martensite, the dependence of the lattice parameters on the carbon
composition, nc, is given by [39]
.
c=2.861 + .116nc
a=2.861-0.013nc
for the body-centered tetragonal martensite structure and

a = 3.548 +O.044nc
for the FCC austenite structure.
Now consider the case where residual stresses are to be measured in the
austenite phase of a steel specimen in which the carbon concentration changes
from the surface to the interior (e.g., a caburized specimen). Assume, for
simplicity, that the residual stress tensor in the surface is biaxial. In this case, the
dependence of the measured plane spacing, d~, on the residual stresses is given by
Eq. (5.17):

d~",-do
. 2
(1-V) (all +( )
do = (1+V)
E
a~sm 'V+ E
22

(5.17 )

Here do is the unstressed plane spacing and the other terms have been previously
defmed. In this case, however, if the carbon concentration exhibits significant
changes over the maximum depth penetrated by the x-rays, the average
equilibrium plane spacing, do, measured at any 'V-tilt will be given by an equation
similar to that given for strains in Eq. (5.42):

J (A+Bnc(Z) )e-z/T'Pdz

Top

<do)~=

Tip

J e-z/T'Pdz

( 5.67)

Here nc (z) is the variation of carbon concentration with depth (z) and the other
terms are defined in Eq. (5.42). Thus, Eq. (5.17) becomes (by substituting
Eq. (5.67) and integrating)

(d)~",=Kl +K2(nc('t) )",. { (1 ~v) a~sin2'V+ (1 ;v) (all +(22) + 1} ,


(5.68 )
where K 1 ,K2 are constants and (nc ('t)", is the x-ray average of the concentration
gradient at depth 't given by Eq. (5.67) for a particular 'V-tilt.

5.17 X-ray Determination of Yielding

159

It can be seen from this equation that, even when all the stress terms aij are
zero, d~ will change with tp-tilt creating a fictious stress. Thus, residual stress
values determined from such specimens must be corrected for the variation of
concentration with depth. This is possible if the actual concentration gradients
with depth and the functional dependence of equilibrium lattice parameters on
solute concentrations are known. In such a case, the terms K 1,K2 and <nc( t) >1p
can be determined from Eq. (5.67) after which Eq. (5.68) can be solved for the
correct stress values. A similar procedure can also be used for the triaxial formulas
[Eqs. (5.22) - ( 5.26). In this case, the procedure described automatically
corrects for the variation of the unstressed lattice parameter with depth. (On the
other hand, one must make sure that the total error introduced into the calculation
from the composition terms is within acceptable limits. This can be done by
straightforward propogation of error, similar to the procedures discussed in
Chap. 6.)
It must be noted that the effect described above will be pronounced only for
those cases where concentration changes appreciably over the penetration depth
for the radiation employed. This correction may also be especially important in
stress determination with neutron diffraction because of the much larger volumes
involved. Similarly, in the case of surface doped or irradiated thin films, or
otherwise surface treated materials with sharp concentration gradients, it may be a
consideration with x-ray diffraction. The use of highly absorbed, long-wavelength
radiations will minimize this effect, similar to the case discussed for the stress
gradients.

5.17 X -ray Determination of Yielding


The various measures of yielding, such as the true yield point, the elastic limit, etc.,
discussed in Sect. 2.14 are usually determined by standard extensometers or with
LVTD (linear variable differential transformer) gages. In both cases, the values
determined are averages over the total gage length. Diffraction methods can also
be used to determine the onset of yielding. This can be accomplished, as discussed
in the case of the x-ray elastic constant determination, by loading a material insitu on the diffract9meter and measuring the spacing d of a particular family of
planes. As long as the deformation in the diffracting volume is elastic, the plane
spacing d will change in proportion to the applied load (Fig. 5.21 ). However, once
dislocation movement starts in the diffracting grains, the plane spacing will remain
constant since plastic extension (or compression) of the distance between atomic
planes is not' possible. Thus, in this case the proportional limit will be equal to the
x-ray yield point.
Measurements such as this can be used in a number of ways: if one measures
the yield point of various sets of planes; hlklll' h2k212' h3k3l3' etc., one obtains a
measure of the distribution of the "x-ray yield point" in the irradiated volume.
Such a measurement will be particularly easy at a synchrotoron, where the
wavelength can be varied easily and the appropriate plane spacings measured at
high Bragg angles for accuracy.

5 Determination of Strain and Stress Fields by Diffraction Methods

160

0'1

'u

Mechanical stress

a.

III
Q)

0..

a:::

Fig. 5.21. Variation of interplanar spacing d in a residual-stress free specimen under


applied stress. The "x-ray elastic limit (s) (for tension and compression)" are
defined as the stresses at which the plot becomes horizontal

Another interesting measurement of this sort is the measurement of the elastic


limit of different regions in a given sample. For example, with x-rays one can
determine the onset of yielding in the surface regions [40J, or with neutrons, onset
of yielding in the interior.
It must be noted that the values obtained from such experiments will be x-ray
averages of the actual distribution within the diffracting volume. Thus, their
interpretation may not always be straightforward and one may have to use
concepts similar to those developed in Sect. 5.13 for the x-ray elastic constant
determination for the correct analysis of data.
In those cases, however, where the composition of the material under
investigation changes appreciably within the volume irradiated by neutrons or xrays, errors may be introduced into the stress analysis because of this variation. In
such cases, corrections must be applied to the "d" vs. sin 2 \j.l data before the stress
analysis is performed.
In addition to stress analysis, it is also possible to use diffraction techniques to
determine the onset of plastic deformation in various subsets of the specimen
under investigation.

5.18 Summary
From the discussions given in this chapter it may be seen that all diffraction
methods of measuring stress measure strains along various directions in the
specimen and then use these values to calculate the stresses along any direction
through tensor transformation formulae and the appropriate form of Hooke's
law. While this basic methodology permits the determination of general triaxial

5.18 Summary

161

stress/strain fields, traditionally modified forms, such as single-exposure, two-tilt


or sin 2 ", methods, have been used for the determination of biaxial stresses in the
near-surface layers of materiais. In some situations (such as ground two-phase
alloys for example) these approaches are not adequate and the general solutions
must be used. Two properties of measured data may be used to check the
dimensionality of the strain tensor:
1) Presence of ",-splitting in "d" vs. sin 2 ", (d~+ =Fd~- ), which indicates thatthe
shear strains E 13 ,E 23 are finite within the irradiated volume.
2) Presence of curvature in "d" vs. sin 2 "" which indicates strain gradients with
depth.
When these effects are present, the properties of the material under examination
(isotropic or anisotropic) and of the deformation suffered by the specimen must
be considered in the selection of a given method of strain/sress analysis.
The diffraction methods are valid for isotropic and anisotropic materials.
However they require a homogeneous strain distribution within the irradiated
volume. Experimenlally determined "regular" "d" vs. sin 2 ", plots indicate that the
strain distribution within the irradiated volume is indeed homogeneous. Oscillatory "d" vs. sin 2 ", data, on the other hand, indicate the presence of an
inhomogeneous strain/stress distribution within the irradiated volume for the
particular lattice plane hkl used in the measurement. Analysis of such data with the
"regular" procedures may cause large errors in the results.
The strains/stresses determined by diffraction methods may contain macro
and micro stress components. The macro components are due to the mutual
constraint of macroscopic volumes of material or to loads applied at the
boundary of the material. The micro components arise from grain interactions in
response to macro stresses (applied or residual) or from inhomogeneous
partitioning of plastic deformation. The macro and average micro components
may be separated by x-ray diffraction methods if the total strains from all phases
of a multiphase solid can be measured, as well as the volume fractions of these
phases.
,
The elastic constants coupling strains determined by x-rays to the stresses
existing in the specimen volume may be measured by x-rays or calculated from
elasticity theory. If the distribution of crystalline directions within the material
volume is completely random, theoretical calculations show good agreement with
experimental values. 'In the presence of non-random distributions (texture), deviations occur. These deviations are due to interaction components which contribute to the effective elastic constants of the material. These effective elastic
constants can be measure by x-rays even when the "d" vs. sin 2 ", is oscillatory.
These parameters contain configurational terms as well as "real" elastic constants
like Young's modulus and Poisson's ratio.
X-ray diffraction techniques cannot be used to measure strain fields deep
within the material ( > 100 microns) without electropolishing surface layers. In
those cases where such measurements are desirable, one may use neutron
diffraction. Because of their small absorption coefficients, neutrons penetrate
much deeper into the material and thus it is possible to obtain strain data without
layer removal.

162

5 Determination of Strain and Stress Fields by Diffraction Methods

Problem
5.1. Given the following "d" vs. sin21V data,
sin21V

d,A

sin21V

d,A

0.00
0.10
0.20
0.30

2.87206
2.87082
2.86965
2.86856

0.40
0.50
0.60
0.75

2.86753
2.86656
2.86563
2.86445

a) It is also given that dcp.p = d,p for all <1>. Plot "d" vs. sin21V, what can you conclude
from the shape of this plot? Describe the possible stress tensors. What can you
conclude about the strain distribution in the surface of the material?
b) Given tha~ the material is shot-peened steel

;v = 5.77 .10- 6 MPa,

~ = 1.25 .10

MPa)

Determine the surface stress cr", using the biaxial sin21V method and the two
tilt method.
c) Using the triaxial analysis determine the stresses tensor over three IV-ranges,
sin21V=0-O.75, sin21V=O-O.3, sin21V=0.4-0.75. Assume d o =2.8665A.
d) Calculate the unstressed lattice spacing ( do) 'P* using the approximate formula
given in the text for the three IV-ranges. What is the error in do at each tilt?
e) Re-do (c) using do = ( do ) 'P*. Compare your answers with the results from
( c ); what are the errors.

References
1
2
3
4

H. Dolle, J. Appl. Cryst., 12,489 (1979)


H. Dolle and V. Hauk, Z.f. Metalkde, 68, 728 (1977)
H.H. Lester a,nd R.M. Abom, Army Ordnance 6, 120, 200, 283, 364 (1925 -1926)
Soc. Automotive Eng., Residual Stress Measurement by X-ray Diffraction, SAE J784a, 2nd
ed. (1971)
5 H. Dolle and J.B. Cohen, Met. Trans. A, ll-A, 831 (1980)
6 V. Hauk, P.J.T. Stuitje, and G. Vaessen, Hiirterei-Tech. Mitt., Beiheft, 129 (1982)
7 I.e. Noyan, Met. Trans. A, 14A, 1907 (1983)
8 e.N.J. Wagner, Acta. Met., 5 427 (1957)
9 H.M. OUe, J.,Appl. Phys., 33,1436 (1962)
10 R.L. Rothman and J.B. Cohen, J. Appl. Phys., 42, 971 (1971)
11 V.M. Hauk, R.W.M. Oudelhoven, and G.H.J. Vaessen, Met. Trans. A, 13A, 1239 (1982)
12 I.C. Noyan, Adv. X-ray Anal. (in print)
13 T. Imura, S. Weissman, and J.J. Slade, Jr., Acta Cryst., 786 (1962)
14 S. Weissman and W.E. Mayo, ASM Metals/Matis. Tech. Series, 8311-006, 185 (1984)
15 W.E. Mayo, J. Chaudhuri, and S. Weissman, ASM Metals/Matis. Tech. Series, 8311-005,
129 (1984)

5.18 Summary
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39

163

A. Segmiiller and M. Murakami, in "Analytical Techniques for Thin Films", K.N. Tu and R.
Rosenberg Eds., Tratises on Mat. Sci. and Technology, Academic Press, New York (in
print)
M. Murakami, CRC Critical Reviews in Mat. Sci., 11, 317 (1983)
V.S. Speriosu and T. Vreeland, Jr., J. Appl. Phys., 56, 1591 (1984)
V.S. Speriosu, J. Appl. Phys., 52 6094 (1981)
A. Segmiiller, P. Krishna, and L. Esaki, J. Appl. Cryst., 10, 1 (1977)
A. Segmiiller, J. Angiello, and S.J. La Placa, J. Appl. Phys., 51, 6224 (1980)
U. Wolfstieg, Harterei-Tech. Mitt., 31 83 (1976)
I.e. Noyan, Mat. Sci and Eng. (in print)
e.M. van Baal, Laboratory for Metallurgy Report, Delft Uni. of Tech., Rotterdamsurg,
Netherlands (1982)
e.M. Brakman, J. Appl. Cryst., 16, 325 (1983)
V.M. Hauk, Adv. in X-ray Anal., 27, 101 (1983)
I.e. Noyan, Ph. D. Thesis, Northwestern University, Tech. Institute, Evanston II. 60201
(1984)
I.C. Noyan and J.B. Cohen, Adv. in X-ray Anal., 27, 129 (1983)
N.R. Draper and H. Smith, "Applied Regression Analysis", Wiley-Interscience, New York,
NNY (1966)
R. Marion, Ph. D. Thesis, Northwestern University, Tech. Institute, Evanston. IL 60201
(1972)
e.M. Sayers, Publication MPDjNBS/233, Materials Physics Division, Harwell Oxfordshire, England.( 1983)
L. Pintschovius, V. Jung, E. Macherauch, and O. Vohringer, Mat. Sci. and Eng., 61, 43
(1983 )
A. Krawitz, J.E. Brune, and M.J. Schmank in "Residual Stress and Stress Relaxation",
E. Kula, V. Weiss Eds., Plenum Press, New York, NY, 139 (1981)
C.W. Tompson, D.F.R. Mildner, M. Mehregany, R. Berliner, and W.B. Yelon, J. Appl.
Cryst. 17, 385 (1984)
A.D. Krawitz, R. Roberts, and J. Faber, Adv. in X-ray Anal., 27, 239 (1983)
W.B. Pearson, "A Handbook of Lattice Spacings and Structures of Metals and Alloys",
Pergamon Press, New York, 19-54 (1958)
C.S. Barrett and T.B. Massalski, "Structure of Metals", 3rd ed., McGraw-Hili, New York,
357-379 (1966)
C.S. Roberts, Trans. AIME, 191, 203 (1953)
S. Kodama, Proc. Int. Congo Mech. Behav. of Materials, Soc. Mat. Sci. Japan, Kyoto, 111
( 1972)

6 Experimental Errors Associated with the X-ray


Measurement of Residual Stress

6.1 Introduction
The basic methods of residual stress measurements with x-rays and neutrons, as
well as the theory underlying the measurements, have been covered in the previous
chapters. In any scientific experiment, however, the error associated with the
measured quantity is just as important as the measured value itself and must be
known for the correct interpretation of the results. Evaluation of errors by
theoretical foanulae is also of primary importance in automating a measurement
such that the error in the measurement can be specified a priori by the operator.
In this chapter, the errors associated with the residual stress measurement will
be discussed. The discussion will be in two parts. In the first part, the selection of
the diffraction peak, methods of peak location and the statistiCal errors associated
with the maximum 28 value (and therefore the plane spacing "dcjl1jl") obtained
from these methods will be reviewed. This error will then be used to obtain the
statistical error in stress determined from the analysis of "dcjl1jl" vs. sin 2 \j.l data
through error propogation techniques. In the second part the instrumental errors
arising from diffractometer misalignment, specimen curvature and various other
geometrical factors, as well as errors due to specimen preparation techniques such
as electropolishing, will be reviewed and various techniques of minimizing such
arrors will be discussed. It must be noted, however, that the equations describing
the errors in stress/strain throughout the chapter have been derived assuming
regular "d" vs. sin 2 \j.l behavior and are not valid for oscillatory data.

6.2 Selection of the Diffraction Peak for Stress Measurements


In Chap. 4 it was shown that when a polycrystalline sample is irradiated with xrays, diffracted intensities can be obtained from all sets of planes that:
(i) satisfy the Bragg condition,
(ii) have non-zero structure factors.
Thus, if a polycrystalline specimen is examined with a diffractometer employing
monochromatic radiation, various reflections can be brought into the diffraction
condition at the appropriate 28 values and may be used for stress measurement.
Generally the peak at the highest 28 is chosen since the peaks at high 28 values are

6.2 Selection of the Diffraction Peak for Stress Measurements

165

700.-------------------------------,

MPa

""2"8

-+-__

- j -_ _

~~=35

500

500

t
~

400
I
I

I
300

I
I

Ktwotilt

Fe211

Cr

9Fig. 6.1. Variation ofthe stress constant K as a function of diffraction angle 9 for single exposure
and two-tilt methods

more sensitive to strains/stresses, i.e., for a given stress, a high angle peak will
display a larger shift in 29. This may be seen by differentiating Bragg's law:
nA.=2dsin9
0= 2d cos9L19 + 2L1d sin 9
or:

L19= -stan 9,

(6.1 )

L1d
s=d
Since tan 9 is larger for larger 9, the term L19 is larger for a given stress. It may be
seen from Eq. (6.1) that for a given strain in an iron specimen examined with Cr
radiation, the peak shift of the 211 peak at 1560 29 will be 3.6 times the peak shift of
the 200 peak at 1050 29 and 6.9 times the peak shift of the 110 peak at 680 29.
Because of this effect peaks with 29 values above 1400 are preferred for routine
stress measurements. It may be noted that, the term tan 9, which causes this
behavior, also appears in the stress factors K for the single exposure and two tilt
methods [where 0" = KL19, Eqs. (5.19), (5.20)]. Thus, the stress factor also varies

166

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress

with 29. In Fig.6.1 the variation of K for steel specimens over the 29 range
frequently used for stress measurments in steels is shown. It may be seen from this
figure that the sensitivity of different methods to strains/stresses are different even
over the same angular range (the higher the term K, the poorer the sensitivity) .

6.3 Peak Location


The data obtained with a diffractometer at a given \P is either a strip-chart
recording or a digital output of the variation of the diffracted intensity over the
pertinent 29 range (Fig. 6.2). These raw intensities contain contributions from
angle-dependent intensity factors, such as the Lorentz-polarization factor,
absorption, etc. and the intensity variation caused by such factors must be
removed from the profile before the "true" peak maximum is determined (Sect.
4.24). Depending on the subsequent method of stress analysis, various approaches have been used to determine the 29 value corresponding to the maximum
intensity from the corrected data. If a biaxial stress determination method is used,
the stress is proportional to the slope of the "d" vs. sin2\p line which, in turn, is
directly proportional to the peak shift between \p = \Po and \p = \p. Thus, any
consistent feature of the line profile may be utilized to define a reference point
whose position is monitored as a function of the \p-tilt. This peak position
located by anyone method need not correspond to the one defined with any other
or to the actual peak position, as long as the reproducibility of the peak shift is
sufficient for the measurement. In triaxial stress measurements, on the other hand,
the analysis [Eqs. (5.4) - ( 5.15 ) ] is based on the strains Ecj>1jl' given by
EcI>1V=

dcj>1jl-d o
do
'

(5.1 )

where do is the ll;nstressed lattice spacing. Since the strains Eij are less than .01 % for
elastic behavior in most materials, the absolute value of the peak position, i.e.,

b
a
Fig. 6.2. Intensity vs. 29 profiles obtained from an (a) annealed and (b) cold-worked 1090 steel
specimens

6.3 Peak Location

167

the true peak maximum, is required for the accurate calculation of d..., from
Bragg's law and an appropriate method of peak location must be used [1].
In the following the various methods of peak location are reviewed. In all cases
it will be assumed that the appropriate angular corrections to the raw intensities
have already been performed.

6.3.1 Half-Value Breadth and Centroid Methods


In the half-value breadth method, the 29 value corresponding to the peak
maximum is determined by the mean of two angles:
xpeak=1/2(L1/2+X1/2)'

(6.2)

where x 112 is the intersection of the profile and the straight line parallel to
the background at half the maximum intensity after the background has been
subtracted (Fig. 6.3a) .
In the centroid method, the center of gravity or centroid of the peak is defmed
as [2]
J29 I ( 29 ) d29
(6.3)
(29)peak = II (29) .d29
The centroid of a given profile may be calculated by a procedure proposed by
Baucum and Ammons [3]. This procedure is based on Simpson's rule for the area
under the curve. Consider the profile given in Fig. 6.3b. The centroid of this profile
is given by
k

LIDj

;=1

(6.4 )

X=-k-'
LA;
;=1

where

X;-h...J

x;

LX;+h

x-

a
b
Fig. 6.3. Calculation of the (a) half-value breadth and (b) centroid of a diffraction peak

168

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress

and

Here h represents the increment between data points and


k=

=11=

of data points - 3 + 1
2

(6.5 )

since three data points are required for the first incrment and two are required for
the rest.
6.3.2 Functional Representations of X-ray Peaks
The methods in this group represent the x-ray peak as mathematical functions
which can approximate the various peak shapes encountered in practice with
simple modifications. These functions are expressed in terms of various parameters describing the peak shape and location. These parameters (which also
include the peak maximum) are then determined by matching the function to the
observed peak profile. Among the functions used for this purpose are the
modified Lorentzian [4, 5], Pearson type VIII [4, 6], Gaussian and Cauchy [4,
7] profiles.
The modified Lorentzian is one of the more popular functional forms used in
profile fitting [4, 5] and its applicability to peak location determination has been
investigated by Devine [4]. The function is of the form
1(28);=A { COS1t

( 28i - 28p a

())

}n . K2+ (28.-28
K2
)2 +SL28 +INT.
i

(6.6)

Here, I (28i ) is the calculated intensity profile, A,a,8 are parameters related to
the shape of the line profile (where' A' is related to the maximum intensity at the
peak, 'a' is related to the full-width of the profile from background to background
and '8' is used to account for small asymmetries in peak shape) , K is a constant
Lorentzian term and 28p ,28i are the two-theta values at the peak and at position i
along the profile respectively. The cosine term raised to the appropriate power 'n'
is used to force the curve to drop more rapidly in the tail regions of the line profile.
The terms SL and INT fit the tails of the peak profile to the background. If the
background is approximated by a straight line under the peak, SL is the slope of
this line and INT is the intercept respectively. This function generates profiles that
fit well to most observed x-ray line profiles (Fig. 6.4 ).
The Pearso~ Type VII function [4, 6] is of the form
1(28).=1 . {1 + (28i -28p )2} -m + SL28+INT ,
I
0
ma 2
I

(6.7)

where 10 , and a are constants optimized in the profile fitting process, the power
term 'm' is user-defined according to the sharpness of the peak, and the other
terms have been defined in Eq. (6.6).

169

6.3 Peak Location


1i.00
cps

.. .

1200

:c
'iii

Observed
Calculated

1000

:s
<lJ

800

600
i.

.
0067.6 0

00

..
..'.

.. .

'

... '

'

'

68,0 0

68.i.
68.8
26-

69.2

69.6

Fig. 6.4. Modified Loientzian function fitted to the 110 ferrite peak obtained from a cold-worked
1008 steel sample. (From Devine [4] )

Equation (6.7) takes the form of the Cauchy function for m = 1:


1(29 L =l o.{1+

(29j~229p)2} -1 + SL'29j+ INT

(6.8)

and for m--+oo, it becomes the Gaussian function


(6.9)

This can be seen by expanding Eq, (6.7) in the form of a binomial series:
1(29d = 1 - m + ( - m)( -1) - m -

1.

X+ m ( - ; - 1 ) ( -1) - m - 2X 2 + ... ,

j
p
d ' fi'
h'
. b
h
were
x = ( 29 - 29
2) 2' A
S m ten s to In Imty, t e IntensIty ecomes
rna

1(29.) =1 '{1
I

+.

- (29 j-29p)2 - (29 j-29p)4 0


+
2'.a4
+
a2

This is equivalent to the exponential expansion


y2 'y3
eY = 1 + y + 2! + 3! + ... ,
where y= - (29j-29p)2/a2. Thus:
1(29d =Io.exp { _

(29j~229p)2}

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress

170

...;.ii ..,

1200
cps

Pearson type-m

900

I'

I'

,.

'

I'

:c- 500

......

'\Ii

c(\)

C
....

A; q~served
D:

.:

300

O.

aO., am:1

DD

DD

...

DO

"AOD

m=2

-3005~7~5~0--~5~8.-0~0----5~8-.4~0-----5L8.~8~--~5~9.-2~---5-"9.50
28Fig. 6.5. Pearson':type VII functions with m = 1,2, and 3 fitted to the 110 ferrite peak shown in
Fig. 6.4. (From Devine [4])

500r----------------------------------,

400

>.

300

~
c(\)

:s 200
100

-.-

I
r
/

,,---

.......

53.00
53.50
29-

54.00

54.5 0

Fig. 6.6. A neutron diffraction peak (AI, 311) fitted


with a Gaussian peak.
(From Rudnik, unpublished
research)

The functions based on the Pearson type VII function are easier to use than the
modified Lorenzian because of the smaller number of unknown parameters. These
functions, however, do not exhibit as good a fit as the modified LQrenzian
function (Fig. 6.5 ) .
Various mathematical procedures may be used in fitting the functions to the
observed proflle. In the case of the simpler functions [Eqs. (6.7) - (6.9)] the
solutions may be found using either first or second order differential techniques
such as nonlinear least-squares procedures. In the case of the modified Lorentzian
function more complicated procedures must be used because of the larger number

6.3 Peak Location

171

of unknown parameters involved [seven in the case ofEq. (6.6)]. Devine [4] used
a modified "simplex" method, capable of handling up to twelve parameters, in this
case. Similar routines are also available in various program libraries. In the case of
neutron diffraction, the peaks are Gaussian in shape [8] (Fig. 6.6) and other
functional forms are not used.
6.3.3 Peak Determination by Fitting a Parabola
This method, which was first used by Ogilvie [9], is one of the most popular
methods today. In this method, it is assumed that a region in the vicinity of the
peak can be represented by a parabola. Expanding the Cauchy [Eq. 6.8 )] or the
Gaussian [Eq. (6.9)] functions as power series,
I ( 29i ) = 10 - 210 (2 9i - 29p) 2 + ...
a

(6.12 )

for the Gaussian case and similarly for the Cauchy equation.
If the third and higher order terms are neglected, it is seen that both equations
assume the form
(6.13 )

where C is a constant. Equation (6.13) is the equation for a parabola whose axis is
parallel to the intensity ordinate. Therefore, the peak position can be determined
by measuring a number of points (intensity vs. 29) along the peak profile, fitting a
parabolic equation to the measured data and determining the apex of this
parabola. Two factors are critical in this procedure: the region of parabolic fit and
the number of points used in fitting the parabola.
The region of the diffraction profile which can be fitted by a parabola is not
well defined. Thomsen and Yap [10] have shown that, for x-ray profiles
satisfactorily represented by the modified Lorenzian function [Eq. (6.6)], a
parabolic assumption is valid when the parameter V, given by the function
(6.14 )

is less than 0.32. Here W is the full width at half the maximum height of the
diffraction profile, Z9p is the peak position, 29min is the minimum 29 that is within
the parabolic region and it is assumed that the peak is symmetric around 29.
An empirical rule proposed by Koistinen and Marburger [11] indieates that a
parabola should be fitted over the region of the peak which is bounded by the
points that have 85 % of the maximum peak intensity (Fig. 6.7). This so called
"top 15 pct." rule has found wide acceptance in practice since, in contrast to
Eq. (6.14), the. total peak profile need not be determined accurately. Furthermore, both methods yield similar results for most cases as shown in Table 6.1
[12].
The number of points measured in the appropriate 29 range and used in
calculating the parabola describing this region affects the reproducibility of the
apex of the parabola, i.e., the calculated peak position. Originally Kostinen and

172

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress


I

1200. en

Q.

900.-

..

>-

I-

en
~

600.1-

IZ
H

,.

.. ..

Observed

Calculated

'

..

300.1-

68.8
69.2
28_
Fig. 6.7. A 7-point parabola fitted to the top 15 % of the peak shown in Fig. 6.4 and 6.5. (From
Devine [4])
68.4

68.8

Table 6.1, Region of parabolic fit for the 211 Diffraction peak from various criteria. From James
& Cohen [12]
Steel
specimen

FWHM* Peak/back- peak


(29)
ground ratio ( 29)

29 min
Eq. (6.15)
(29)

29 min
15% rule
CZ9)

29 min
15% rule
after background
subtraction (29)

1090-1
1045-1
1045-3
1045-2
1090-2
TBAG-5

0.45
1.5
3.4
5.1
6.0
5.8

156.07
155.65
154.99
154.68
154.50
154.33

156.05
155.61
154.76
153.20
152.67
152.69

156.05
155.65
154.93
154.40
154.16
154.29

60
3.5
2.0
2.6

1.3
1.3

156.14
155.89
155.63
155.50
155.46
155.29

* FWHM = full width at half-maximum intensity

Marburger [11] suggested the fitting of a three-point parabola to the top 15 pet. of
the peak. Further work by Marion and Cohen [13] showed that reproducibility
was dramatically improved by the use of multiple points. This was later confirmed
by James and Cohen [12,14], who showed that given a fixed time for data
accumulation, use of multiple data points (five or more points, typically seven
points with background subtraction is used) could improve the reproducibility by
a factor of three.
The formulae describing the parabola fitted to a given angular range may be
derived in the following manner [14]. Assume that the data is obtained at an odd
number of observation points, spaced at equal increments of29. Thus, there are 'n'
points on each side of the center point (usually termed the working origin) and

6.3 Peak Location

173

the total number of points is 2n + 1. If the increment between the points is 3, the
parabola to be fitted to these points is defined by
a+b3j+c3 2j2=Ij

(6.15)

and the maximum of the parabola is obtained from

01.
J
-0
o( 3j) - .

(6.16 )

The parameters a,b,c in Eq. (6.15) are obtained from a least-squares best-fit to
the data by minimizing
(6.17 )
with respect to the unknown parameters a,b,c. Equation (6.17) may be solved by
setting the partial derivatives of S with respect to a,b,c to zero.
For example: .

as

oa =0=2 j~O (a+b3j+c3 2j2-Ij)


=a

where ni =

j=-o

j=-o

j=-o

j=-o

L jO+b3 L j+c32 L j2_ L Ij=noa+n232c-3-1mo,


o

(6.18 )

L land Mi , the ilh moment of the observed line profile, is given by

j=-o

Mi =3i + 1

L {Ij

(6.19)

j= - 0

The set of equations obtained by evaluating the remaining derivatives are


noa+O+ n232c- 3 -1 Mo =0
0+n 23b+0-3 72 M1

=0

( 6.20)

n2a+0+n432c-3-3M2 =0.
Simultaneous solution of Eqs. (6.20) for a,b,c yields
a= (Mon432-M2n2 )/(nOn4 -n~)33
b=Mt/3 3n 2

(6.21 )

c= (n232Mo-M2nO)/( 35 (n~-nOn4) .
The apex of the parabola is [from Eq. (6.16)]
2ep = 2eo- b/2c ,

( 6.22)

where 2eo is the working origin. Thus, from Eqs. (6.22) and (6.21), the peak
position (that is, the apex of the parabola) is given by
(6.23 )

174

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress


Initialize
angular

positions

step and
accummulate

move to
maximum
intensity
position

set 2,\ and

wgff~rl\

no

step scan for


the first
parabolic
fit

intensity
corrections
and curve
fit
calculate new

stePa~aan

step increment
and preset
counts

accummulate
data

Fig. 6.8. A typical flowchart for fitting a parabola to the top 15 % of a diffraction peak. (From
James [14])

6.3 Peak Location

175

The parabolic fit lends itself well to computer controlled data acquisition and
various software packages are available for such a fit. A representative flow chart
is shown in Fig. 6.8.
James and Cohen [12] have shown that subtracting of the average background from measured intensities before the parabolic fit is applied increases the
reproducibility of the peak position. The point where the background value is
determined is not critical in this case since it only affects the range over which the
parabola is being fitted and an approximate value for the 28, as well as for the
intensity, may be used.
6.3.4 Determination of Peak Shift
The methods discussed above determine either the true "peak" location or the
position of a reference point on the peak profile at any given 1.p-tilt. Peak shift
determination methods, on the other hand, determine the angular shift, A28, of
this point upon tilting. Significant among such methods are the "difference" and
"cross-correlation~ methods.
1) Difference Method
This method, originally developed for film methods [15], compares the variation
of intensity over the total profile for two 1.p-tilts and yields the angular shift, A28, in
the peak-maximum between these tilts. The difference function D (A) is defined as
[16]
D (A) = J[10 (28) - J'l' (28-A) ]d28,

(6.24 )

where A is the step size, 1,I'P are the intensities at 28 for 1.p = O,1.p respectively. The
peak-shift A28 is the value for which D(A) is a minimum [aD(A)/aA=O].
If a position-sensitive detector or a step-scanning counting method has been
used in the data acquisition such that, instead of a continuous profile, a discrete set
of points are available, Eq. (6.24) becomes [16]

Dr =

L (I

j;

j -

Ii"- J ' j = - (n -1 ), - (n - 2 ) , ... , ( n -1 ) .

(6.25 )

The peak-shift in this case can be determined from


1

,:l28''I'=C { k+

L Dj(j-k) }
j;k

'

(6.26 )

j~k Dj
where C is the increment in 28.
The difference method can also be used in a multiple 1.p-tilt procedure by
referencing the profiles obtained at all tilts to the profile of the initial1.p-position,
1.p0. If the peak position at the initial1.p-position is known, this method may also be
used to determine the absolute values of 28 at all other 1.p-tilts
(28~=28~ +A28''I').

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress

176

2) Cross-correlation Method
The cross-correlation method also yields the peak shift .12<]> of the intensity
maximum between two tp-tilts. In this case a correlation function F'" (.129) is
defined such that [16, 17, 18]
F'" (.129)

00

J 1 ( 29 ) .I'" ( 29 -

(6.27 )

.129 ) d ( 29) ,

-00

where 1, J'l' are the (corrected) intensities at tp = 0 and tp = tp and .129 is the step
size.

/:,.-

(step size)
28-Fig. 6.9. Shift of the x-ray diffraction profile upon tilting in a stressed sample and the maximum in
the cross-correlation function P'

0.5 . . . - - - - - - - - - - - - - . ,
en

g-

0.4

~ 0.3
en

~ 0.2

I-

0.1
0.0

l...-_ _L -_ _-'---_ _......l...-_ _.-J

0.0

128.0

256.0

384.0

512.0

CHANNEL No.

0.5 . . . - - - - - - - - - - - - - . ,
en 0.4
Co
(,)

>- 0.3

I-

en 0.2
z

Fig. 6.10. Intensity vs. 20 profiles from a


PSD measurement for IjJ angles 0 and 22.5
from s steel specimen. (From Knuuttila
[16])

I-

z 0.1
H

0.0
0.0

128.0

256.0

384.0

CHANNEL No.

512.0

6.3 Peak Location

177

20.0.----------------.,

16.0

UJ

12.0

Q.

*
UJ

Q.

8.0

4.0

0.0
-500.0 -300.0

-100.0

100.0

300.0

500.0

Fig. 6.11. Cross-correlation function obtained from the diffraction


profiles shown in Fig.6.10. (From
Knuuttila [16])

CHANNEL INDEX

The function P exhibits a sharp maximum when the unknown peak shift
is equal to ~2e (Fig. 6.9). Furthermore, when the peak profiles are
similar in height and breadth, F'P is symmetric about ~2eQ,'P [16].
If, instead of a continuous profile, the data is measured at n discrete points
spaced in equal intervals along the profile for each 'V-tilt, Eq. (6.27) is replaced by
a summation [16, 17]

~2eo.'P

Fj= L Ip-Jr- j

(6.28 )

i= 1

and the function Fj must be evaluated at all points along the line profile. Once the
Fj vs. ~2e profile is evaluated, the position of the maximum in Ft can be estimated
from the center of gravity [16]:
1

~2e'P=C {k+

,L Fj.(j-k)}
j=k

LP.
j=k .J

( 6.29)

In Fig. 6.11, Fj evaluated from the line profiles shown in Fig. 6.1 0 is shown. It can
be seen that the profile is quite symmetric and exhibits a sharp maximum even
though the intensity profiles are not smooth.
The procedure described above may also be used in a multi-tilt procedure. In
this case, a series of cross-correlation functions are obtained by referencing the
line profile at any 'V-tilt to the profile measured at the preceding tilt [17]. The
peak position at any 'V-tilt is then given by
(6.30)

178

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress

Here 28"', 28",-1 are the peak positions of the proflles measured at the nth and
n_lth 'V-tilts respectively, and d28",-1,,,, is the peak shift calculated from these
proflles by Eq. (6.27). Of course, the peak position of the first proflle must be
determined by one of the methods previously discussed if absolute 28 values rather
than peak shifts are required in the subsequent stress analysis.

6.4 Determination of Peak Position for Asymmetric Peaks


In most of the procedures above a symmetric peak was assumed. This is usually
the case for materials that have suffered severe deformation such that no
resolution of the K"t - K"2 doublet is observed. When the doublet is totally or
partially resolved, however, most of the procedures discussed above must be
modified to take into account the peak asymmetry. In the case of the functional
representation:of the diffraction proflles (Sect. 6.3.2) the equations are expressed
as
( 6.31 )
Here the term 1/2 describes the ratio of the intensities of the K"t to K"2 for a given
28 and d is the angular separation of the K"t - K"2 peaks. For example, the
modified Lorentzian equation becomes
1(28i ) = A [ COS1t
A[

+2

( 28i - 28 - 0) ] n
K2
a p
'K2+ (28i -28p )2

cos 1t

(28i -28p -O-d)


a

+ SL28i + INT.

In

K2
. K 2 + (28i - 28p - d )
(6.32 )

In Figs. 6.12a,b" the theoretical fit to experimental data, assuming a doublet


[Eq. (6.32)] (a) and without this assumption (b), are shown. It may be seen
that the correction is especially important in the tail regions. Similarly, the top
15% of the peak can still be fitted by a parabola. The absolute position of
the peak itself, 'however, may be shifted to higher 28 since what is observed
by the detector is the summation of intensities in this range, as described by
Eq. (6.31 ) . If because of defocusing errors (discussed in Sect 6.7) the shape, i.e.,
the degree of resolution of the doublet, changes from one 'V-tilt to the next during
a stress measurement, there will be a shift in the peak position due to this effect
even for a stress-free material. (Since the effect of defocusing changes with 'V-tilt
only in the stationary slit method, this shift is not important in parafocusing.)
This shift can be calculated theoretically by describing the peaks by one of the
functions, Gaussian, Lorenzian or Cauchy, summing up the two K" components
over the appropriate 28 range and then determining the shift of the peak position
of the summed peak by fitting a parabola (Fig. 6.13). This peak shift is then
plotted as a function of half-width (H.W.) Figs. 6.14a,b, and the correction is

179

6.4 Determination of Peak Position for Asymmetric Peaks


1.20
cps

360

..
',',1"%0
0

.0

::- 300

0
0

'Vi

0
0

"

D _.

..

"
06"

:5 21.0 -

'

o~
.

0>

.D
a

..

..

""

,"D~

180

..... "'00

;I

120

1.50

<> "
<> "

.- -. '.

cps

1.00

o~
',
,
a.

ca,'

350 -

....C

300

0
0'

a'
a

00 , --,
0

0"

"

o.
o

250 -

a.
o

"'

'Vi
~

0 ,

<>

00.

.!> .

200 rb

I
150
11.3,1 11.3,1.

1l.17 11.1.,0 1l.l.,3


28 -

11.1..6

11.1.,9 11.5,2

Fig. 6.12. Modified Lorentzian function fitted to the 211 peak from the p-phase of 60 - 40 Brass
(a) with doublet term [Eq. (6.32] ) and (b) without doublet term [Eq. (6.6)]. (From Devine
[4] )

+
Z9 K

([1

29

,R

29 K

"' I

29 K

"'2

Z9 K

29

IXI

Fig. 6.13. Summation of two peaks (K."K. 2 ) and the shift in the apparent maximum of the
summed-up peak. The shift is proportional to the relative intensities of the peaks and to the halfwidth of the summed-up peak

180

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress

H.W.

100

Correction
for
220 peak
0 - brass
28 K = 122.550
01

.050

Fig. 6.14. Variation ofthe peak shift with half-width in degrees 29 for (a) the 220 peak for IX-brass
and (b )211 peak for l3-brass. It may be seen that the correction is also a function of the 29 range

directly read from the chart for any given halfwidth. If such a correction is used, the
wavelength employed to determine the lattice spacing corresponding to the peak
position from Bragg's law must be that of Kill' not the weighted average of
K lll -K..2
Other methods of peak determination such as the centroid, etc., also require
this correction. The cross-correlation function discussed in Sect. 6.3.4, on the other
hand, is a symmetric function even for asymmetric peaks and does not require

6.5 Statistical Errors Associated with the X-ray Measurement of Line ProfIles

181

such corrections. One cannot, however, decide on a given peak determination


method based on the criteria discussed so far: ease of application, effect of
asymmetry etc. In addition to these factors, the inherent statistical error in the
particular method must also be considered.

6.5 Statistical Errors Associated with the X -ray Measurement


of Line Profiles
The intensity measured for a finite time at any 29 position on the line profile
contains a finite statistical error. These errors arise since the arrival of the x-ray
quanta in the detector is random in time. Consequently, the number of pulses (N)
counted for a fixed time t and a given 29 position will have a Poisson distribution
about the true value, with an error proportional to
This uncertainity in the intensity data causes an error in the peak location,
determined from the measured line profile by any of the various methods
described above. The variance, cr 2 (29) , may be obtained from the propagation of
this error through the functions utilized in a given method. In the following this
error propagation is discussed for the parabolic fit.
The statistical variance associated with the 29 location of the apex of a leastsquares parabola fitted to an x-ray peak was first derived by Wilson [19] who
treated the case where intensity was recorded for a fixed time at each observation
point. The extension of these procedures to include the 'fixed count method' (in
this case a predetermined number of counts are accumulated at each observation
point) was given by James and Cohen [12,14]. The extended treatment discussed
here follows that in Ref. [14].
The variance of a general function X, of several variables, X = f( Xl ,x 2,... ,xn ) is

tIN.

cr 2 (X) = (dX)22
dX 1
cr (xd

+ (dX)22
dX 2 cr (X2) + ...

dX dX)
+2 ( dx l 'dx 2 COV(X 1 ,X2 )+ ...

(6.33 )

plus higher order differential terms. Here cr 2 (X) is the variance of X and
cov (xj,Xj ) is the covariance between variables Xj,X j. Neglecting the higher order
differential terms, Eq. (6.33) becomes
dX) 'cov(Xj,xj ) .
cr 2(X) = ~ (dX
-d '-d
1,]
Xj Xj
For linear functions, X =
cr 2 (X)

= L:afV(xd .
j

(6.34)

L: ajxh the standard error is given by


j

(6.35 )

The count rate, Ij == Xj is a random variable having a finite variance. Thus, all
quantities obtained by manipulating it also have a finite variance. Since the

182

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress

measured intensities are statistically independent


cov (xi,xj ) =0 i =l=j
cov ( Xi,Xj ) = 0"2 ( Xi) i = j ,

(6.36 )

Equation (6.34) then becomes


0"2( X) =

~ (ddXX; . ddXXj ) cov (x;,Xj) .


I,J

( 6.37)

This equation is also accurate for non-linear functions if the standard deviation of
the mean is small (less than 20 % of the mean);
From Eqs. (6.23) and (6.37), and substituting the counting rate I j for Xj one
obtains
0"2(29p ) =

)2
t ( d29
dI j
.0"2(Ij)'
P

(6.38)

The variance of the peak is dependent on the variance ofthe count rate, I j, which in
tum depends on how the data are accumulated. The two common methods of
determining the counting rate I j at the diffractometer setting 29j are the fixed time
and fixed count techniques. Defining tj as the time of data accumulation and C j as
the accumulated counts:
Ij = Cit

for fixed time

Ij = C/tj

for fixed counts.

(6.39 )

The variance (which is the square of the standard deviation) of the count rate is
derived by Wilson for both cases [19]:

cri-r (Ij ) = lit

for fixed time

(J~c (Ij ) = If/C for fixed counts

( 6.40)

Equations (6.40) predict an asymptotical approach to the true value of the count
rate with increa'sing time or total number of counts collected. This behavior is
shown in Figs. 6.15a,b. It can be seen that there is a very good match between
measured and predicted errors for both cases.
The differential in Eq. (6.38) is now derived. Rewriting Eq. (6.23),
29p =290 and

(52 (n~-nOn4) {
2
2
Mt/(n 2() M o -noM 2 )}
n2

where

aMI = ()I"+1'"
__
J'
OIj

(6.41 )

(6.42)

6.5 Statistical Errors Associated with the X-ray Measurement of Line Proflles

183

50
+ Observed
o Statistical

Vi ['0 C
::l
0
U

-=~ 30 "a;

en

: 20
c:

j
II

::l
0

Gl

!jl

10 _

rIlJjl
I!!
0

L.OOOO

II

120000
80000
Counts

160000 200000

[,2
35 ~

+ Observed
o Statistical

Vl

628 i-

u
L

~ 21 i-

en

c:

1[,

i-

+
0

+
0

7i-

b
0

!jl

10

20

30
Time

III
I

['0

50

s 60

Fig. 6.15. Comparison of statistical error in intensity measurements with observed error for (a)
fixed counts and (b) fixed time measurement

Combining Eqs. (6.38), (6.40), and (6.44), the error in peak location due to
counting statistics can be obtained.
For accumulation of data by fixed time (at each observation point j) the
formula is
2
o4(n~-nOn4)2
cr (2ap ) = 4n~(n202Mo-noM2)4
n

L ( noo3M J2 + (n202Mo - nOM2 ) o2j - n 203M d 21/t, (6.45)

j= -n

where cr ( 2ap ) is the standard deviation in the peak position and the term Mi is the
ith moment of intensity about the peak. For perfectly symmetric peaks the odd
moments of intensity will be zero and for small deviations from symmetry (at least

184

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress

Table 6.2. Comparison of exact statistical formulae to approximate formulae for the counting
error in peak location using a parabolic fit". (The data is corrected for Lorentz-polarization and
absorption factors as discussed in Sect. 3.2 )
0"
er( 29p )
er(29p )
W
% Chang!
("29) (29) Eq.(3.32) (29) Eq.(3.33) (29) in er(29p )

Sample

II'
Degrees

1090-1

0
0
45
45

3
7
3
7

0.11
0.03
0.15
0.05

0.45
0.45
0.50
0.50

0.00085
0.00105
0.00139
0.00181

0.00085
0.00105
0.00139
0.00180

0
0
0
0.3

TBAG-5

0
0
45
45

3
15
3
15

2.83
0.38
2.35
0.38

6
6
6
6

0.01023
0.0134
0.0147
0.0146

0.01023
0.0133
0.0147
0.0144

0
0.8
0
1.37

No. of data
points

" 0 represents the 29 step increment between data points. W is the full width at halfthe maximum
intensity. The o/er. change represents the pet. difference in 29p between the statistical errors as
calculated from Eqs. (3.2) and (3.3). From James & Cohen [12]

in the peak region being fitted by the parabola) these terms will be negligible. In
such cases Eq. (6.45) becomes
2 9 _
04(n~-nOn4)2
n!:2
2 4'2
cr (2 p) - 4.. 2( 02M
M )4.L (n2u Mo-noM2) 0 J Ij
0 - no
2 J =- n
t n2 n 2
04 (n~ - nOn4) 2

4-t-n~(n202Mo-noM2)2

'04

f.

j=-n

j2I.
J

05 (n~-nOn4)2'M2

= 4.t'nHn202Mo-noM2) 2 .

(6.46 )

Similarly, in the case of fixed counts:


04(n~ -nOn 4 ) 2

L (no03 MJ2+ (n2 02Mo-noM2)02j

j= - 0

-n 203 M 1 ) 2If/C.

( 6.47)

Neglecting the odd moments of intensity, Eq. (6.47) becomes


cr

2(29 ) _
P

08(n~-nOn4)2
0'2 2
4Cn 2(
02M 0 - no M 2 )2 J. =L- 0J I j .
2 n2

(6.48)

The validity of Eqs. (6.45) - (6.48) have been checked experimentally by James
and Cohen [12, 14] (Table 6.2). In practice however, the exact formula (6.45)
or (6.47) should be used. Although longer, these equations are easily in-

6.5 Statistical Errors Associated with the X-ray Measurement of Line ProfIles

185

corpora ted into software for microcomputer controlled experimentation and take
into account the presence of some asymmetry in the peak profIle.
Similar procedures may be utilized to determine cr 2 (2ep ) for the other
methods of peak location. For example, for the cross-correlation method the
variance in peak shift due to counting statistics is [16]
cr 2 (A2e) =

1(a (:~je) )

2 cr 2

(6.49 )

(Fj ) ,

where
(6.50)
and
cr 2 (Fj ) = L
i

{(12

(lP) + cr 2 (lr- j ) }

(6.51 )

(where the terms are defined in Sect. 6.3.4).


If one uses a PSD to determine the total peak profIle (this is usually the choice,
since step-scanning through the total profIle may take a long time to obtain the
required accuracy) the intensity at all 29 is counted for the same time (i.e., the
fixed time method). Thus, from Eqs. (6.40) and (6.49),

cr' (Ale"") ~ t

{c. (j-k) L{~,r (i-k) }

'cr' (F

j ) ,

(6.52 )

where
(6.53 )
The calculation of the error in the case of the functional representation is not as
straightforward as the methods discussed above. A much larger number of
parameters are involved and propagation of error through non-linear regression
procedures may be necessary. One other problem associated with these methods
arises directly from the large number of definable parameters; the calculated
solution may not be unique and may not be the correct one.
In the discussion above, the general methods of peak location and the errors
associated with them have been discussed. There are numerous articles comparing
these methods for general and particular applications and usually there is little
agreement as to the"best' method of peak location for particular cases, although
the work by Knuuttila [16], James and Cohen [12, 14] and Devine [4] indicates
that the parabolic fit is usually the best choice. This method, and the error
equations associated with it, has been extensively treated in the literature and may
be used with confidence in most cases. In general a parabolic fit with 7 or more
points to the top 15 % of the peak with background correction appears to be the
most versatile method that can be applied over a wide range of peak shapes with

186

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress

minimal error. The effect of doublet correction must be taken into account if the
stationary slit method is used.

6.6 Statistical Errors in Stress


The variances in the measured peak positions result in an error in the residual
stress calculated from them. This error can alsQ be determined through error
propagation techniques. In the following this procedure will be examined for
various methods, as per James [14].

6.6.1 The

8in 2,&,

Technique

In this method, the residual stress ad> is determined from the slope m of the
interplanar spacing "<1,.," vs. sin 2tp (Sect. 5.5). The terms "<1,.," are obtained from
the 29 values determined from the measured line profIles and, thus, will also
contain statistical errors. The variance in "<1,.," due to a 2 (29p ) can be determined
through Bragg's law:
(6.54)

nA = 2<1,.,sin 9p
From Eqs. (6.35) and (6.54),
a

2 ( rl

..",

= [() (1../2 sin 9p )


(}9

2 2(9 ) = (A cos 9p ) 2. a 2 ( 29p )


a
p
2sin29p
2

~) 2

180'
(6.55 )

where 29 is in degrees and a 2 (29 p ), the variance in 29p is determined from the
variance equation for the particular method of peak determination [Eqs. (6.45),
( 6.47) for example]. The surface stress is given by m', the regression coefficient
describing the slope of a least squares line fitted to the data

a.

L
m'=

(Yi-Y)
L{Xi -X)2

(~-x)

(6.56)

Here x is the independent variable (sin 2tp) and y is the dependent variable (<1,.,).
Assuming that tp is determined exactly (this is a valid assumption for a well
aligned diffractometer and a flat sample) , and that all <1,., are measured to the same
precision, i.e., a 2 ( <1,.,) is constant (this would be approximately the case if a fixed
count procedure was used for all profIles) , the variance in m' is [from Eqs. (6.35)
and (6.56)]

L (sin2tp-sin 2tp )2a 2(doJ>1j

a 2 (m') = -=Nc:..::..,F-------~,___-

[~ (Sin2tp -sin2tp )2T

N..,

(6.57)

6.6 Statistical Errors in Stress

187

where cr 2 ( d,.,) is given by Eq. (6.55). The surface stress cr, is given by

thus, the variance in the surface stress is given by


(6.58)

6.6.2 Two-Tilt Technique


Application of Eq. (6.34) to the two-tilt method [Eq. (5.17)] yields

(6.59)
where the stress constant K is given by
K=

( 1 + v) . sin 2 ",

( 6.60)

If the peak shift is small, the two-tilt equation is commonly expressed in terms of 28
[Eq. (5.19)]. The variance in surface stress cr, in this case is given by
(6.61 )
In this equation it is assumed that the stress constant K' [which is, in fact a
function of 28, Eq. (5.20)] is known exactly. This assumption is valid for small
errors in 28.

6.6.3 Triaxial Stress Analysis


The extension of the procedure discussed above to triaxial stresss analysis
[Eqs. (5.15), (5.23)] was given by Rudnik and Cohen [20]. In this analysis the
terms a l and a 2 are used (Sect. 5.22):

188

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress

The variance in aI' for example is [from Eq. (6.34)]


2
(dal
(J (ad= d(d",'I'+)

2
( d a l ) 2 2 (d
)
.(J (d",'I'+)+ d(d",'I'_)
.(J
"''1'-

dal ) 2 2
+ ( dd o
. (J (do).

(6.63 )

Assuming that do is known exactly 1, Eq. (6.63) becomes


(J2(ad = (d( ::;+) ) 2'(J2

(d~'I'+) + (d( ::;_) ) 2'(J2 (d~'I'_) ,

(6.64 )

where the variance of the terms d~+,d~- are given by Eq. (6.55). Thus, the
variance of the slope of the a l vs. sin tp plot becomes [from Eqs. (6.55), (6.56),
( 6.57), (6.64)]

(6.65 )

(6.66 )
The triaxial analysis determines the strain E33 from the intercept of a l vs. sin 2tp.
The intercept la" of the least-squares line is given by

-.= L:(al )i
Y
,
(Na,) ,
No. of tp-tilts.

( 6.67)

la, =Yi-ma,X;,
N a ,:

The variance for the intercept of a l vs. sin 2tp can also be determined from the
application of Eq. (6.34) to Eq. (6.67). Determination of strain terms usually
involves the slopes of a l vs. sin 2tp at various <1>, for example for E 12 :

E12 = (ma.)~=45-"2 (ma.)~=o-"2 (ma.)~=90+ (la.)~

(6.68)

If this value is experimentally determined, the error term must be evaluated from the particular
"do" determination method and substituted in Eq. (6.63).

6.6 Statistical Errors in Stress

189

Thus, the variances of the strain terms are then determined from the variances of
the slopes of a 1 vs. sin2\j.l for tI> =0,45,900 and from the variance of the intercept
at a given tI> tilt. The variance of E12 is
cr 2 (E 12 ) = cr 2 (la!)4> + cr 2 (ma.)4>=45 + ~ (cr 2 (rna.) 4>=0 + cr 2 (rna! )4>=90) ) .
(6.69 )
After the strains are determined, stresses are determined from Hooke's law:
( 6.70)
The errors in stress can be determined from the propagation of the error through
Eq. (6.70) for any particular case.

6.6.4 Statistical

~rrors

in X-ray Elastic Constants

In the procedures discussed above, it was assumed that the elastic constants viE,
( 1 + v) IE or CjjkJ have no errors associated with them. If, however, the x-ray
elastic constants 8 1 ,8 2/2 (8ect. 5.13) are used in the analysis, the error associated
with these terms must also be included in the overall analysis. In the following
formulas that can be used in evaluating the statistical errors in 8 1 ,8 2/2 are
discussed.
The experimental evaluation of 8 1 ,8 2/2 and the errors associated with this
measurement have been discussed by Perry and co-workers [21, 22] and was
reviewed in 8ect. (5.13). The basic flowchart of the measurement is shown in
(Fig. 6.16). Using the procedures discussed above, the error in the slope of the
"m" values versus applied load (Fig. 6.16b) will be
cr 2 ( mil) =

L [( crOPI No-a [

No-a

crOPI) 2 cr 2 (m j ) ]
2] 2 '
(cropi - crOPI)

(6.71 )

where N"a: no. of applied loads, cr4>1 is the applied load at ith experiment, and
cr 2 (m') is given by Eq. (6.72). The error in 8 2/2 is
cr 2 ( 82 /2)

cr 2 (mil)

(6.72 )

= ( do ) 2

Here do is the unstressed lattice spacing.


8imilarly, the error in 8 1 , which is evaluated from the slope of d4>'V=o vs.
applied load line is

1
cr 2 (8 ) 1 -(d o )2

L
No-a

[(crop-crop)2cr 2 (d'V=0)]
(

[L(crop-crop)2]2

6 73 )
.

No-a

where the variance cr 2 ( ~= 0) can be calculated from Eq. (6.55) for


load levels.

\j.l =

0 at all

190

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress


d

d
a'opp

./

= a'<t>1

, ------,
i) Data acquisitron

b
Analysis

Fig. 6.16a-c. Schematic representation of the x-ray


elastic constant measurement. a The slopes (m';) and
intercepts (d,p=o); of least-squares lines fitted to
experimentally determined d vs. sinz\jJ data at
various applied loads are calculated. b The slope of
m'; vs. applied load, m", is proportional to the x-ray
elastic constant Sz/2. c The slope of the intercept
(d,p=o) vs. applied load is proportional to the x-ray
elastic constant S1

6.7 Instrumental Errors in Residual Stress Analysis


These errors dep,end on the particular configuration and instrument parameters
used in the measurement of the line profiles. They may be broadly classified into
two groups: errors dependent on beam optics and alignment errors. In the
following these errors are examined following the work of Zan to pulos and Jatczak
[23J, Marion and Cohen [24, 25J , James and Cohen [14, 26J and others [7, 27J.
a) Errors Due to"Beam Optics

Beam focusing depends on the geometric arrangement of the diffraction


apparatus, the particular slit arrangement in the beam path, and on the vertical
and horizontal divergence allowed in the beam by these slit systems. Various
important effects of the beam optics for diffractometer geometry are discussed
below.

6.7 Instrumental Errors in Residual Stress Analysis

191

6.7.1 Variation of the Focal Point with 9 and 'I'


In a diffractometer operated in the focusing condition (Fig. 6.17), the divergent
rays produced by the tube D will diffract from the specimen S and focus at a point
A on the goniometer circle, which is then scanned by the receiving slitdetector combination or detected by a PSD. The radius of the focusing circle
(Fe) is related to the radius of the goniometer circle (GC) by the equation
R
FC

(6.74a)

RGC

2sin9

for 1.1' = O. The correct focusing position is maintained in diffractometer by the 1:2
ratio of the 9/29 movement at all times. If, however, the specimen is rotated
around the center of the diffractometer such that 1.1' =F 0 (Fig. 6.18) , as in a stress
measurement, Eq. (6.74a) becomes

R
FC

RGC

(6.74b)

2 ( sin 9 +"'P )

Since the radius of the focusing circle changes with 1.1', the rays arriving at A will be
defocused and the mean position of the maximum intensity will be shifted,
creating an error in 29. One method of avoiding this error is to move the receiving
slit to the focusing point B. The distance from the sample to B is given by
R' - R
p-

Fig. 6.17

GC

. cos{l.p+ (90-9)}
cos{l.p- (90-9)}

E
0/=0

(6.75 )

Fig. 6.18

t=l/t

Fig.6.17. Focusing geometry for a specimen of ideal curvature for '1'=0. Here GC is the
goniometer circle and FC is the focusing circle
Fig. 6.18. Focusing geometry for a divergent beam when a specimen of ideal curvature is rotated
ljJ-degrees from the normal focusing position. It may be seen that the radius of the focusing circle,
and thus the radius of curvature for such an ideal specimen, is different than the one required for
focusing in the normal condition (Fig. 6.17)

192

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress

This procedure is called the "parafocusing method". To maintain true focusing, the
position of the receiving slit must also be changed with 28 (at a given tp) as the line
intensity is being recorded. This is not always practical, however, and Marion and
Cohen [24, 25] recommend moving the slit into a compromise position
corresponding to the center of the peak and leaving it there as 28 is changed. If the
position of the slit is changed with 28, on the other hand, the length of the
diffraction cone seen by the receiving slit changes, and the scattering factor IF21 is
proportional to
IF21(XP2oRp,

(6.76)

where R~ is the distance from the specimen to the receiving slit [Eq. (6.75)] and
P~8 is the intensity measured at any point 28 along the line profile. The correct
intensity at 28 along the profile must be obtained by multiplying the observed
intensity at each point by R~.
Parafocusing requires the accurate positioning of the receiving slit during the
measurement. This is not always practical and usually the 'stationary slit' method
is used. In thi~ method the slit and the detector are left on the goniometer circle at
all times, deliberately not fulfilling the focusing condition. This procedure results
in lower diffracted intensity received at the detector. However, the complication of
moving the slit or the PSD exactly along a given radius is eliminated. Any
deviation from the radius will cause an apparent shift in 28 since the angular
relationship between the direct beam and the receiving slit or PSD changes during
the measurement. Even if parafocusing geometry is correctly used, however, xrays focus to an area, not to a point, due to the effect of sample shape and there is
still some residual error in the peak location. This effect is treated below.
6.7.2 Effect of Horizontal Divergence on Focusing

For finite x-ray beam dimensions, true focusing requires the conformation of the
specimen surface to the focusing circle at all tp-tilts. This requires a concave
specimen whose radius of curvature is continuously variable during the stress
measurement. Such specimens, however, are rarely encountered in practice. Nonconforming specimens produce a focusing aberration such that the focus is no
longer a point (such as A or B in Fig. 6.17, 6.18) but has a finite area. This
departure from the ideal focusing condition is shown in Fig. 6.19, 6.20 for a flat
specimen and a convex specimen respectively.
Marion and Cohen [24, 25] have given a simple equation relating the error in
peak location to beam divergence. Consider Fig. 6.19; using the law of sines for the
triangle XOD one obtains

o D = RGcs~n ( (X/2)

(6.77 )

smx

From triangle AOD and Eq. (6.77):


AO= ODsin(x- 2<j
sin ( (X/2 )

RGc sin (X-2<j


sin X

(6.78 )

6.7 Instrumental Errors in Residual Stress Analysis

193

Sample surl ace

\\

\ J

t---I

ormal to
diffracting plane

Fig. 6.19. Schematic of the effect of the horizontal beam divergence for a flat sample (IX is the
incident beam divergence, q,=90-9, d=90+1p-q,-1X/2 and X=90-1p+q,-1X/2)

?II

~J
/

I'

,I .{
a

l....
- --'r--- - /

Fig. 6.20 a,b. Departure from the ideal focusing condition for a convex specimen. Due to
the non-ideal curvature, the rays from left
(I), right (r) and center (c) do not meet at
one point causing a broad focus spot

1/

(L

Similarly, from triangles XCO and BCO:

co = RGcs~n
( r:t./2 )
sm.1
BO __ CO'sin (.1 + 29)
sin ( r:t./2)

(6.79)
RGc'sin (.1 + 29) /sin.1.

194

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress

-D2
c:i

CD

Q)

<I
t.o

_DID

130

140

28

150

160

Fig. 6.21. Variation of the peak-shift as a function of 29 due to a horizontal beam divergence of 1
(a.=1)

The width of the focusing aberration P is


AB
(BO-AO)
p= T tan (ex/2) =
2
tan(ex/2) ,

( 6.80)

where AB = BO - AO and can be evaluated from Eq. (6.78), (6.79). The "peak
shift" due to this aberration in degrees 29 is given by

A29=~180
,
R'
p

(6.81 )

where R~ is given by Eq. (6.75).


If a biaxial stress is assumed, usually the peak shift between 'I' = 0 and 'I' = 'I' is
of interest. This quantity is defined as:
( 6.82)
variation of 0 (A29) DO as a function of 29 and 'I' is shown in Figs. 6.21, 6.22 for
ex = 10 and 30 respectively. If 0 (A29) is negative, the shift is to a lower value of 29
with increasing '1'. Thus, a stress-relieved flat plate will show an erroneous tensile
stress if this errQr is not taken into account in the analysis. Similarly, a convex
specimen will exhibit a peak shift to lower 29 values. This is also the case for
concave specimens whose radius of curvature is greater than that of the focusing
circle. On the other hand, concave specimens whose radius is smaller than that of
the focusing circle at a given 'I' cause peak shifts to higher values.
Because of the unequal distribution of diffracted intensity between the central
beam and the right and left hand portions, the peak shift given by Eq. (6.82) is an

6.7 Instrumental Errors in Residual Stress Analysis

195

_.10 0
c

III
Q)

C\J

<J

IX)

-.05

t= 15~
130

140

28

150

160

Fig. 6.22. Variation of the peakshift as a function of 29 due to a horizontal beam divergence of 30
(a.=3)

upper limit. Zantopulos and Jatczak [23] have shown that the deviation of the
centroid [Eq. (6.4)] of the diffracted beam from the true value approaches a
limit of (1/3) of 0 (L\29) given by Eq. (6.82) (the centroid in this case was
determined by a computer iteration method in which the divergence angle was
divided into 1000 parts). James [14] suggests the use of 1/2 of the value of
o(L\29). As long as the divergence angle is small, however, (1 0 for example), the
error due to this effect is small over the tp-ranges usually employed (Fig. 6.21 ).

6.7.3 Effect of Vertical Beam Divergence


A peak shift arises due to vertical divergence when there is strong texture. This
effect is important only in the case of weak peaks where the grains that diffract are
those with planes slightly tilted from the orientation dictated by Bragg's law for a
parallel incident beam. Pike [28] has analyzed this situation for a variety of slit
conditions and showed that
(6.83 )
where 2h is the beam height (dictated by the height of the receiving slit) and
Ql,Q2 are constants that depend on the slits in use and are given Table 6.3.
Marion and Cohen [25, 26] have shown that a very strong texture is required to
produce significant peak shift due to this effect. Since in the presence of such
texture "d" vs. sin 2 tp is usually oscillatory, stress analysis may not be possible. This
correction may be important, however, in the determination of the effective elastic

196

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress

Table 6.3. Values to the constants 01 and 02 for textured specimens (From Ref. [24])

No soller slits
One soller slit
Two soller slits

Strong peak

Very weak peak

01 02

01 02

2
1
0

1
0
0

2
2
2

constants of textured materials where analysis of oscillatory "d" vs. sin 2 '1' plots is
required.
In the discussion above, the effect of deviations from the focusing geometry
were reviewed. A quantitative comparison of the error introduced in the stress
determined from peak position data measured by the parafocusing and the
'stationary slit' methods has been given by James and Cohen [14, 26]. Using an
automated receiving slit bracket under computer control, they performed fifteen
replicate measurements on a steel sample (211 peak, Cr radiation) with and
without slit movement. They report an average stress value of -164 MPa for the
parafocusing geometry and -170 MPa for the stationary slit geometry. The
observed error of one standard deviation from the mean value is 3.6 MPa for the.
stationary slit geometry and 7.2 MPa for the parafocusing geometry. Even
though the repositioning of the slit during focusing was very accurate, it caused an
error of 4 MPa. Based on these considerations, the use of the stationary slit
method with appropriate slits limiting the horizontal divergence to less than 2 is
recommended.

b) Alignment Errors
As the name implies, these errors originate from instrumental misalignment. Three
important cases of misalignment contribute to errors in residual stress measurement: sample displacement from the center of the diffractometer, effect of'l'-axis
not corresponding to the 29 axis, and missetting of the true zero of the 'I'-rotation.
These effects are discussed below.

6.7.4 Effect of Specimen Displacement


If the effective diffracting volume is not located at the center of the diffractometer,

there is a relative peak shift between '1'=0 and '1'='1'. This is shown in Fig. 6.23
for the case of the O-goniometer. The shift in degrees, 29, is given by
() (t129) SD = t129..,= 0 - t129..,=..,

= 360t1X cos 9 . [_1_ _


1t

RGC

. .sin 9
] .
Rpsm (9 + '1')

(6.84 )

6.7 Instrumental Errors in Residual Stress Analysis

197

DISPLACED
SAMPLE

CORRECT
- - - - SAMPLE
POSITION

PEAK SHIFT
FROM 4.1= 0
TO 4.1 =\.1.1

Fig. 6.23. Schematic of the effect of the sample displacement on the


peak position when the specimen is rotated Ijl degrees

VI

Q)

C\J

<I

130

140

28

150

160

Fig. 6.24. The variation of the peak shift due to sample displacement of 0.0254 mm back from its
proper position over the center of the diffractometer (Ax = - 0.025 mm)

Here.1X is the sample displacement, R~ is given by equation (6.75) and Rocis the
goniometer radius. The peak shift due to this effect is shown as a function ofe and
\jJ in Fig. 6.24 for .1X= -0.025 mm.
For cubic materials a simple method for checking sample displacements on a
diffractometer has been discussed b, Cohen [29]. Consider Fig. 6.25 (\jJ = 0):
.1X/sin e
sin .12e

'" Roc _
Roc
Roc
sin ( 2e - .12e) - sin 2e 2 sin e cos e

( 6.85a)

198

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress

+------~**~=-~\

sin 8

Fig. 6.25. Definition of the pertinent angles in calculating the


effect of sample displacement on the position of 28

for small angles:


sin A2e ~ A2e ~ 2AX cos e/RGC .

(6.85b)

Differentiating Bragg's law:


0= 2d cos eAe + 2Ad sin e

(6.86 )

substituting from Ae from equation (6.85b):


Aa
Ad
A2e
AX cos 2 e
a =(f=-2tane =-RGc ' sine
=

ahkl-a O
AX
= - - 'cot ecos 9 .
RGC
ao

(6.87 )

Here ahkl is the lattice parameter calculated from the peak maximum and a is the
extrapolated true lattice parameter.
Equation (6.87) can be used in measuring AX via x-ray diffraction. The lattice
parameter ahkl is determined for various reflections (at 'I' = 0) and plotted vs.
cos 2 e/sin 9. The slope of this line is equal to - AX/RGc, For a o one can take the
intercept of the plot. In this type of analysis a positive slope indicates negative
displacement, that is the sample is displaced to a position behind the center of the
diffractometer, -away from the x-ray source and the detector.
If the material is susceptible to stacking faults, however, this method may yield
an erroneous displacement. The lattice parameter in such cases is also affected by
other factors, as discussed in Sect. 5.7. When the specimen displacement is finite
Eq. (5.27) becomes

+ (_

AX 'cot 9 cos 9)1,


RGC

(6.88 )

where the terms are defined in Sect. 5.7. In such a case A29hkl must be measured for
at least five reflections and the unknown parameters, including AX, must be
determined from a least-squares analysis of the measured values as discussed in
Sect. 5.7.
1 This term is for stationary slit geometry. If another method is used, the appropriate equation
for sample displacement must be substituted here.

6.7 Instrumental Errors in Residual Stress Analysis

199

6.7.5 Effect of 'P-axis not Corresponding to the 20-axis


For correct alignment, the axis of inclination of the specimen for tp-tilts must be
coincident with the axis of29 rotation. If this condition is not satisfied, a peak shift
will occur upon tilting the specimen (Fig. 6.26). Marion and Cohen have given an
equation to evaluate this error. From the triangle ABC:
AC=

ABsintp
sin ( 180-tp-9)

AX'tan ( tp/2 ) sin tp


sin ( 180-tp-9) ,

(6.89)

similarly, from triangle ACD:


A29=

AC~~29

( 6.90)

where R~ is given by Eq. (6.75). From Eqs. (6.89) and (6.90):


A29 ~ AX:tan (tp/2) sin tp si~29
sm( 180-tp-9) Rp

(6.91 )

which upon simplification becomes


A29 ~ 2 (AX') si? ~ cos 9 ( 1 - cos tp )
R psm(9+tp )

( 6.92)

The relative peak shift, 3 (A29) 'l'A' is given by


3(A29)

'l'A

=A29 _ -A29 _ =2AX'sin9cos9(1-costp)


'1'-0
'1'-'1'
R'psm
. (9
.
+tp)

(6.93 )

The effect of this displacement is shown in Fig. 6.27 for various 29,
tp (AX' = - 0.025 mm).

--+-------~~-7~.-~--+------

SAMPLE
0/=0

Fig. 6.26. Definition of the geometry for the


displacement of the Ip-axis from its proper
position; point A is the axis of 2e and point 0 is
the axis ofe.IjI=90-e

200

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress

.02

<l

Q)

<J
fA)

.01

130

140

28

150

160

Fig. 6.27. Variation of the relative peak-shift due to the presence of a 'I'-axis displacement of
-0.025 mm

6.7.6 Error Equations for the tp-Goniometer


In the procedures discussed above, the error equations applicable to the Q
goniometer geometry (Fig. 4.22) were discussed. For the case of the 'Pdiffractometer similar considerations apply. In this geometry, the 'P-tilts are around
an axis parallel to the plane of the diffractometer (Sect. 4.20). This method of
tilting preserves the 1:2 relationship between 9 and 29 at all times during the
measurement and thus defocusing errors due to the movement of the focusing
point are not important for the 'P-goniometer. There will be, however, focusing
errors due to the deviation of the specimen surface froJ;ll the focusing circle
(horizontal divergence) and effects from vertical divergence as well as specimen
displacement errors. In the following these terms are briefly discussed following
the treatment given in Ref. [27].
1) Errors Due to Sample Displacement
If the irradiated area C on the specimen is displaced from the center of the
diffractometer (as shown in Fig. 6.28), the error in 29 at a given 'P is given by

~ (29)1jl =

R2 cos 9 [cos 'P(X.)1jl + sin 'P.(z.)1jl] .

(6.94 )

GC

The peak shift between 'P = 0 and 'P = 'P is then given by

3(~29) = R2

GC

.COS9[(_1cos 'P

-1)X+~ tan'PJ '

( 6.95)

6.7 Instrumental Errors in Residual Stress Analysis


incident
beam

201

_specimen

.-'

." aXIs

..

..

Fig. 6.28. Definition of the displacement terms in "'goniometer geometry

where Roc is the diffractometer radius. In a typical diffractometer where


Roc=145.54 mm (5.73") and 29~156 (211 Fe peak with Cr radiation),
equation (6.95) becomes
3 (a29) = 0.067 X+ 0.164~, ('1'2 = 450 for this case)
where 3 (a29) is in degrees 29 and the displacement terms are in

( 6.96)
Mm.

2) Effect of Specimen Curvature

This effect is due to the deviation of the specimen surface from the focusing circle.
For flat specimens, the error in 29 is given by
1
sin 29
a(29)= 12sin2 (9+'I')
where

(1.

( 6.97)

.(1.2,

is the horizontal divergence of the x-ray beam.

Circle

/
RGC

Fig. 6.29. Schematic representation of diffraction from a


cylindrical specimen in "'goniometer geometry

202

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress

In the case of cylindrical geometry (Fig. 6.29), the displacement of the x-ray
peak due to this effect is
_ x( 3) . e. {RGC
1-4 sin 2e sin2'P}
- 3R2 cos
R cos 'P +
. e
GC
1
sm
+

A 2 -A 1
.
3
3
R
cos e sm'P, x(3) = (AI +A 2 )/(A 1+A 2 ),
GC

(6.98 )

where the terms are defined in Fig. 6.29. The peak shift error between two 'P-tilts
'P = 0, 'P = 'P is given by
.O(A2e)=cose.z 2 .(_1__ 1),
3RGC Rl
cos 'P

(6.99)

where Z is half the height of the irradiated region.

6.7.7 Effect o( Errors in the True Zero Position of the '1'-Axis


Assume that the zero position for the 'P-movement, read from the odometer of a
diffractometer, is different from the geometric definition of'P = 0 by A degrees, i.e.,
(6.100)
where A is defined in Fig. 6.30. Substituting Eq. (6.1 00) into the Eq. (5.15)
(assuming 0'13 = 0'23 =0):

(e~3)"''''= d"''''d~do =K",{cos 2A-sin2A}sin2'P+ ~"'sin2Asin2'P


(6.101 )

From Eq. (6.101) it is seen that an error A from the true 'P-position will cause
splitting proportional to K",sin2A for positive and negative 'P. The shape of the

X-ray
tube

detector

Fig. 6.30. Definition of the tp-missetting


error in a diffractometer (a-goniometer)

6.7 Instrumental Errors in Residual Stress Analysis


irradiated

region

I
I
I
N8 I
I
I

203

~I

I
I
I
I
I
I

~----:::::=r-::::;:::::::----I TA

90-t t

Fig.6.31. Variation of true 'I' with


position over the irradiated area of a
curved specimen

split d vs. sin 2 \jJ pf-ot will be the same as the psi-splitting caused by 0" 13,0"23 (with
.::\ = 0) since in both cases the effect is caused by the argument sin2\jJ. However,
while for true \jJ-splitting the dcjJljJ+ ,dcjJljJ_ will be interchanged if <I> is rotated 180, no
such change will occur in the case of the split caused by .::\ since
(6.102 )
If the specimen surface exhibits curvature, the true \jJ-position will vary along the
surface. In those cases where the dimensions of the beam are large compared to the
radius of curvature, the measured d value will be an average over a range of \jJtilts. If the odometer of the diffractometer is showing \jJ = \jJ A for the center of the
specimen, the true \jJ at any point within the irradiated area will be
(6.103 )
where 90-\jJI is the angle between the tangent to the surface at that point and the
normal to the diffracting planes at that point in the irradiated area (Fig. (6.31 ). It
may be seen from Fig. 6.31 that the maximum deviation from the mean value will
depend on the specimen radius and width of the i~radiated area.
Various estimates of the error this effect causes have been given. For biaxial
stress states the errol is less than 5 % [27] for situations where the ratio of the halfwidth of the irradiated area to the specimen diameter is less than 0.13. No work on
triaxial analysis is yet available. The best method of dealing with this situation is to
use a narrow beam such that there is little curvature within the irradiated region.
The loss of intensity due to the narrow incident beam may be compensated by
increasing the height of the beam.
6.7.8 Alignment Procedures
In Table 6.4 the effect of some of the errors discussed above are shown. It can be
seen that the most important sources of error are the displacements of the sample
and the \jJ-axis from the center of the diffractometer. The effects of the sample
displacement may be minimized by using the analysis method discussed above to

204

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress

Table 6.4. Typical instrumental errors in peak location for 2!bd56 (approximately the 211
CdC" peak from iron) (From Ref. 14)
Cause

Horizontal beam
divergence (IX = 0.5)
(Taking 1/2 (o(~26hD) see text
Vertical beam divergence
(Assuming strong texture
using divergent Soller slit
no receiving Soller slit)
Sample displacement
~= 0.025 mm
II'-axis displacement
~x' = 0.025 mm
Maximum total errors+
a) in - 26 direction
b) in + 26 direction
Maximum error in
stress for steel'

Peak shift between


11'=0 and 11'=45

Peak shift between


11'=0 and 11'=60

-0.0006 26

-0.0025 26

- 0.002 26 or
+0.002 26

-0.002 26 or
+0.002 26

-0.0034 26 or
+0.0034 26
- 0.002 26 or
+0.002 26

-0.0088 26 or
+0.0088 26
-0.0068 26 or
+0.0068 26

-0.008 26
+0.0068 26
-4.74 MPa ( -690 psi)
or
+4.0 MPa ( -585 psi)

-0.0201 26
+0.0149 26
-7.0 MPa ( -1150 psi)
or
+5.9 MPa (+860 psi)

+ Note: Maximum error is either one of these but not the total range.
Calculated for steel from 0"<1> = K.v (26J. - 26'l') where K45 = 593 MPa;o26 and
396 MPa;o26.

K60

measure the offset and displacing the specimen to the correct position. Alternately,
it is also possible to evaluate the error in stresses such a displacement causes and
correct the measured stress values. One method of stress measurement that
minimizes the effect of specimen displacement is the parallel beam method
(Fig. 6.32) where the 20 position is determined uniquely by the angular
relationship between the parallel slits (these are usually soller slits placed with the
baffles perpendicular to the goniometer plane). In Fig. 6.33 the effect of specimen
displacement is shown for parafocusing, stationary slit and a parallel beam. The

~etector

X-ray tube

5011""'\'
\

,\ ,,
,, \ \ \ \ \
\

-------

'

\ , \' ,
~~
\

Specimen position

---- ------

Fig. 6.32. Schematic of parallel beam geometry. The diffraction angle 26 is uniquely defined
by the angular relationship between the two
sets of soller slits and is independent of sample
position (dashed lines or the solid line)

6.8 Corrections for Macrostress Gradients

205

350.-.----.-----,----.-----,-----,----,-----.-----r~

300
250
o

~ 200

il

.:: 150

V1

7-

___Parallel beam

-_-_~_--_--_6--

Stationary slit

50/

100

O~~--~--

-2.0

-1.5

__- L_ _ _ _
-1.0
-0.5

J __ _ _ _~_ _- J_ _ _ _~_ _ _ _~_ _ _ _~

0.5

1.0

1.5

2.0

Displacement (mm)

Fig. 6.33. The dependence of stress upon three geometric focusing conditions on a 1405 steel
sample. The solid lines represent the error as calculated from Eq. (6.84). The dashed line
represents the slope of the actual data with a parallel beam. From James [14]

error is maximum for parafocusing and negligible for parallel beam geometry. The
barnes, however, restrict the amount of diffracted intensity received by the
detector and for a comparable statistical error counting times up to three times
longer may be necessary.
An alignment method that will indicate the presence of all aberrations except the
tp-missetting error [Eq. (6.101)], or variation oftp within the irradiation surface, is
the "annealed powder" method. In this case a very thin film of annealed powder is
brushed on the surface of the specimen irradiated by x-rays, and the stress in the
powder is measured by the residual stress determination procedure that will be
used on the actual specimen. The position of the specimen is then adjusted until the
stresses measured f:t;"om the powder are negligible within the statistical error. The
error from the powder can then be used as an estimate of the error in the
stress determined from the specimen after the powder is brushed away.
However, since all errors are strongly dependent on 29, the powder peak from
which the errors are measured must be close in 29 to the position of the specimen
peak. It must be noted that this method can not detect the presence of "tpmissetting error" since the stress term Kq, in Eq. 6.101 and thus the term Kq,sin2tp
which causes the split is zero, independent of!l. and an unsplit plot will be obtained
even if !l. is finite.

6.8 Corrections for Macrostress Gradients


In Sect. 5.12 the effect of gradients in the direction of the surface normal was
discussed. Two types of stresses may contribute to such gradients: macrostresses
which have to act in the plane of the surface (0"11,0"12,0"22) and average micro

206

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress

stresses which can also act in the direction of the surface normal. It was also shown
that if steep stress gradients exist within the maximum penetration depth 't'P' "d"
vs. sin 2 \jJ would exhibit curvature. One way of minimizing the effects of gradients
is to use high 'l'-points in the analysis as discussed in Sect. 5.12. This method is valid
both for micro- and macroresidual stress gradients. An electropolishing method
has also been suggested to compensate for the variation of stress with depth [7].
This method, which is valid only for macros tress gradients, is discussed below.
For an x-ray beam penetrating a distance 't into the material, the average 29
measured will be a function of the variation of the true 29 with the penetration
depth:
J 29(x)dI
x=o
9
(6.104)
2 obsv. = - - 0 0 - - - dI = Ke -lI x /'dx
<X)

x=o

dI

which is similar to Eq. (5.42).


If a layer of thickness z is electro polished without changing the stress
distribution, this equation becomes
<X)

J29(x)dI

(29obsv .) z=z = -z-oo-::----

JdI

(6.105 )

which can be written as


(6.106 )
z

differentiating Eq. (6.1 06) with respect to z and simplifying:


29(z) = (29(z)obs.) -

:/1 .d(2:;bs.) .

(6.107 )

Thus, if one obtains the variation of the observed 29 angle as a function of depth
(by successive measurement and electropolishing) it is possible, theoretically, to
determine the true 29 as a function of depth. Of course, observed 29 data over all
the \jJ-tilts examined must be measured at every etching level to obtain the
necessary corrected data for stress analysis. The problem with this method is the
assumption that electropolishing does not alter the stress distribution. Some
relaxation will invariably occur. If a very thin layer is removed this relaxation may
be negligible. However, if successive electropolishing oflayers results in substantial material removal, relaxation effects must be taken into account.

6.9 Corrections for Layer Removal


Residual stresses form due to the mutual constraint of discrete volumes in a given
body which have suffered differential deformation. If one of these volumes is
preferentially removed, the residual stress distribution will change because of

6.9 Correction for Layer Removal

207

relaxation effects. In the limit, if one part of the body that is constraining the other
part is totally removed, the stress will decay to zero in both parts. Since with x-rays
one measures the stress after a volume is removed, the measured stress must be
corrected for the effects of such relaxation. The initial stresses in the body are
related to the relaxed stress state through the equations of compatibility and the
equations of equilibrium, and for a given specimen/stress distribution combination
it is possible to use these equations to determine corrections to the measured
values. Such treatment has been formalized in closed form for various geometrical
shapes [7,30,31]. These solutions involve evaluation of definite integrals of the
measured residual stress distribution. For example, in the case of a flat plate,
which contains a biaxial macroresidual stress tensor, the components of which
vary only as a function of depth, the correction is given by
er (z ) = er
x

Xobsv

(z)
1

(z)
+ 2 Her
S XobsvZ
ZI

Her
(z)
dz - 6z S Xobsv
dz
1

ZI

Z2

'

(6.108 )

where H is the original thickness of the plate, ZI is the distance from the lower
surface to the uncovered depth of interest, erx (ZI) is the true stress and erXobsv' is the
measured stress value. For complicated shapes and/or stress distributions, the
correction becomes more involved.
A series correction for layer removal is possible where the integrands in
Eq. (6.108) are expanded in the form of Taylor series referred to the surface
values and the integration is performed term by term:
1 )
H-Z
erx(d
z -erXmeas ( z d -- -4erXmeas (H) . ( H-

+ [er

Xmeas

H -_
z 1 ) 2 + ...
(H) +2H oerXmeas ( H ) ]. ( _
oz
H

(6.109 )

For small etching depths and shallow gradients, the first terms of the series may be
used:
Az
Aer=erx(zd -erXmea.(zd = -4erxmeas (H)'If'
(6.110)
This equation may be used to determine the polishing thickness that will yield a
small correction. For example, for Aer to be less than 5 %:
AZI

= 4'0.05 H.

(6.111)

Various other specimen geometry/stress distribution combinations have been


treated in the literature in closed form and by series analysis. All of these corrections, however, are based on macro or applied stress considerations. No work is yet
available on the effect of average microstress components on such corrections. In
case of critical applications, it may be better to use neutron stress analysis (if
possible) to determine the stress state in the deeper layers of the specimen, or to
verify the correction by using a combination of x-ray and neutron measurements
on polished and unpolished samples before a procedure is decided on.

208

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress

6.10 Summary
Residual stress values determined by diffraction methods have two types of error
associated with them:
a) statistical error, which arises due to the random arrival of x-ray photons at the
detector, causing an uncertainty in the absolute intensity at any 29 position;
b) instrumental error, which arises due to specimen alignment, effect of beam
optics, etc.
The variance in the peak position is the sum of the variance due to these effects
[26]:
0- 2 (

stress) = 0- 2 ( stat.)

+ 0-2 ( instr.) .

(6.112)

The statistical variance in peak position can be obtained from the variance in the
intensities over the pertinent 29 range. This is accomplished by propagating the
error through 'the equations used in peak location determination. The instrumental errors are frequently determined from geometrical considerations and are
dependent on the particular experimental configuration and beam optics.
Various strategies are available for peak determination. The most widely used
method is the fitting of a parabola to the top 15 % of the peak and then
determining the apex of this parabola. Various studies have shown that this
procedure can be used for the majority of peak shapes encountered in practice.
Furthermore, the calculation of the statistical error for this method is relatively
straightforward. These formulae have also been tested by comparing the standard
deviation calculated from one measurement with the standard deviation obtained
from multiple measurements, and excellent agreement has been reported [12, 14,
26].
If the equations for error determination are known for a given measurement
configuration, the stress measurement may be automated such that an experiment
can be carried out to an operator-specified precision. This may be accomplished by
writing a code that first requests the various experimental parameters from the
opera,tor, calculates the 0- 2 (instr.) from these parameters using the equations
given in this chapter, determines the maximum statistical error permissible for a
specified total error from Eq. (6.111) and then calculates the total accumulated
counts or time required for such an error. Intensity vs. 29 data is then obtained to
this specification. If the diffracted beam intensity is known (this may be obtained
from a fast preliminary scan of the peaks involved) , such a procedure can also be
used to estimate the time required for the stress measurement.

6.10 Summary

209

Problems
6.1. Given the following five points on a diffraction profIle.
Angle (29)

Corrected CPS.

Time

159.845
160.305
160.765
161.225
161.685

502.5
524.5
526.7
513.4
499.2

39.9
38.2
37.9
38.9
39.9

(i) Calculate the peak position by fitting a parabola to these points. (ii)
Calculate the statistical error. (iii) Calculate the "d" value (Co radiation is
used). (iv) Calculate the statistical error in "d".
6.2. Given the following "d" vs. sin 2 tp data.
sin 2 tp

d,p(A)

Statistical
error (d,p)'10- 4 (A)

0
0.1
0.2
0.3
0.4
0.5

0,90795
0.90735
0.90505
0.90669
0.90698
0.90607

0.569
0.432
0.568
0.390
0.367
0.398

(i) Calculate

0"<1>'

(ii) Calculate the statistical error in

0"<1>'

References
1
2
3
4
5
6
7
8
9

J.B. Cohen, H.Dolle, M.R. James, NBS Special Publication 567, 453 (1980)
A.J.C. Wilson, "Mathematical Theory of X-ray Diffractometry", Philips Technical Library
(1963)
W.E. Baucum and A.M. Ammons, Adv. in X-ray Anal., 17, 371 (1973)
T.J. Devine, Comparison of Full-Profile Peak Finding Methods For Use In X-ray Residual
Stress Analyses, M.S. Thesis, Northwestern University (1985)
J. Mignot and D. Rondot, J. Appl. Cryst., 9, 460 (1976)
M.M. Hall, V.G. Veeraaghavan, H. Rubin, and P.G. Winchell,J. Appl. Cryst., 10, 66( 1977)
Soc. Automotive Eng., Residual Stress Measurement by X-ray Diffraction, SAE J784a, 2nd
Ed. (1971)
H.M. Rietveld, J. Appl. Cryst., 2, 65 (1969)
R.E. Ogilvie, Stress Measurement with the X-ray Spectrometer, M.S. Thesis, MIT (1952)

210

10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31

6 Experimental Errors Associated with the X-ray Measurement of Residual Stress

J. Thomsen and F.Y. Yap, J. of Res. of Nat. Bureau of Standards, 72a, 187 (1968)
D.P. Koistinen and R.E. Marburger, Trans. ASM, 51, 537 (1959)
M.R. James and J.B. Cohen; Adv. in X-Ray Analysis, 20, 291 (1977)
R. Marion and J.B. Cohen, Adv. in X-ray Anal., 18,466 (1975)
M.R. James, An Examination of Experimental Techniques in X-ray Residual Stress Analysis,
Ph. D. Thesis, Northwestern University (1977)
M.A. Korhonen and V.K. Lindross, Scandinavian Journal of Metall., 2, 100 (1973)
M. Knuuttila, Computer Controlled Residual Stress Analysis and its Application to
Carburized Steel, Ph. D. Thesis, Linkoping Univ. of Tech., Linkoping, Sweden (1982)
H.K. TonshofT, E. Brinkmeister, and H.H. NOlke, Ztsrft. f. Metalkde, 72,349 (1981)
H.K. TonshofTand E. Brinkmeister, Annals ofCIRP, 29,519 (1980)
A.J.C. Wilson, Brit. J. Appl. Phys., 16, 665 (1965)
P.J. Rudnik and J.B. Cohen, Adv. in X-ray Anal., 29, 79 (1986)
K.A. Perry, Experimental Determination of X-ray Elastic Constants, M.S. Thesis, Northwestern University, (1982)
K.A. Perry, I.C. Noyan, P.J. Rudnik, and J.B. Cohen, Adv. in X-ray Anal., 27 (1983)
C. Jatczak and H. Zantopulos, Adv. in X-ray Anal., 14, 360 (1970)
R. Marion, X-ray Stress Analysis of Plastically Deformed Metals, Ph. D. Thesis,
Northwestern University (1973)
R. Marion, ibid. Appendix B
M.R. James and J.B. Cohen, O.N.R. Technical Report, No. 14, Northwestern University,
Mat. Sci. & Eng. Dept., Evanston, 11. February, 1977
Centre Tech. des Indus. Mech., Les Mem. Tech. du Cetim, 33 (1978)
E.R. Pike, J. Sci. Instr., 34 355 (1957)
J.B. Cohen, "Report on Tungsten Lattice Parameter Round Robin", X-ray Subcomm. of
SAE Iron and Steel Tech. Comm., Division 4, (1964)
M.G. Moore and W.P. Evans, SAE Trans., 66,340 (1958)
D.P. Koistinen, R.E. Marburger, M. Semchyshen, and W.P. Evans, SAE Information Rep.
TR -182, A.L. Christianson ed., SAE, New York (1960)

7 The Practical Use of X-ray Techniques

7.1 Introduction
In this chapter, we consider the various ways to make measurements of residual
stresses on real parts, either in an ordinary x-ray laboratory (Sect. 7.2) to which
such samples may be brought occasionally, in a factory where frequent inspection
is required, or in the field, for example at large construction sites, on oil-rigs or
pipelines or power stations, or for parts which are too large for a commercial
diffractometer. Tliere are several units now available for such situations, but this
equipment is new and can be expected to change frequently. Therefore we prefer
here to provide a list of ideal requirements for the software and hardware
(Sect. 7.3), against which the reader can examine what is available commercially,
emphasizing his own priorities in these lists. (Also, it is always advisable to contact
several users of such equipment, as well as the supplier.) We will then briefly
describe the available equipment at this writing (Sect. 7.4 ), and conclude with a
detailed examination of several examples of the use of stress measurements in such
situations (Sect. 7.5).

7.2 The Use of Ordinary Diffractometers


For small parts, or samples whose shape is not too complex or bulky, ordinary
commercial diffractometers are quite adequate and offer the added advantage that
such instruments can also be used for a variety of other measurements such as
retained austenite, or preferred orientation. In some cases it will be necessary to
add an attachment for the 'P or n tilt, although with some equipment this is already
available. If such an additional tilting device is needed, a suitable one can often be
obtained from manufacturers of single crystal devices; large goniometers exist that
provide the necessary motions. It is only necessary to check carefully that such a
device is pinned to the diffractometer so that all centers of rotation coincide to
0.001 cm. Parallel beam geometry can be arranged by merely rotating the soller
slits supplied with the unit by 90, so that the plates are perpendicular to the
diffractometer. One' of the distinct advantages of a commercial instrument is that
it is often available with a mini- or microcomputer control and will permit software
development for stress measurements. In some cases such software is available for
a unit but at this writing it is minimal. We turn next to the ideal requirements for an
instrument to be employed continually for stress determinations, and capable of
field use. (The software requirements are also applicable to commercial
diffractometers. )

212

7 The Practical Use of X-ray Techniques

7.3 Software and Hardware Requirements


A. Hardware

1) There should be a high level of safety, such as scatter shields. Also, the shutter
should remain closed unless the device is in contact with a part, and there
should be warning lights. In-field equipment may be operated by persons
with minimal skills in diffraction!
2) Some form of position-sensitive detector should be incorporated to minimize
measurement times.
3) The equipment should be rugged and transportable.
4) Protection from dust and moisture may be necessary.
5) To reduce weight and bulk, use should be made of miniature components,
such as the x-ray tube and generator.
6) There should be flexibility so that the device can be employed in a variety of
situations.
7) It should 'be possible to choose parallel or divergent beam geometries.
8) A microprocessor control and data processor is essential.
9) There should be a choice of (\jl or Q) rotations.
10) The <l> rotation should be possible, so that the entire three-dimensional tensor
can be measured if necessary.
11) Is there a means for oscillation to reduce the effects of coarse grain size? (This
is especially important with a PSD, as with such a detector seeing the whole
peak at one time, texture can distort the peak shape.)
12) It should be simple and rapid to change the x-ray tube.
13) It should be simple and rapid to choose different Bragg peaks.
14) It should be possible to use it at large distances from the electronics (8 -10m
or more).
15) Light weight is important for ease in handling and positioning.
16) Completely automated operation should be available (including tilts,
shutters et~.)
17) Suitable procedures and attachments are needed to measure the x-ray elastic
constants.
18) A graphics display (and final hard copy) of a peak during measurements is
essential, as well as "d" vs. sin 2 \jl.
19) Is maintenance available? Can parts be replaced in a matter of days?
20) Are there adequate yet rapid means of locating the unit relative to a part.
B. Software

1) An easily readable menu of all options and required input should be available
on the screen.
2) "Help" instructions should be provided.
3) Are appropriate corrections (f2, L.P. etc.) made to the data (including
detector nonlinearity, background)? For this non-linearity, how long does it
take to gather the data, and how often must it be repeated.
4) The output should be well organized and clear.

7.4 Available Instruments

5)
6)
7)
8)
9)

10)
11)
12)

213

Can data be stored for future use?


How is the peak determined, and with what number of points?
Are available x-ray elastic constants in storage in the program?
Is the program written in a language (such as FORTRAN) so that
modification by the user is feasible (which also implies that the source listing
is made available)?
Are the geometric and statistical errors evaluated and reported? Some care is
needed in examining this point. Some software may estimate the statistical
error from a least-squares fit of"d" vs. sin 2 tp, without taking into account the
errors for each point.
What types of stress tensors can be measured? .
Operation should be possible in a mode that permits specifying the desired
total precision in stress in advance. (The computer "decides" how long to
count at each tilt.)
How fast is the data processing immediately following a measurement (more
than a few seconds is excessive).

There is no available equipment that provides all of these requirements, so the


prospective buyer must decide which are most important to her or him.

7.4 Available Instruments


A. General Purpose Instruments

This class of field devices offers sufficient flexibility to permit at least some choice
of techniques.
1) Strainflex (RigakujUSA Inc., Danvers, MA. 01923)
This instrument can be described as essentially a full diffractometer with a
standard x-ray tube (and its associated large transformer) mounted on a
mobile hoist. An ordinary detector is employed. Calculations and operations
are "hard-wired" so that changes are not feasible. A similar instrument is
available from the Siemens Corp. and Shimadzu Co.
2) PARS (Portable Analyzer for Residual Stresses, Technology for Energy
Corp., Knoxville, TN 27922).
A position-sensitive gas filled proportional counter is employed, as well as a
miniature x-ray tube and a miniature high-voltage generator. It incorporates
micro-processor control and calculation. [This device was invented by M.
James and one of the authors (J.B. Cohen) J.

B. Single Exposure Instruments


While the machines described in this section are primarily intended for the single
exposure methods, in some cases somewhat more flexible operation is feasible.
While the original efforts with this technique used film to record peaks, with the

7 The Practical Use of X-ray Techniques

214

attendant delays for development and reading, these commercial products all
involve detectors.
1) F ASTRESS (American Analytical Co., Grafton, OH 44044 ).
Two standard x-ray tubes with different incident directions and two conventional detectors are employed. Peak position is located automatically when the
intensity reading is the same in both halves of each detector.
2) EMSAM, Paris, France
Two gas fIlled position sensitive detectors straddle both sides of a standard xray tube to record both sides of a diffraction cone in the back reflection region.
Control and data processing is through a microprocessor.
3) HITECH Canada, Ltd., Ottawa, Ontario, Canada KZC3N6
This is similar to item # 2.
4) Dr. C. Ruud, Materials Research Laboratory, The Pennsylvania State
UniversitY'eUniversity, Part, PA 16802
This is similar to both (2) and (3), but utilizes a miniature x-ray tube and
special miniature position sensitive detectors built with channel plate. The
instrument is sufficiently compact to be placed inside a large pipe. The detectors
produce analog signals rather than the counting pulses produced by gas-filled
detectors, so that calculating the statistical error is not straight-forward.
We turn now to several examples of in field measurements. Our aim in this
section is three-fold: 1) to demonstrate the actual use of the techniques, 2) to
show how such measurements can be employed to resolve important problems,
and 3) to illustrate typical output data.

7.5 Selected Applications of a Portable X-ray Residual Stress


Unit
(By W.P. Evans!)
A photograph of the x-ray stress unit manufactured at Caterpillar Tractor Co.'s
Research Department is shown in Fig. 7.1. The goniometer consists of an x-ray
tube, position-sensitive proportional counter, and track. The psi ("') tilts are
available at several fixed positions up to 45 degrees. The goniometer is held in
proper location by means of an alignment plate and magnetic "hold-downs".
Several alignment plates are available for various geometries. Specific positioning
is observed through a telescope whose focus is set for proper location. The
goniometer is supported with an articulated arm (dental x-ray support) which in
turn is mounted on an elevating tower for coarse adjustment from floor level to 1.9
m (6.2 ft.) above the floor. A minicomputer is programmed to utilize either the
two-tilt or multi-psi technique. Intensities are corrected for the Lorentzpolarization factor and absorption. An average option (taking consecutive points
1 Staff Engr. Tech. Center, CaterpiIIar Inc., Peoria, IL (retired).

7.5 Selected Applications of a Portable X-ray Residual Stress Unit

215

Fig. 7.1 a,b. Photographs of the portable x-ray unit built at Caterpillar Tractor Co.

in the peak 2,3,5 or9 at a time) is available in the software to help smooth the data.
This data is screened (using 80 pet. of peak maximum) and a least-squares fit
to a parabola is employed to determine peak position. Statistical and geometric
errors are included in the calculations. Correction for doublet overlap is also
included in the two-tilt technique.
In what follows, the two-psi tilt technique (0 and 45 degrees) was used unless
otherwise noted. Counting times of 40 - 60 seconds were usually employed and
data on peaks were collected in the multi-channel analyzer at gain settings of
(usually) 512. The averaging option of 3 was most often used. The detector was
covered with 0.004 mm of V foil and Cr radiation was employed. Channels were
calibrated with a Cr-Fe powder mix using literature values for the peaks positions.

Example 1: Measurements on Ring Assemblies


An assembly consisted of a heavy ring with a welded backing or wrapper. The ring
was made of SAE 1524 steel. An angular span of 1200 on the ID of the ring was
induction hardened to 3 mm minimum depth. The ring was welded into the
wrapper prior to further assembly. A key processing variable was the order of
hardening ahd weld,ing.
Residual stress measurements were made near the ID of the ring in the
circumferential direction. A template was made to fit the ring on which the
goniometer alignment plate was attached. (The articulated support arm and
elevating tower were not available at the time.) In this manner various angular
locations could be set quickly. The template (and goniometer) was rigidly held in
position during psi-tilts by C-clamps.
Results are shown in Table 7.1, for 30 degrees on each side of the center line of
the hardened area, where the order of hardening and welding are compared. The data shows that hardening done before welding minimizes tensile stress. Ifhardening is
done after welding, tensile residual stresses result due to constraint by the heavy
wrapper (i.e., the ring is not free to expand during induction heating of the ID). In
fact, delayed cracking was observed on some specimens if hardening was done last.

216

7 The Practical Use of X-ray Techniques

Table 7.1. Residual stresses in ring assemblies


Size

Order of
Hardening (H)
Welding (W)

Loc,
deg

H-W

-30

Measured

+30
W-H

-30
+30

H-W

-30
+30

W-H

+30
+30

W-H

Re.sidual Stress, MPa

-30
+30

-302.0
- 358.5
-271.7
- 351.6
-307.5
-3123
175.1
317.8
236.5
112.4

(127.6)"
( 38.6)
( 78.6)
( 38.6)
( ~6.2)
( 66.2)
( 50.3)
(136.5)
( 60.0)
( 66.2)

- 6.9 ( 57.9)
- 43.4 ( 42.1)
- 557.8 ( 65.5)
-644.0 ( 57.9)
816.3 ( 55.2)
538.5 ( 40.0)
160.6 ( 64.8)
201.3 ( 45.5)

Mean

-310.7 ( 44.1)"
- 323.8 ( 24.2)
246.5 (100.9)
174.4 ( 87.8)

- 25.2 ( 25.8)
- 600.9 ( 60.9)
677.4 (196.5)
181.0 ( 28.8)

- 48.3 ( 40.7)
- 28.3 ( 42.7)

Statistical and geometric counting error (geometric error due to beam divergence, IV axis
misplacement, and sample displacement) .
.. Std. Dev. of the Mean.

Example 2: Measurements on A Bolted Assemply of Plates


The plates are of SAE 15B36 hot rolled steel. They were cut to size, approximately
1732 x 482 x 35 mm, drilled and countersunk, and then hardened, tempered, and
roller straightened. Straightening was done while hot following the temper. This
was an intended "leveling" operation which introduces a controlled bending
followed by straightening, as processed out-of-flatness of 5.5 mm was not
uncommon.
Residual stress measurements were made near the bolt holes in the long
direction of the plate. This direction is normal to delayed cracking that
occasionally re~ulted after bolting. Surfaces were cleaned with a dilute
hydrochloric-nitric acid mixture prior to stress measurement. Stress measurements were made before and after bolting, as shown in Fig. 7.2.
Results are given in Table 7.2. The inside face was inaccessible after bolting. The
actual change due to bolting could be measured on the outside face. The stresses
on the inside face after bolting were estimated by inference, as indicated. Plate No.
3 was fairly straight, so out-of-flatness was intentionally introduced by inclusion

217

7.5 Selected Applications of a Portable X-ray Residual Stress Unit

Fig. 7.2. Set-up for measurement of


residual stresses in a plate assembly
prior to bolting

of a piece of 5 mm weld rod near the right end to simulate a "worst case" situation.
The data shows an average change ofl147.21MPa after bolting. This could result in
residual tensile stress on the inside face as shown in the table. To eliminate this
possibility, the ()Ut-of-flatness tolerance was reduced and delayed cracking was
subsequently minimized.
Table 7.2. Residual stresses in a bolted assembly of plates
Plate

Assembly No.

Outside face
middle
2

3**

Inside face
left
middle
right
Outside face
left
middle
right
Inside face
left
middle
right
Outside f~ce
left
middle
right"

Residual stress, MPa


Before bolting

After bolting *

Change due
to bolting

- 265.5 (76.5)

-357.8 (77.2)

- 92.4

-115.1 (24.1)
102.7 (55.2)
- 35.9 (51.0)

151.1
- - 64.9
34.5

179.3 (64.1 )
- 296.4 (40.0)
-168.2 (44.8)

- 86.9 (41.4)
-128.9 (50.3)
- 238.6 (45.4)

22.8 (71.0)
-319.2 (70.3)
- 206.2 (70.2)

- 5.6
-458.5
70.3

-263.4 (47.6)
-432.3 (77.2)
-463.3 (62.1)

- 280.6 (64.1)
- 293.0 (58.6)
-739.8 (78.6)
Ave. change

Presuming "change" is additive, but opposite in sign to outside face .


.. Intentional spacer of 5 mm at right end (using weld rod).

-266.2
-167.6
- 70.4

- 17.2
139.3
-276.5
1147.11 (97.6)

218

7 The Practical Use of X-ray Techniques

Example 3: Measurements in Depth on A Solid Circular Cylinder


This is an example oflayer removal technique to establish the residual stresses in
depth. The results were to be used as a reference in a later investigation.
The cylinder was 41.07 mm dia. x 146 mm in length. It was made of SAE 4116
steel. The austenization at 1550F (843C) in a neutral atmosphere was followed
by severe water quenching in a four pole submerged spray. Surface hardness was
Rc 46-47.
X-ray residual stress measurements were made on the cylindrical surface at
mid-length in both axial and circumferential directions after successively removed
layers. Layers were removed by grinding followed by electropolishing to remove
grinding stresses. A typical set-up is shown in Fig. 7.3 near the end of the test after
many layers had been removed. An alignment plate was used whose plane contains
the axis of focus of the goniometer. Mid-length measurements were taken at three
locations 120 deg. apart and averaged. The average standard deviation of the
means was about 30 to 56 MPa for 16 depths (96 measurements in all).
Corrections for "stressed layers removed were carried out by methods discussed in
Chap. 6. Corrected stresses are plotted in Fig. 7.4. The radial stresses and
corrections for layer removal followed [1]. The trends are typical of residual
stresses produced by thermal and transformation stresses. The hardened surface
layers are in compression with balancing tension in the inside. That conditions of
equilibrium are satisfied is shown by the force balance in Table 7.3. Note that each
component of stress sums to zero over the cross-section. Longitudinal (or axial)
forces are balanced over circular segments of width equal to the layer removed by
the relation
(7.1 )
where <O"z> is the average axial stress at radius R over the increment dR.
Tangential (or circumferential) forces are balanced over incremental rectangles
of unit length and width dR by the relation
.I:,< O"o>dR ,
where <0"0> is the average circumferential stress over dR.

(7.2)

Fig. 7.3. Experimental arrangement for measurement of residual


stresses remaining in a circular
bar after many layers had been
removed. Positioning is for circumferential direction

7.5 Selected Applications of a Portable X-ray Residual Stress Unit

r__ *_

600

RC~4~

/HARDNESS__

--"'--*~ ~

400

219

35

___ ~~:.

200

~ O/~/ )(/~
~

c91j

-200

~ -400
<
:::)

x
x/x

o - LONGITUDINAL
.=RADIAL

~ -600
-800~011
0

_1000X~~
-1200

x - TANGENTIAL

-1400~1__~~'~~~'~~~'__~~'__~~
' ~~~'~~!

0.0

2 .0

4.0

6.0

8.0

10.0

12.0

14.0

DEPTH, mm

Fig. 7.4. Residual stresses in a cylindrical bar due to furnace hardening


Table 7.3. Equilibrium check, round bar
Depth

Long forces
( x W"N)

Tang forces
( x 103 N)

Depth

Long forces
( x 10 3 N)

0.0508
0.1270
0.2692
0.5079
1.2699
1.4071
1.5240
2.7939
2.9971
4.9530

- 5.37
- 9.10
-16.09
-25.12,
-59.73
- 8,95
- 8.03
-64.33
- 7,03
-27.20

- 1.40
- 2.58
- 4.66
- 7.04
-17.13
- 2.45
- 2.21
-20.47
- 2.75
-17.76

5.0799
7.5183
7.6199
10.0329
12.5729
15.2399
17.7799
20.3199

0.29
30.28
2.08
54.50
58.45
45.81
26.31
9.70

Totals

-3,58

Tang forces
(xl0 3 N)
0.33
2.94
0.38
11.25
14.55
15.93
15.61
15.61
-

2.52

The force balance is an interesting exercise but can only be done if measurements
are made near the bar center, and only if microstresses are negligible. It is an
indication of reliability in the test, that as stressed layers are re~oved the bar
relaxes and obeys conditions of equilibrium (as it certainly must) , and that most
of the stresses are macrostresses.

220

7 The Practical Use of X-ray Techniques

Example 4: Measurements on Crankshaft Journals

The test crank was a steel forging ofSAE 1034 steel. It was normalized, machined,
and heat treated to give a hardness of Rc 48 minimum. The journals are finished
ground to size and polished to improve surface finish. The problem was to
quantify the effect of polishing on residual stresses in the journal bearing areas.
Measurements were made at mid-length of both pin and main journals as
shown in Fig. 7.5. The outboard side of the journal was used which was set
adjacent to the goniometer alignment plate. In this way, measurements could be
made non-destructively, without sectioning the journals. Results are shown in
Table 7.4, comparing as-ground to ground and polished journals. Significantly
higher compression is seen in the polished journals. Scatter is also somewhat less.

Fig.7.S. Set-up on a crackshaft journal to


determine residual stresses in the surface
layers after grinding or polishing. Positioning is for the longitudinal direction in the
journal

Table 7.4. Residual stresses in crankshaft journals comparing as-ground and ground and
polished journals
Residual stress
Measured
As-ground

Ground and Polished

Pin
Pin
Pin
Pin

- 504.0
-600.5
- 562.6
- 277.2
-431.6
- 251.0

(46.9)*
(40.7)
(42.1)
(53.8)
(60.7)
(49.9)

No.2 Pin
No.4 Pin

-934.9
- 955.6
-920.5
- 862.5
- 662.6
- 753.6
-819.1

(49.0)
(62.1)
(59.3)
(43.4)
(45.5)
(73.8)
(80.7)

No.2
No.4
No.6
No.5

No.6 Pin
No.1 Main
No.5 Main
.... Same as meaning in Table 7.1

MPa(ksi)
Mean

-555.7 (48.6)"
-319.9 (97.6)

-918.4 (39.9)
-745.1 (78.6)

221

7.5 Selected Applications of a Portable X-ray Residual Stress Unit


Table 7.5. Data bank display of crankshaft data
MFG
PNA

Engine PLT
Crankshaft

T
HAR
DEP
DIR
BEAM

400F
RC 52-58
SURF
AXIAL
1 DEG BEAM

Location

Stress, MPa

Standard Deviation,
MPa

FILLET, FRONT
FILLET, FRONT
FILLET, REAR
FILLET, REAR
MID-LGTH
FILLET, REAR
FILLET, FRONT
FILLET, FRONT
FILLET, REAR
MID-LGTH
FILLET, FRONT
MID-LGTH
FILLET, FRONT
FILLET, REAR
MID-LGTH
FILLET, FRONT
MID-LGTH
FILLET, FRONT
FILLET, REAR
MID-LGTH
MID-LGTH
FILLET, FRONT
MID-LGTH
FILLET, FRONT
FILLET, REAR
MID-LGTH
FILLET, FRONT
FILLET, REAR
MID-LGTH
FILLET, REAR
MID-LGTH
FILLET, FRONT
MID-LGTH

-275.8
-341.3
-377.8
-341.3
-548.8
-295.1
-299.9
-255.1
-233.0
-320.6
-302.7
-722.6
-396.4
-395.8
-490.2
-471.6
-571.6
-455.1
-330.3
-509.5
-604.0
-365.4
-334.4
-313.7
-295.1
-481.9
-454.4
-351.6
-485.4
-268.9
-348.2
-408.9
-364.7

36.5
46.9
37.9
63.4
46.9
42.1
44.8
41.4
37.2
44.8
38.6
45.5
50.3
49.0
57.2
38.6
60.0
31.7
35.2
43.4
52.4
42.1
53.8
35.2
37.2
55.2
34.5
35.2
48.3
38.6
42.1
42.1
40.0

LNO
Date
REQ
COM
TC
MAT
HT

30224
6-3-83
R. Shafer
Fillets were peened
Jour were sectioned
SAE 1552
IND HARD

SID
No 1 PIN
No 1 PIN
No 1 PIN
No 1 PIN
No 1 PIN
No 1 PIN
No 2 MAIN
No 2 PIN
No 2 PIN
No 2 PIN
No 3 MAIN
No 3 MAIN
No 3 PIN
No 3 PIN
No 3 PIN
No4 MAIN
No 4 MAIN
No 4 PIN
No 4 PIN
No 4 PIN
No 4 PIN
No 5 MAIN
No 5 MAIN
No 5 PIN
No 5 PIN
No 5 PIN
No 6 PIN
No 6 PIN
No 6 PIN
No 6 MAIN
No 6 MAIN
No 7 MAIN
No 7 MAIN

This is believed typical of polishing (or lapping) and is often used as a residual
stress leveler for this reason.
The data bank display of data for another sized crank shaft, material, and its
heat treatment is shown in Table 7.5. The material was an SAE 1552 steel forging.
Such data represents a base line and is of great value in quality control (QC) and
for comparing processing and manufacturing trends.

222

7 The Practical Use of X-ray Techniques

Example 5: Measurements on a Gear Tooth


In this example, measurements were made along the flank of a cracked test gear
tooth in an effort to delineate possible delamination at the case-core boundary.
The gear was of SAE 8822H alloy steel, carburized, quenched and tempered to
give a minimum Rc hardness of 59. The tooth had been overloaded in a test until
cracking had occurred. Ultrasonic tests had indicated possible delamination at a
point approximately 12 mm from an edge. It was suggested that residual stresses at
the surface over the separation should show a decreased value if separation had
occurred. If this was true, stress measurements might prove of value in finding
such cracks.
The experimental arrangement is shown in Fig. 7.6. Proper location was
achieved with the alignment plate. Measurements were made along a line
approximately 5 mm from the tip in the radial direction with respect to the gear

Fig. 7.6. Arrangement with a gear tooth to


determine surface residual stresses at selected points along the flank in the radial
direction
-10

""
-200
Q.
:Iii
u)
(/)

w
a:

!n -300
..oJ

""

::::I
Q

en

~ -400

-500
0.0

10.0

20.0 30.0 40.0 50.0


TEST LOCA TION-mm

60.0

70.0

Fig. 7.7. Measured residual


stresses at points along the flank
of the gear tooth. A marked
decrease in magnitude of the
stress is seen at a location approximately 5 -13 mm from one
edge

7.5 Selected Applications of a Portable X-ray Residual Stress Unit

223

Fig. 7.8. Photograph on an etched section of


the gear tooth showing a crack at approximate
location of the x-ray stress measurement. The
section was taken at 5.5 mm from the edge
referred to in Fig. 7.7

hub. The multi-psi tilt method was used with tilts of 15,30 and 45 degrees; the
zero-degree position was also used in a few cases. Counting times of 120 seconds
were used for each peak. The beam was 2 rom in diameter.
The results at various points along the flank are shown in Fig. 7.7. A significant
decrease in residual stress is shown in locations 5 -13 mm from the suspect edge.
Sectioning verified the findings as shown in Fig. 7.8. After the fact, measurements
closer to the tooth tip should have enhanced the effect, and a smaller x-ray beam
might have helped as well.
Example 6: Measurements on a Shot-Peened Reference
A known reference is of great value to any laboratory making routine residual
stress measurements. Such a sample can be run from time to time to examine
equipment stability, electronics, detector, and calibration. It is also useful in
evaluating particular running conditions.
We use a shot-peened reference for this purpose. It is a plate ofSAE 1045 hot
rolled steel with dimensions 140 x 28 x 6 mm. Hardness before peening was RB
92.5 (Rc 12). One face was shot-peened with 230 shot to an estimated Almen A
intensity of 0.25 - 0.30 mm.
Over the past several years, we have run many hundreds of measurements on
this specimen. A sampling of recent data is shown in Table 7.6. The average of29
measurements of surface residual stresses is - 438.08 (22.46) MPa. The effects of
some important miming conditions are shown in Table 7.7 and show good consistency over the ranges examined.
Example 7: Triaxial Residual Stresses
In situations where the surface has plastically flowed or when high stress gradients
exist close to the surface, triaxial equations and methods must be used, as

224

7 The Practical Use of X-ray Techniques

Table 7.6. Reference data semi-portable x-ray residual stress data file
LAB NO. 30632
MFG. PLTfDATE ID:TECH CENTER/JUNE 1980
SAMPLE IDENT: PEENED REF
MAT SPEC: SAE 1045 STEEL
H.T. SPEC/QUENCH: HOT ROLLED
HARDfTEMPfYlELD: RB 92.5 (RC 12)
TEST LOCfDEPTH: MID-LGTH/SURF
DATE

BEAM GEOMETRY

1-12-84
1-19-84
1-19-84
1-20-84
1-20-84
1-20-84
1-20-84
1-20-84
1-23-84
1-30-84
2-16-84
3-05-84
3-07-84
3-13-84
3-23-84
3-30-84
3-30-84
3-30-84
3-30-84
3-30-84
3-30-84
3-30-84
3-30-84
4-03-84
4-04-84
5-21-84
5-21-84
5-24-84
6-08-84

0.1" DlA
0.1" DIA
0.1" DIA
0.06" DlA
0.06" DlA
0.06" DlA
0.06" DlA
0.06" DlA
0.1" DIA
0.1" DlA
0.1" DlA
0.06" DlA
0.06" DlA
0.1" DlA
0.1" DlA
0.1" DlA
0.1" DlA
0.1" DIA
0.1" DIA
0.1" DlA
0.1" DlA
0.1" DlA
0.1" DlA
0.1" DlA
0.1" DlA
0.1" DIA
0.1" DlA
0.1" DlA
0.06" DlA

Requested
by:WPE
STRESS
DIRECTION:
A-B
COMMENTS:
ROUTINE
CHECK

STRESS
MPa (KSI)

-450.2
-403.3
-446.0
-447.4
-444.7
-408.8
-434.3
-433.6
-443.3
-418.5
-443.3
-457.8
-466.7
-448.1
-431.6
-461.2
-405.4
-467.4
-421.9
-448.8
-462.6
-454.3
-425.4
-448.8
-393.0
-410.9
-474.3
-398.5
-454.3
-438.08
.< - 63.54)

ERROR
MPa (KSI)

33.0
34.4
29.6
41.3
51.7
112.3
30.3
26.8
35.8
33.7
29.6
50.3
37.9
28.2
28.2
31.0
31.0
28.2
33.0
36.5
37.2
29.6
30.3
29.6
128.2
31.0
39.3
30.3
35.1
22.46
(3.26 )

discussed in Chap. 5. As was shown there, this is recognized by 1p-splitting, that


is, d vs. sin 2 1p is different for +1p and -1p; these plots mayor may not be linear.
In any case biaxial equations will give incomplete and quite erroneous results.
In this present example, flow was introduced by directionally shot-peening the
surace (i.e., a 30 deg. peen) . Comparisons were made with a 90 degree peen which is

7.5 Selected Applications of a Portable X-ray Residual Stress Unit

225

Table 7.7. Effect of selected run conditions


Sample identity:
Mat. Specification
Heat treatment:
Hardness:
Test location/depth:
Stress direction:
Count time:
Beam geometry:
Average:
Gain

Counting time

Beam geometry

Averaging

Gain

Screen
(with 0.06 in. dia. beam)

Peened reference plate


SAE 1045 steel
Hot rolled
RB 92.5 (RC 12)
Mid-lgth/surf
A-B
40 s/or as noted
1 Deg masked/or as noted
By 3/or as noted
512/or as noted
Condition

Residual stress MPa (KSI)

lOs
20s
40s
80s
160s

-498.4
-375.7
-450.9
-466.0
-413.6

(86.8)
(59.9)
(28.2)
(24.8)
(27.5)

0.1 in. dia.


0.08 in. dia.
0.06 in. dia.
0.045 in. dia.
0.03 in. dia.
4 deg. masked
2 deg. masked
1 deg. masked
0.4 deg. masked

-432.9
-424.0
-419.2
-464.9
-386.1
- 509.5
- 448.8
-418.8
-430.2

(33.0)
(44.8)
(40.6)
(44.8)
(87.5)
(40.6)
(34.4)
(35.8)
(38.6)

Av'g by 1
AV'g by 2
Av'g by 3
Av'g by 5
Av'g by 9
Av'g by 9
Av'g by 5
Av'g by 3
Av'g by 2
AV'g by 1

-460.5
-424.7
- 421.9
-404.7
-431.6
-461.2
-490.9
-457.1
-457.8
-461.9

(49.6)
(33.0)
(36.5)
(35.8)
(26.2)
(24.1)
(39.3)
(28.9)
(33.0)
(41.3)

256
512
1024
2048

-418.5
-404.0
-460.5
-478.4

(36.5)
(38.6)
(83.4)
(92.3)

0.7 in.
0.75
0.8
0.85
0.9

-433.6
- 434.3
-447.4
-444.7
-408.8

(26.8)
(30.3)
(41.3)
(51.7)
(112.3)

-440.9 (47.5)
(-64.0 (6.9

-437.1 (34.6)
( - 63.4 (5.0

-447.2 (25.6)
(-64.9 (3.7

-440.4 (34.9)
(63.9 (5.1

-443.8 (15.2)
( -62.9 (2.2))

226

7 The Practical Use of X-ray Techniques

Table 7.8. Tensor input"


Input and/or lattice graphics information
K=1 (",=0); N(K) =4 ('" tilts)
J

PO(K,J)
('" tilt)

Lattice Parameter (A)


UP(K,J)
UN (K,J)

EP(K,J)

(micro inches/inch)
(+",)
(-"')

( -"')

2
3
4

0
15
30
45

1.172026
1.172053
1.171510
1.170636

EN(K,J)

1.171727
1.171209
1.170140
1.169298

*
*
*
*

1521.6
1545.2
1081.0
334.4

1266.1
823.9
- 89.4
-808.9

1.171786
1.171172
1.170236
1.169330

*
*
*
*

1446.5
1327.0
735.8
136.8

1316.6
792.5
- 7.6
-781.7

1.171803
1.171457
1.170785
1.169606

*
*
*
*

1439.5
1095.1
413.3
-358.1

1331.3
1036.1
461.3
-545.5

K=2 (",=45); N(K) =4


1
2
3
4

0
15
30
45

1.171938
1.171798
1.171106
1.170405

K=3 (",=90); N(K) =4


1
2
3
4

0
15
30
45

do = 1.170245

1.171929
1.171526
1.170729
1.169826

" for SAE 52100 steel, 30 degrees peen

nondirectional. Plate specimens of approximate dimension 110 x 38 x 6 mm were


used at three ca~bon levels: SAE 1018, 1045, and 52100 steels. Heat treatment was
as follows:

Austenitizing, C
Quenchant,
Temper,oC
Hardness, Rc

1018

1045

52100

870
water
430
29

820
water
430
30

840
water
430
50

Basic data was obtained by the multi-psi tilt method. Lattice parameters were
determined as a function of both positive and negative psi-tilts. Psi-tilts of zero, 15,
30 and 45 degree were used. This is done for three <p tilts (zero, 45, and 90

7.5 Selected Applications of a Portable X-ray Residual Stress Unit

227

Table 7_9. Example of tensor results (same sample as Table 7.8)


Strain matrix ( x 10 6 )

E12=
83.0
E22= -2269.5
E32=
-5.0

E11 = -1 949.3
E21 =
83.0
E31 =
475.6

E13= 475.6
E23= -5.0
E33=1346.6

Stress tensor
(MPa)

SI 12=
15.4
SI 22 = - 688.0
SI32=
0.9

SIll = -628.5
SI 21 =
15.4
SI 31 =
88.3

SI13= 88.3
SI 23= - 0.9
SI 33= -16.6

Principal stresses
(MPa)

SI 1 =642.6
Vl (for SI 1)

SI 2= -636.4
V2 (for SI 2)

X -0.283
Y 0.958
Z 0.038

0.948
0.286
-0.134

SI 3= -4.1
V3 (for SI 3)
-0.140
-0.001
-0.990

Orthogonality check of eigenvectors


OR (1.2) =3.897221E-07
OR (1.3) =2.142041E-07
OR (2.3) =7.462222E-08
Solutions check (check of eigenvalues in characteristic equation)

Fl =2.343750E-02
F2 = 4. 765625E - 01
F3 = 1.044921 E - 01
D (0) = 1.170245

degrees). Phi is the angle between a reference line in the surface and the plane of
the goniometer. In effect, six plots of d vs. sin 2 ", were obtained.
Running conditions for all peaks were as follows:
2.5 mm dia. beam
30 seconds count time
averaged by 2's
512 gain

7 The Practical Use of X-ray Techniques

228
Table 7.10. Key triaxial results for three steels

Material

Peen direction
(degrees)

Tensor results
Stress tensor, MPa
Normal

Shear

SI11

SI22

SI33

SI 12 SI13 SI23

52100

90
30

-695
-629

-703
-688

-33
-17

24 -2
15 88

4
-1

1045

90
30

-627 -627
-572 -699

9
-12

9 -2
8 56

-3
-5

1018

90
30

-547 -561
-526 -612

-2
-7

-5
15

5
30

4
1

Principal stresses and directions


Unit vectors

Peen
Principal Stresses
direction
(degrees)
S1
S2
S3
52100

1045

1018

90

-723

-675

-33

30

-693

-636

-4

90

-635

-618

30

-700

-577

- 7

90

-563

-546

-2

30

-614

-525

- 5

V1

V2

V3

-0.65
0.76
-0.006
-0.28
0.96
0.04

0.76
0.65
-0.002
0.95
0.29
-0.13
0.71
0.70
0.004
0.99
0.07
-0.01
0.95
-0.30
-0.006
0.98
0.17
-0.06

0.002
-0.006
-0.99
-0.14
-0.001
-0.99
0.002
0.004
-0.99
-0.1
0.006
-0.99
-0.008
-0.008
-0.99
-0.06
-0.002
-0.99

0.70
0.71
0.001
-0.07
0.99
0.01
0.30
0.95
-0.01
-0.17
0.99
0.007

Sample computer printout of input used for the triaxial calculation is shown
in Table 7.8. Strains were calculated from the lattice parameters computed for
no stress. All terms are matrix arrays. Using methods discussed in Chap. 5, a
strain matrix is computed which also defines the stress tensor, principal stresses,
and directions. An example of the complete triaxial (tensor) results for one
of the steels is shown in Table 7.9 Psi-splitting for that steel is shown in Fig. 7.9.
Key results for the three steels are compared in Table 7.10. As is seen,

7.5 Selected Applications of a Portable X-ray Residual Stress Unit

229

1. 173
r.J)

~ 1. 172

:. ----0______...
~

!'17' .~ ~O
~~

t;jl . 170
~

~~

Fig. 7.9. Lattice parameter vs.


sin 2 1jl showing "psi-splitting".
Sample was 52100 steel peened
at 30 degrees. Data was taken for
both psi tilts at a phi tilt of zero
degrees (corresponding to the
direction of peening)

~ 1.169

i=

~ 1.168

...J
1.167L.---~--...J....._-""""----'-----'

0 .0

0.1

0.2
0 .3
SIN2 (PSI)

0.4

0.5

shear components of stress are introduced by directional working. The effect


increases with carbon content. Thus, as mentioned earlier, directional
working of the surface (or gradients) causes triaxial stresses. We have also
observed marked effects produced by grinding (not shown). In this case high
stress gradients close to the surface compound the effect.
It is interesting to note the incompleteness (and considerable error) if only
biaxial analysis is used; e.g., a 52100 steel, with a 30 peen at phi = 0, gave measured
stresses in the longitudinal direction (cr 11) of the plate:
0

- 462.6 (62.0) MPa (positive psi tilt)


- 760.4 (31.7) MPa (negative psi tilt)

Reference
1 M.G. Moore and W.P. Evans, "Mathematical Correction for Stress in Removed Layers in
X-ray Diffraction R;esidual Stress Analysis", SAE Trans, 66 (1958)

8 The Shape of Diffraction Peaks


X-ray Line Broadening

8.1 Introduction
During processing, the shape of a diffraction peak can change (as well as shift)
and this is called "X-ray Line Broadening". This broadening can be an important
signature. For example, as shown in Fig. 8.1 (from [1] ), for peened soft steels this
can be even more important in controlling fatigue life than the induced stress. For
a well-annealed sample, the peak shape depends on the size of the source, finite
divergences of the slits at the x-ray tube and the receiving slits, the range of wavelengths in the beam, and the fact that in a diffractometer a flat specimen is only
tangent to the focusing circle at one point. 1 As a soft sample is deformed, there are
local strains due to the strain fields from dislocations and dislocation arrays. The
material is broken into regions with slight tilts with respect to one another due to
these arrays, subgrain boundaries, etc. Constructive interference occurs only
within each such region, and the peak (from such a region) is larger in angular
extent the smaller is such a region. This is so because when the region is small there
are not as many planes to cause destructive interference away from the exact Bragg
angle. To see this more clearly we adopt a treatment by Cullity [2].
Suppose that the crystal has a thickness t measured in a direction perpendicular to a particular set of reflecting planes (Fig. 8.2 ). Let there be (n + 1 )
planes in this set. We call e p the angle which exactly satisfies the Bragg law for the
particular values of A and d involved, or
A=2dsinell

(8.1 )

In Fig. 8.2, rays A,F, ... ,K make exactly this angle with the reflecting planes. Ray F',
scattered by the first plane below the surface, is therefore one wavelength out of
phase with A', and ray K', scattered by the nth plane below the surface, is n
wavelengths out of phase with A'. At the diffraction angle 2ell , rays A',F', ... ,K' are
completely in phase and add to form a diffracted beam of maximum amplitude,
hence the beam of maximum intensity (as intensity is proportional to the square
of the amplitude).
When incident rays that make Bragg angles only slightly different from e p are
considered, destrllctive interference is not complete. Ray B, for example, makes a
slightly larger angle e 1 , such that ray L' from the nth plane below the surface is
At low scattering angles, all of these factors are important. However, at high angles, where
stresses are normally measured, the dominant effect in focusing geometry is the wavelength
distribution. Even if only a K" line is employed ~},p. . = 1/2000. By differentiation of Bragg's
law (with d constant), ~2e, the range of angles, is 2 (!!J,}. .j)...) tan e ~ 10- 3tan e. At 1562e this
~0.3.

8.1 Introduction

231
Surface macro residual stress
cancelled
Surface macro residual stress
not cancelled

40

en
c

'cOJ

30

.8

OJ

- 20
::J

"0

Benefit due to macro


residual stress

OJ

::J

.Q\

. 10
OJ
Ul

OJ

L..

ct.

0
20

30
40
50
Hardness before peening, Rc

60

Fig. 8.1. Percent increase in fatigue limit due to peening as a function of hardness, showing the
separate contributions of x-ray line broadening and stress; SAE 86B45 steel. (These percentages
were obtained by cancelling the stresses with applied loads.) From Ref. [1]
F

O----~~~~_7~------~

1 ____________

2 ____________

__________

~_

__________

3 __~------------~-----K
L
L/ K'
N

'

"

/./ / -"N'
Fig. 8.2. Effect of crystal dimension on
scattering

(n + 1) wavelengths out of phase with B', the ray from the surface plane. This
means that midway in the crystal there is a plane scattering a ray which is an
integer plus one-half wavelength out of phase with ray B' from the surface plane.
These rays cancel one another, and similarly for the other rays from other pairs of
planes throughout the crystal, so that rays scattered by the top half of the crystal
cancel those scattered by the bottom half. The intensity of the beam diffracted at
an angle 2e 1 is zero. It is also zero at an angle 2e2 where e2 is such that ray N' from
the nth plane below the surface is (n -1) wavelengths out of phase with ray C'
from the surface plane. We have therefore two limiting angles, 2e 1 and 2e2 , at

8 The Shape of Diffraction Peaks - X-ray Line Broadening

232
I

Measured peak

Peak from a coherent


region

29
Fig. 8.3. A diffraction peak is made up of many peaks from different subgrains. Their shape
depends on crystal size, their position on the local microstrain

which the diffracted intensity drops to zero. Thus, the diffracted intensity at angles
near 29 p (yet not greater than 29 1 or less than 29 2 ) has a value intermediate
between zero and the maximum intensity of the beam diffracted at an angle 29 p
The smaller the thickness, t, the wider is the range 29 1 - 29 2
We can think of the diffracted beam as a sum of many peaks. Each of these is
shifted due to the local strains in it, and each has a shape that depends on the size of
the region. This is shown in Fig. 8.3. The measured peak shape will depend on the
range oflocal strains and on the size distribution. Our goal in this chapter is to see
how to obtain such information from x-ray line broadening, and to compare and
contrast it to the strain obtained from peak shifts.
The shape of a diffraction peak from a monochromator on a neutron source or
synchrotron source can sometimes be adjusted to be a simple function, such as a
Gaussian curve. However, the shape of an emission line from a normal (sealed) xray tube defies such a simple desc~ption. Because of this, corrections for the shape
of a peak (takeI!: with a sealed tube) due to instrumental factors can be done only
very approximately by assuming such simple functions. Fofexample, if the shape
is assumed to be a Lorentzian function, the half-width of a peak from a well
annealed standard is merely subtracted from the broadened profile's half-width to
obtain a corrected profile's width. If a Gaussian function is assumed, the squares
are subtracted. Because it is well recognized that neither approach is generally
correct, an average of these two procedures is sometimes employed when
examining broadening. After this correction, with some further simplifying
procedures, we can get some qualitative information on the subgrain size and
microstrain. The size, D, affects peak broadening in the following manner:
A290=

A 180
Dcos9 1t

( 8.2)

By differentiating Bragg's law, the effects of strain are


A290= -!Ad tan 9 1!0 .

(8.3 )

8.2 Slit Corrections

233

Actually, from our discussion above, peak broadening is related to the standard

deviation, (( ~: )
~2eocose=

2)

1/2. Then, including the effect of subgrain size:

180[ D
A -2 ((~d)2)1/2
]
1t
d
sine.

(8.4 )

From a plot of ~2eo cos e vs sin e, D is obtained from the intercept and the rms
microstrain from the slope. Multiple orders of an hkl peak should be employed,
not just any peaks, as most materials are elastically anisotropic. The size is actually
an average over the volume of the x-ray beam.
Actually, there is a better way with fewer assumptions - using Fourier series.
Twenty years ago this was a bit tedious, but with small ( or large) computers this is
a relatively simple task. We will follow this approach in this chapter. If the reader is
unfamiliar with such series, Appendix C will provide the appropriate background.

8.2 Slit Corrections


We tum first to consider how we can correct a peak's shape for the various effects
that cause the peak from a sample with even few defects to have a finite breadth.
Suppose a slit in a measuring system can be described by a function f( y) 1, and the
true function to be measured is g (z) , Fig 8.4. The receiving slit takes any element
g (z) dz as it "sees" it, and spreads it out into a function of the form f( y). Some
point on this spread function is a contribution, d[h (x) ], to the measured function.
Therefore, we can write
d[h (x)]j (g(z) dz) =f( y) lAy,
where Ay is the area of f( y). Then,
d[h(x)] = (ljAy)g(z)f(y)dz.
As can be seen in tl!e figure, y = x - z. Hence,
1
h(x) = Ay

J g(z)f(x-z)dz.

+00

oo

(8.5 )

Our resulting function h (x) is proportional to what is known as the "convolution


integral" of the slit function and the true function. This measuring process is
actually a very good way to visualize convolutions in setting up any analytical
treatment. For example, suppose we wish to convolute two triangles f( Xl) and
f( x2 ) as in Fig. 8.5. The various stages of the convolution, with one triangle acting
as a slit and moving past the other, are shown in the figure, along with the required
integrals for part of the evaluation.
1 The function f( y) really represents all instrumental effects.

8 The Shape of Diffraction Peaks - X-ray Line Broadening

234
f(y) -slit

h(x)

9 (z) - true
function

measured function

Fig. 8.4. A slit f( y) redistributes


an element of a true function
g (z) dz, so that it has the shape
fey). (After Warren, B.E., "Xray Diffraction". AddisonWesley, Reading, MA, 1969.)

_~+~ u>O

flu)

={

!!+~.
2, 2

u< 0

I -3S x sol

IUI>3

u< x
u >x

/S/&,

1-6sxS-31

y(xl =

-3

f(x)~g(x)

x+3

J (~+~ )(-~+~+~)

J\Q+~)(Q+~-l!.)dU
2 2 2 2 2
-3

u 3
u 3 x
(-+-)(--+-+-)du
2 2
2 2 2

J f-~+~)(- ~.~.~)du
"3

du =?

-3

y(x)=f(x)~g(x)

x+3

f(u)

_L---X.._..L..---'-_"'---"o._u
-3
x
0

= ;>

Fig. 8.5. The convolution of two triangles

We now write the functions in the convolution Eq. (8.5) as Fourier series:
1
h(x) = Ay

= ~y

+00

Joo g(z)f(x-z)dz

):

L:Fnexp( -21tin;X-Z) )L:Gnex p ( -2:in'z )dZ

/
"Gn,exp (-21ti[n'-n]z) exp (-21tinx)dz .
-~~
A +S3 2 "F
~ n~
y -a/2

n'

(8.6 )

The limits on a convolution are really - 00 to + 00; by truncating the limits to


represent practical evaluations, we are effectively mUltiplying our convolution by
a function which is unity in the interval - a/2 to + a/2 and zero elsewhere. Then
the Fourier transform ofh (x) is its true transform convoluted with the transform
of the unit function.

8.2 Slit Corrections

235

{To see this we examine the Fourier transform of the convolution y (x) .
y (s) = Jy( x)exp( + 21tisx)dx

= Hf( u) g( x-u)exp( 21tisx )dxdu.

(8.7 )

Letting x-u=w,
Yes) =Hf(u)g(w)exp[ +21tis(u+w)Jdudw,
=Jf(u)exp( +21tisu)duJg(w)exp( +21tisw)dw
=F(s)G(s) .

(8.8 )

Following the same approach, the reader can readily prove that if
w(x) =f(x)g(x), the transform ofw(x), Tw(x), is
Tw(x) = W(s) =JF(s-u)G(u)du

(8.9)

Now G (u) can be considered as our true convolution, F (s - u) the transform


of our unit function. As the transform of the unit function exhibits oscillations [see
Appendix C, Eqs. (C.20a,b) J these will appear in the transform of h (x) when
h (x) is evaluated with finite limits. Because a transform and the determination of
Fourier coefficients are closely related processes, the Fourier coefficients of h (x)
will also exhibit such oscillations. Care is needed not to truncate the data too close
to a peak to minimize these oscillations.}
The integral in Eq. (8.6) involving n - n' is zero unless n = n', in which case it is
over a period, therefore,
a
( 21tinx )
( 21tinx )
hex) = Ay LFnGnexp - - a - =LHnexP - - a - .

(8.10 )

Thus, to obtain the coefficients of the true function g (x) ,


G n= (HJaFn)Ay.

(8.11 )

We need only have tbe Fourier coefficients for our broadened peak and those for a
peak from a sample that has few defects to obtan the coeffiCients of the true
diffraction peak without the effects of slits, source size, etc. To facilitate
computation, this equation is generally separated into sine and cosine terms (G~
and G~, respectively):
G~ +iG~ = [(H~ +iH~)/(F~ +iF~)] (Ay/a) .

Multiplying top and bottom of the right-hand side by


Gr (H~F~ + H~F~ ) Ay
n= (F~)2+ (F~)2

a'

Gi _
n-

H~F~-H~F~ ) Ay

(F~)2+ (F~)2

(8.12 )
F~ - iF~,

(8.13a)
(8.13b)

This entire procedure was first developed by Stokes [3J and his original paper is
still well worth reading! A program for carrying out the Stokes' correction is given
in [4].

8 The Shape of Diffraction Peaks - X-ray Line Broadening

236

If the measured functions are symmetrical, the sine terms of the series vanish
and we have
G~ = (H~ap") A,.

(8.13c)

In other words, if the slit and measured functions are symmetrical, so is the true
function.
With these coefficients G, we can synthesize the true peak. And if you wish to
have the true function in absolute terms, instead ofjust its shape, it is vital to keep
track of all the constants, a and Ay (An easy way to do this is to choose a function
for which the Fourier coefficients can be readily evaluated analytically, and with
the same period as the function to be evaluated" numerically. The numerical
solution can then be normalized to the analytical solution.)
It is possible to obtain the Fourier coefficients of any real function by replacing
the integral for these coefficients by a sum for practical evaluation on a computer: 1
1
(21rinx)
Gn=-Lg(x)exp
- - ax.

(8.14a)

For evaluation, let x/a be t/120 (the function is divided into 120 parts,
ax=a/120). Then:
a 1 1=+60
(21tint)
G n= 120 1=~60 g(t)exp 120

or:
(8.14b)
where the sum is over integers t in the interval a. Now the true peak g (t) = g (x)
can be written as:
(8.15 )
Therefore,
1
G n= 120

+60

1=~60 ~Gn'

(21ti[n-n /]t)
exp
120

(21ti[n-n /]t)]
1
[
= 120 ~Gn' ~exp
120
.

(8.16 )

The exponential term is zero unless (n-n') /120 is an integer. Thus, G n is the sum
ofG n, values for '(n-n') equal to an integral multiple of 120. That is, n-n' =0,
120,240, etc. Hence n'=n, n'=n-120, n' =n+ 120, etc. If the coefficients fall to
nearly zero values as n increases in the range of n from - 60 to + 60, there will be
1 Alternately, a least-squares fit to the series can be employed; uncertain regions of the peak can
then be ignored. [Kidron, A., and DeAngelis, R.J., "Symposium on Computer Aided
Engineering" (G.M.L. Gladwell, ed.), p. 285. Univ. of Waterloo Press (1971).J

8.2 Slit Corrections

237

only one coefficient in this sum for each n; this then is the required test that our
numerical analysis is correct. Furthermore, only as many independent coefficients
as there are data points can be obtained, in this case, 120. Additional coefficients
are related to these by the periodicity. The function is only defined by those
coefficients at the discrete data points and all 120 coefficients are, of course,
required to properly reproduce the function.
It often happens that it is not convenient to use the same period for the "slit"
function and the measured function because the latter is so much broader; if we do
use the same period with the sharper function, we will have many intervals (t)
with zero height. A Fourier series representing a diffraction peak, say a 001' peak
for an orthorhombic cell (we can always choose such a cell) , is in terms of a period
which is the inverse of the fictious cell parameter a~ 1:
21tnx

- - = 21tnxa 3

(8.17 )

In general, we can"compare coefficients at equal values of L where L = na~. The


fictitious cell edge, a~, is calculated for a 001 peak from the high - and low-angle
positions of the peak, where the tails join the background, eH and eL , rather than
from the distance between orders, i.e.,

[( l' + 1/2) - (1' -1/2) ]/a~ = (2 sin eHl"-)

(2 sin ed"-)

(8.18 )

In this way the large regions between two orders of a peak where there are no data
are not included in the analysis.
The sum involving the Fourier coefficients will produce peaks with period
(l/a~) even though all of these but the first really do not exist. However, it is only
the actual peak that concerns us in this analysis and we may ignore the others.
Thus we see how to actually obtain the Fourier coefficients of our true peak
without the effects of slits and other broadening factors. Our "slit" function is
simply the same peak recorded from a sample that is well annealed. If we wish to
do so, we can then ~ynthesize a diffraction peak from a sample with distortions,
without these "slit" effects, employing the Stokes-corrected Fourier coefficients.
Then the effects of various treatments of the specimen could be compared by, say,
comparing the breadth of the peaks. This is really unnecessary, however, for we
shall see in the next section that we can learn a great deal about the specimen from
the corrected Fourier coefficients themselves! Note below that it will only be
necessary to obtain the coefficients and normalize them in this procedure; it is not
necessary to keep track of constants.
In this Fourier analysis, it is a good idea to record at least four to five times the
breadth of a peak on either side of it, and to compare the background of a
standard and the pattern to be analyzed; the background in both should be the
same and this comparison will help in any extrapolation of overlapping peaks on
the broadened pattern. These precautions will minimize oscillations in the
coefficients. The peaks to be analyzed should be at least 20 % broader than the
1 The series is in terms of a reciprocal space variable b~ = lja~. That is
21tnxja = 21tnxjb~ = 21tnxa~. lfthe reader is not familiar with reciprocal space he should consult
any text on x-ray diffraction.

8 The Shape of Diffraction Peaks - X-ray Line Broadening

238

standard peaks, or else the analytical procedures will result il! considerable scatter.
(This means, e.g., that a particle siZe larger than about 1000 A cannot be measured
this way, because of the instrumental broadening.) If the broadening is less than
this, there are special techniques to eliminate the K"'2 peak [5] that enable sizes as
large as 2000 A to be studied. The errors in the particle size and in the strains are
generally about 10 % via this procedure. We turn now to see how these sizes and
strains can be obtained from the Fourier coefficients.

8.3 Fourier Analysis of Peak Broadening


We shall follow the treatment by Warren and co-workers [6]. The derivation that
we give again involves reciprocal space and the kinematic theory of diffraction. If
the reader is unfamiliar with this he should consult [6], or skip to the result,
Eq. (8.31). The material will be considered to be cubic, with a cell edge, ao. The
intensity I can be written as

The subscript "eu" means that we have left out such terms as the polarization
factor. We will bring this back later. The term s is the diffraction vector
(s=htb t +h2b2+h3b3 where the b i are reciprocal lattice vectors, and the hi are
continuous variables) and Cn is the position of a unit cell in terms of the axial
system: cn=nta t +n2a2+n3a3+Acn. The n i are integers. The term ACn is the
displacement of the nth cell. Carrying out the dot products:
Ieu(s) =

L
n

LF 2exp{21ti[h t (nt-m t ) +h 2 (n 2 -m 2 ) +h3(n3- m 3)


m

(8.19)
We will henceforth assume that each and every cell has the same structure factor,

F.

We measure power (P) not intensity, that is, the intensity integrated over the
number of grains (n) oriented to diffract and the area (A) of the receiving
surface (Fig. 8.6) .
P=HIdndA.

( 8.20)

Now:

( 8.21 )
where m hkl is the multiplicity. The angle 90 - e is that between the plane normal
and the incident beam. The equation for dn is the fraction of the surface of a sphere
of radius R covered by plane normals in the angular range dcx (assuming a random
distribution) .

8.3 Fourier Analysis of Peak Broadening

239

Fig. 8.6. The range of incident x-rays is the angle


dC(. The angles d~ and dy define the position of a
diffracted beam on the detecting plane

da : d(6 S Ja

Fig. 8.7. The three angles dC(,d~,dy define a volume in


reciprocal space. Note that dy and dC( are perpendicular to So and s so that the included angle is 2e

Also, dA=R2d~dy, so that:


P = III In~hkl cos 9 dcxR 2 d~dy .

(8.22 )

Now, the angles dcx,d~,dy also describe a volume in reciprocal space, Fig. 8.7, so
that:
d V = dcxd~dy sin 29 .
(8.23 )
Also

d V = Adh l b l

A,ph 2b2 X Adh3b3 = A, 3dh l dh~dh3


Vunit ceUSlD 29

Vunit ceU2 sin 9 cos 9 .

( 8 2. 4a)
(8.24b)

Thus, the variable's dcxd~dy can be replaced by dh 1dh 2 dh 3, by equating


Eqs. (8.23) and (8.24b). Then:
(P)eu=

mnR 2A,3
4V
c

HI SlD.eu9dhldh2dh3

( 8.25a)

and:
(P') eu = Peu/ (21tR sin 29) ,

(8.25b)

where P~u is the power per unit length of diffracting cone; a powder sample gives a
cone of diffraction with semiapex angle 29, but with a slit, say of unit height, we
measure only a piece of this cone, 1/21tR sin 29. Also, hence forth let M ( = mn)
represent the number of domains or regions of diffraction.

240

8 The Shape of Diffraction Peaks - X-ray Line Broadening

We can always defme a unit cell, such that the hkl diffraction peak is a 001' peak
from this new cell, as mentioned in Sect. S.2. Now, the distance from the origin in
reciprocal space to the diffracting point is the same regardless of the coordinate
system chosen, so that
(S.26 )
We continue with this new cell in order to simplify the mathematics; the variable is
now only h~, which is along the normal to the diffracting planes. Since
h~ b~ = 2 sin 6/A., then dh~ = ( cos 6 d26 ) / (b~ ) A.. Hence, replacing the integration
with respect to h~ with this expression and dropping the integration over 26,
(because we are measuring the shape of a peak vs 26):

x exp{21ti[h'dnl -md +h~ (n2 -m 2 ) +h~ (n3 -m 3 )

(S.27)

+s (Arn-Arm)] }dh~dh~.

The factors sin 26 and F2 are removed from the integral because they vary slowly
compared to the exponential terms. The preintegral factor will be written as
K (6) 1. The integration is a projection along h~ and h~. Consider only one peak,
say the 001'. Then the limits of the integrals with respect to h'l and h~ are from -1/2
to +1/2.
Also:
Afn=xna'l +yna~+zna~,

(S.2Sa)

s~l'b~

(S.2Sb)

and
(S.2Sc)

s (Arn-Arm) =1'(zn-zm) .
Therefore, performing the integrals, and with these substitutions:
(P20)eu =

K(6)L L L L L L sin1t(nl -md x sin1t(n2 -m 2)


nln2n3mlm2m31t(nl-ml)

1t(n2-m 2 )

x exp [21til' (zn -zm)] exp[21tih~ (n3 -m 3)].

( S.29)

The sine terms are each zero unless m 1 = n 1 ,m 2 = n 2 in which case they are unity, so
that the sum over m 1 m 2,n 1 n 2 becomes N h, N h2 , where the N's are the average
number of columns in a plane perpendicular to the [001'] direction in a mosaic
regio~.. The remaining terms
are the sums between pairs of cells,

(L L)
m383

1 K (6) is an angular factor, and we can include in it the polarization we ignored earlier, so
1 +COS226)
that for flltered radiation the angular dependent portion is F2 ( sin 2 6
.

8.3 Fourier Analysis of Peak Broadening

241

n3 - m 3, apart in a column. Let n = m3 - m~ and N n be the average number of cells


with an nth neighbor along the direction [001'], the average being over all columns
in the mosaic regions under the beam. Take the average value of the term involving
z over all these pairs of cells n cells apart, in all the columns, in all the mos.aic
regions:
(P~8) eu = K ( 9 ) N~, N~2

L N n<exp (21til'Zu) )exp ( 21tinh~) .

+00

-00

Multiplying and dividing by N~3' the average number of cells in a column,


expressing the exponents in trigonometric form, and keeping in mind that cross
terms for (n) and ( - n) cancel:

(P~8)

eu

= K (9) Nj,)~j,2Nj,3 {
. cos21tnh~ -

oo

-00

N~,nh3 <cos 21tl'zn )

~n

(8.30)
< sin21tl'zn ) sin21tnh~ } .
h3
Equation (8.30) is in a form suitable for one dimensional Fourier analysis, as it is
a Fourier series with coefficients, A..,Bn. That is, we record power vs. 29 (which
can be converted to h~) and we have an equation which has the same variable.
Now:
A.. = (N.JN~3) <cos 21tI'Zu) ,

(8.31a)

Bn = - (N.JN~3) <sin 21tI'Zu) .

(8.31b)

The coefficients An,Bn are those due to sample only, without instrumental
broadening; it is assumed that the Stokes' correction has already been applied.
Let us look at the various terms in this equation. Those involving displacement
depend on 1', or order, as we predicted in our qualitative considerations, while N n,
the particle size term, does not depend on 1'. The peak will be asymmetric if there is
a net displacement, i.e., if the sine term involving Zn does not average to zero, the
peak will be asymmetric toward lower angles for a positive net qisplacement, and
in the opposite direction for a negative net Zn' If there are many sine terms, i.e., if
the strain exists over a large distance, the peak will actually shift. Note also that:
Bn = -

n-+O

(~nh3 ),21tl<Zn),

(8.32)

and since the strain over a length of n cells, en' is:


Zua3 Zn
e =--, = n na 3
n

(8.33a)

or:
(8.33b)
so:
(8.34 )

8 The Shape of Diffraction Peaks - X-ray Line Broadening

242

Therefore, information about the average strains could be obtained from the
initial slope of Bn vs n. However this is certainly a much more cumbersome
approach than measuring a peak shift!
Because the part of the Fourier coefficients due to particle size does not depend
on l' while that due to distortions does, these parts can be separated. Several orders
of a peak, e.g., hkl = 200,400,600, are analyzed and a plot is made of An versus l' for
each n and extrapolated to l' =0. The ordinate is N.JN~3' Unless the material is
elastically isotropic, the displacements will vary with crystallographic direction;
thus we cannot mix peaks representing different directions. However, with a
powder or a polycrystalline material, the third order often overlaps with another
peak. For example, with an fcc material the 600 peak-occurs at the same position as
the 442. We can eliminate the need for a third-order term if the strains are small, so
that the term cos (21t1' zn) in An can be expanded (cos x = 1 - x2/2 + x4 /24 + ... ).
Including only the first two terms, and taking the logarithm of the expansion:
In ( (cos 21tI'Zo> ) == -

21t2

(1') 2(Z~> .

(8.35 )

With Eqs. (8.26) and (8.33a), Eq. (8.35):


InAn=In (N.JN~J - 21t2[h~ (a~) 2/a~]]n2(E~> ,

(8.36a)

or:
(8.36b)
where L = na~ is the true distance between cells in a column normal to the
diffracting planes. If there are no strains, only small particles, InAL versus h~ (for
each L) will be a horizontal line. If there is no particle size broadening but only
strains, all the lines for various L intersect at L = 0 at InAL = O. Note also that we
obtain (El//2 from the cosine coefficients, not the average strain.
According to Eq. (8.36b) only two orders of a peak are needed to separate the
effects of particle size and strain and obtain N.JN~3 and (El)1/2 vs. L. This is
because InAL vs. h~ is a straight line according to this equation (with a slope
yielding (El> ); it is possible to do the analysis with a powder or polycrystalline
specimen. However, Eq. (8.36b) resulted when we neglected higher terms than the
second in the expansion of the cosine. How valid is this? The term a~/ao is typically
5, (El)1/2 is at most 0.005. Substituting these values in the expansion of
cos 21tl'zn = cos 21t (hoa~/ao) nED' it will be found that nho can be 6 with the third
term in the expansion still only 6% of the second. Thus for the first few harmonics,
n, InAn versus h~ is, in fact, linear. 1
We have seen that if there are net strains over large distances, the peak will
shift. On the other hand, ifin the analysis the center of the actual peak is used, the
measured strain is the value above or below the long range (or mean) strain:
(El>meas = ( ( EL - Elong range) 2> = (El> - 2( ELElong range> + E210ng range'
(El>meaS=(El>-E2Iong range'
1 AL itself, as welI as InAL is approximately linear with h~: AL~Nn;Nh; [1-21t2h~U/a/)
(r.t> )], Delhez, R., and Mittemeijer, E.J., J. Appl. CrystalIogr. 9, 233 ( 1976), and this equation
may be used instead of Eq. (8.36b).

243

8.3 Fourier Analysis of Peak Broadening

En

Fig. 8.8 a,b. The probability of a given microstrain, e", is p (En)

because
(8.37 )
The average <EL> is the long-range strain. In an analysis with the center of the
actual peak, we are therefore measuring the deviation from the mean or longrange strain, not the average strain. There have been attempts to use the values,
<E~>1/2 to calculate the stored energy due to strains. However, it is not the average
strain that we measure, but rather the width of the strain distribution <El>1/2at
each L or n, Fig. 8.8. The average strain could be zero, as shown, but there would
still be a range of strains.
Some typical data are presented in Fig. 8.9.
It is easy to show that NJN~3 ~ (1 = n;N~3) for small L. Take a column of,
say, five cells and evaluate N n. (The terms in the sum involving - n include counts
in the other direction. ) For this column, No = 5, N 1 = 4, N 2 = 3, N 3 = 2, N4 = 1,
N5=O. The term Nl;N~3 is 4/5, which is the same as (1-1/5). This implies that
Nn/N~3 decreases r~pidly with L if N\ is small; the peak will be broader for
smaller regions, as we anticipated. If the values of Nn/N~3 obtained from the
intersections of the curves with the ordinate in Fig. 8.9 [see Eqs. (8.36a,b)] are
plotted vs. L, the initial slope is:
-,
d(NnINh.}/dnln_o = -

N'

h3

(8038a)
(8038b)

Thus, if a plot of N.JN~3 versus L is normalized so that its intercept at L=O is


unity, the slope gives the average value (Derr) of the length of the columns normal
to the diffraction planes - a measure of the effective mosaic size in that direction.
Such a plot is shown in the insert of Fig. 8.9b. (The data in Fig. 8.9b were obtained
after the Stokes correction.)
Often there is a small bend or hook in the data near n = O. This occurs because
with a broad peak it is difficult to estimate background, and hence, as Ao is the

244

8 The Shape of Diffraction Peaks - X-ray Line Broadening

10
v.

z09

...w~ 08
8 0 .7

FROM ANNEALED SPEClloI N

O .OL---~----~-----L-----L----~----~----~--~~--~--J

20

40

60

eo

100

120

140

160

180

(200)

(400)
16

L: 201

0.2

L : 40A
L ~ 60A

0. 3

1.0

N n 0.8

0 .5

0 .6

= BOA

= 100A

N"30.6

L = 120A

0.4

L = 140 1.

0.2
0.0 '---:4;;';0:;---;;8';:;0--:1*20
;:;-~16"::0:-:::*:::~~<;;:'

- 0 .7

L : 160A

UA)_
~--------------------~----------------------~

Fig. 8.9. a Fourier cosine coefficients of 400 peaks, from Ag electrodeposited from
(KCN + AgCN) solution, and from a well-annealed standard. b Separation of particle size and
strain and (insert) the determination of particle size, using the corrected Fourier cosine
coefficients, AL from a. [From R.W. Hinton, LH. Schwartz, and J.B. Cohen, J. Electrochem. Soc.
110,103, (1963).]

8.3 Fourier Analysis of Peak Broadening

245

area of the peak (see Appendix C), its value will be too small. The values from
beyond n = 2 - 4 should be extrapolated to n = 0 in doing this normalization. As
In (NJN~3) versus L is generally linear for n> 2, such a plot is a simple procedure
for this extrapolation. This has been justified theoretically, on the basis that the
strains are due to dislocations [7].
With a minimum of assumptions, we are able with this analysis to obtain the
rms strain (f.t)1/2 as a function of column length in a given crystallographic
direction and the mosaic size (Derr) in this direction. From this information we
can proceed to learn about the degree of anisotropy (from the size and rms strain
in different directions) and the imperfections. In a deformed material, (t/Derr) 2
can be taken as a dislocation density; that is, Derr is the spacing of such
imperfections. From the formulas for strain around a dislocation, the data on
(f.t)1/2 can also be used to obtain a density [8]. These two should agree if the
dislocations are randomly arranged, but if (t/Derr) 2 is smaller than the value
obtained from the strains, the dislocations may be clustered. The densities
calculated in thi~ way are in agreement with those observed in the electron
microscope [8]. (If there are cells or clusters of dislocations, Derr is a measure of
the cell size not the dislocation spacing, but this can generally be ascertained by
comparing the densities obtained with (f.2) and D as mentioned.) This technique
is especially useful for high concentrations of dislocations as this is exactly where
the electron microscope is least useful because of overlap of images. By combining
the two tools, a whole range of concentrations can be studied, with a region of
overlap sufficient to allow comparison of the two,
With materials like polymers, there is often only a single peak. Techniques for
obtaining the mosaic size and (f.~) for a single peak have recently been described
[9]. Finally, equation for the errors in D and (f.~> based on the counting statistics
have been developed [to]. Just as for the residual stresses, software can be
developed to obtain these quantities to an operator-specified precision [10].
Finally, composition gradients can cause peak broadening, particularly peak
asymmetry, due to the range of interplanar spacings produced by the gradient.
Thus, the sine coefficients can be employed to obtain information on the gradient
after correcting these for the strain and particle-size effects with information from
the cosine coefficients; see [11].
We have now examined all the signatures of residual stresses and strains on the
diffraction pattern.>We hope the reader feels equipped and ready to start his own
measurements!

Problem

(Fig. 8.1 0 )

Notes:
1) Given profiles are symmetrical (for simplicity).
2) The period (from 29.4-30.6 deg.9) is the same for both.
3) The number of intervals ainstrumental and abroadened are different. (Usually
ainstrumental should be a multiple of (abroadened) 12 to give Fourier coefficients at
the same L values.)

246

8 The Shape of Diffraction Peaks - X-ray Line Broadening


10
Broadened profile
(6 intervals )

-3

29.L.

81
~------------ a 3 ------------~

10
Instrument profile
(12 intervals)

0
- L.

-6

L.

29.L.

30,0

81

x-

8-

1------------------- a 3 ------------------~
~
Fig. 8.10

4) The period is divided into intervals of equal L\9 (for 9> 60, equal intervals of
L\ sin 9 have to be used, i.e., intervals of h~ ) .
5) Origins, 90 , are at centers of gravity of peaks. The 90 's need not be the same.
6) Background is assumed to be zero.
7) Wavelength is 1.5406 A.
The Fourier cosine coefficients, An, are given by

An= -1 a/2
L
a

x= -a/2

(21tnX )

f(x) cos-- 'L\x,


a

where a = number of intervals,


f( x) = value of intensity at x
x = interval number ( - a/2 ...0... + a/2 )
L\x = 1
n = harmonic number.

8.3 Fourier Analysis of Peak Broadening

247

a) Evaluate A o,A 1 ,A 2 for both peaks; normalize these to Ao = 1.


b) Evaluate a~.
c) Perform the Stokes correction for these coefficients. (This can all be done on a
hand calculator.)

References

2
3
4
5
6
7
8
9
10
11

W.P. Evans, R.E. Ricklefs and J.F. Millan in "Local Atomic Arrangements Studied by X-ray
Diffraction", (eds: J.B. Cohen & J.E. Hilliard) p. 351, Gordon and Breach, New York
(1966 )
B.D. Cullity, "Elements of X-ray Diffraction", 2nd Edition, Chap.3, Addison Wesley,
Menlo Park, California (1978)
A.R. Stokes, Proc. Royal Soc. London 61, 382 (1948)
R.I. DeAngelis in "Local Atomic Arrangements Studied by X-ray Diffraction" (eds: J.B.
Cohen & J.E. Hilliard), p. 271, Gordon and Breach, New York (1966)
R.J. De Angelis, ty1etallography, 6, 243 (1973)
B.E. Warren, "X-ray Diffraction", Chap. 13, Addison-Wesley, Reading, MA (1969)
R.L. Rothman and lB. Cohen, J. AppL Phys. 42, 971 (1971)
D.E. Mikkola and J.B. Cohen in "Local Atomic Arrangements Studied by X-ray
Diffraction" (eds: J.B. Cohen & J.E. Hilliard) p. 289, Gordon and Breach, New York
(1966)
R.K. Nandi, H.K. Kuo, W. Schlosberg, G. Wissler, J.B. Cohen, and B. Crist, Jr., J. AppL
Cryst. 17, 22 (1984)
W.H. Schlosberg and J.B. Cohen, J. AppL Cryst. 16, 304 (1983)
R. Delhez and E.M. Mittemeijer, J. AppL Phys. 49, 3875 (1978)

Appendix A: Solutions to Problems

Chapter 2
1. a) From the formula
cr'ij = aikajlcrkl, where cr;j is defined in L ,
( cr~3) = a3ka31crkl
=a~lcr11 + 2cr 12a 31a 32 +2cr13a31a33 +2cr23a32a33 +a~2cr22 +a~3cr33
The direction cosine matrix is given by

a ij =

035 0.61
-0.87 O. 5
0.35 0.61

-0.71

0.71

Thus, cr~3= - 100.0MPa


cr~1 = - 150.0MPa
cr~2 = - 450.0 MPa

cr~2 =
92.0MPa
cr~3= - 225.0MPa
cr~3 =
31.0MPa.

b) For an isotropic material,


l+v
v
Eij= E crij - "Ecrkk.
For Fe (from Table 2.1): E=211.4 GPa, v=0.293, (Ell )s=2.0910- 3,
(E 11 h=3.73910- 4 and other terms can be solved in a similar way.
c) From Eqs. ,( 2.28) and (2.29):
(cr;;)3+ 700 ( cr;i) 2 + 19100( cr;i) -4.036.10 7 =0.
Solving for (cr;i): cr~ 1= 201.28 MPa
cr 22 = -400 MPa
cr 33 = 501.28 MPa.
2. Eij = Sijklcrkl .
Since crkl=O for all k,l except k=l=l,
Eij = Sijll cr 11
Ell =Sl111cr11 =Sl1cr11 =3.2.10- 3
E22 =S2211o.11 =S12crll = 1.12.10- 3
E33=S3311cr11 =Sl3crll = 1.12.10- 2
E13=S13llcr11 =0 }
E23=S2311cr11 =0
E12 =S12llcr11 =0

for cubic symmetry.

Appendix A: Solutions to Problems

249

Chapter 4
1. a) P = 122 calories/sec. Enthalpy to melt 100 gms of Cu is 11.1 kcalories.
Melting occurs in 91 sec, if the heat capacity is 6.65 cal/deg. mole
b) At steady state all of the heat goes into the water. A1: = 3. 7C.
2. tan(1800-29) =rjD (back-reflection geometry).
a) Bragg's Law:
nA. = 2d sin 9
A. = 1.54178 A
(aO)Fe = 2.8665

d = 1.43325 A .

r = 0.114m

3. For Cr radiation. 1:=40 microns.

1P
I

(2S)b - (2s)A

---+---~-----'";B

= 2.6

Chapter 5
a) Since there is curvature in "d" vs sin2 tp:
(i) O"ij = O"ij (z) .
(ii) The stress state in the surface is homogeneous;

O"~=constant

for all <1>.

sin 2 tp

b) (i) Surface stress 0"., from


analysis", -615 MPa.
(ii) Surface stress 0"., from two-tilt analysis", -612 MPa.
c)

0"33

Total tp-range
Low tp-range
High tp-range

-466
-629
-414

-466
-629
-414

146
73
113

(MPa)

Appendix A: Solutions to Problems

250

d)( d o )1p

'V-range

2.86734 A
2.86692
2.86715

Total
Low
High

e) 'V-range

0'11

0'22

0'33

Total
Low
High

-612
-702
-527

-612
-702
-527

0
0
0

(MPa)

Chapter 6
1. (i) 160.729
(ii) 0.029929
(iii) (~=Q.) =0.90795 A
(iv) Error'{ ~= 0) = 0.00006 A
2. 0'., = 608.8 MPa
. Stat. error = 20 MPa
Chapter 8

a) For each n, take the sum over all values of x. Do it for n=0,1,2 for the
instrumental profIle.
1
a/2
(21t(0)X)
:.Ao=12x="'fa/2f(x) cos 12
(1)
1

= 12 (10+4+1 +0.3+0+0+0+4+1 +0.3+0+0+0)


= 1.716667
1

a/2

(21t( 1

A1=12x="'fal /(x) cos

12

)x) (1)

= 0.833'33 + 0.288675 + 0.0416667 + 0 + 0 + 0 + 0 + 0.288675 + 0.041667


=1.49398.
Normalizing, AdAo = 0.87028.
A2= 1/12

x=~a/2 f(x) (cos 21t~~ )x )

(1)

=0.83333+0.166667 -0.0416667 -0.025 +0+0+0+0.166667


-0.0416667 -0.025
=1.03333.
Normalizing: A21Ao = 0.60194.

Appendix A: Solutions to Problems

251

All sine coefficients are zero because the profIles are symmetrical.
Likewise for the broadened profIle:
1

a/2

(21t (0)
x)
6
(1) =20/6=3.3333

a/2

(21t (61 ) x ) (1)=2.16667.

AO="6 x=L;a/2 f(x) cos


A t ="6 x=L;al/(x) cos

Normalizing: At!Ao = 0.6500.

a/2

(21t (62 ) x ) (1)=0.83333. .

A2="6 x=L;al/(x) cos

Normalizing: A2/Ao = 0.25000.


b) Column Length, L = na' 3' where a' 3 = 4 ( . eA. e
sm 2- sm 0)
For the two simpl~ profIles above, a' 3 = 42.6 A
c) Stokes Corrected Coefficients. If sine coefficients for both peaks are zero, to get
the Stokes corrected cosine coefficients, G n, use

where Hn are the broadened peak coefficients, Fn the instrument peak coefficients
and Ay the area of the instrument peak. The term Ay/a can be set = 1. Thus for our
simple profIles, Stokes corrected coefficients are
0.6500
0.2500
Go=1/1 =1;G t = 0.87028 =0.7469;G2= 0.60194 =0.41532

Appendix B

B.l Introduction
In Sects. 5.13, 5.14, the meaning of oscillations in "d" vs. sin 2 tp and the causes for
such behaviour were discussed. It was shown that such oscillations could result
from an inhomogeneous distribution of elastic residual strains. Such an inhomogeneous distribution can arise from an inhomogeneous distribution of plastic
deformation on a micro scale and/or from interaction strains generated in an
inhomogeneous material in response to an applied or macrostress. Over the years,
oscillatory "d" vs. sin 2 tp plots have been the subject of numerous studies and much
controversy. Various approaches, from elaborate non-linear least-squares curvefitting to tota-l disregard for oscillations by using two-tilt analysis for oscillatory
data, have been proposed and used. In this appendix some of the more applicable
methods are reviewed briefly. The reader is warned that at this point, none of the
techniques discussed below, or indeed any other techniques, have been able to
satisfactorily solve this problem.

B.2 The Marion-Cohen Method


This method presumes plastic deformation followed by partial (localized) elastic
recovery and the following assumptions are used [1]:
1. Only crystal regions of small dislocation density make an appreciable
contributioJl to the maximum in a diffraction peak.
2. In a material subjected to plastic deformation, the regions of small dislocation
density are initially under a generally compressive state of stress (due to
expansion in dislocation-rich regions).
3. Some of the dislocation-poor crystallite regions will be oriented to slip and
rotate during deformation to a more energetically favorable position (A
orientation t A regions) relative to the deformation geometry. This leads to the
formation of deformation texture. The compressive stress is relieved in some
regions by this plastic deformation, corresponding to an increase in lattice
spacing. Therefore, the A regions have a larger lattice constant, "dmax " = d A ,
than other dislocation-poor crystalline regions, whose compressive stress can
not be relieved because of an unfavorable orientation to the applied load
(B orientation, B regions) .

B.2 The Marion-Cohen Method

253

Diffractometer axis
Incident beam

Diffracted beam

Fig. B.1. Specimen geometry and


the definition of the angles cx,/3 in
MaJjon-Cohen method ( From
Marion [1])

4. The lattice constant "d", measured by x-rays, possesses a maximum, dmax> for
each orientation of the sample. From this maximum value, d max, the lattice
constant varies with orientation of the sample according to a distribution
function, until ethe "dB" value of the B regions is reached.
With these assumptions, the dependence of the interplanar spacing on the
orientation may be expressed as:
(B.1)
Here f( (l,~) is the distribution function of a particular (hkl) plane relative to the
sample coordinates and (l and ~ are defined in Fig. B.t.
If in addition to the microscopic strain distribution there are macroscopic
stresses in the material, the combined equation is given by:
(B.2)
{Here it is assumed that the macrostresses contributing to cr~ yield a linear "d" vs.
sin 2 '1' dependence. Thus, reaction strains, tiT, are assumed to be zero

[Eq. (5.53)].}

Thus, if f( (l,~) (which can be taken as the integrated or maximum peak


intensity) and d~", are measured as a function of 'I' for a particular <1>, the
unknowns ~ax,dB and cr~ can be solved by curve-fitting techniques.
As can be seen from assumption 3, these ideas are based on an inhomogeneous
distribution of plastic deformation. The method predicts that the oscillations in
"d" vs. sin 2 '1' and f( (l,~) vs. sin 2 '1' will have the same form, as shown in Fig. B.2.
There are, however, cases in which this relationship is not observed [1] and the
method is inapplicable in such cases. Furthermore, the solution is based on the
assumption that the maximum range in oscillations is observed in a given
measurement, which'may not be the case, as it depends on the 'I'-range examined.
More importantly, this method assumes, a priori, that elastic interaction effects
are negligible 1 . There are, however, some experiments that indicate that this may
Even though the starting assumptions of this method is somewhat restrictive, its fundamental
method of determining macrostress and the range of microstress variation is correct and may
be used when the starting assumptions are satisfied.

254

Appendix B

0.8382 ,..---...,----,-----.--,---,---,
0.8381

t 0.8380
u

0.8379

0.8378
a

1.0
0.8
-:T

0.6

o'

0.4
0.2
b

0.1

0.2

0.3

0.4

0.5

Fig. B.2. "d" vs. sin 2 \jl (a) and f( (X,~) vs.
sin 2 \jl (b) behaviour required for the use
of Marion-Cohen Method. (In both cases
similar oscillations are observed.) (From
Marion [1] )

sin 2 tjJ

not be true for all cases. When a textured material (which had been stressrelieved) is stressed elastically, oscillations appear in the "d" vs. sin 2 tp curve,
which, upon unloading, vanish [2,3]. This observation was utilized by Dolle and
Hauk [4] in a different approach. This method is discussed below.

B.3 Dolle-Hank Method (Oscillation-free Reflections)


Shiraiwa and Sakamoto [2] and Dolle and Hauk [4] proposed that the
oscillations in "d" vs. sin 2 tp are primarily due to elastic anisotropy. That is, the xray elastic constants vary with <f> and tp tilt in strongly textured materials because
grains with different orientations are sampled at each tp-tilt. For such a case, the
equation linking "dq,,,," with the applied stress is [4]

<'

dq,,,,' + t33ij
, ) 'aij >,
do do = E33 >= <(S33ij

(B.3)

where S~3ij are the single-crystal elastic compliances in the laboratory co-ordinate
system, Lj, and t~3ij are the elastic interaction terms of a grain and the surrounding
matrix. The carats in (B.3) indicate that the average is taken over all the

255

8.3 Dolle-Hauk Method (Oscillation-free Reflections)


1.06220 r - - - - - - - - - - - - - - ,

1.06210 rN

-1.06200

f-

CfA-21MPa

1.06190 L....L1_ _Tl...-_-'---I--'.I_-'I__l...-I----LI-----'

1.06220

1.06210

f-

-1.06200

f-

f-

42MPa

I---,I_-,I_ _IL..-_L..-I_-,---I_.L.1---1

1.06190 L..L.-

1.06220 r-

1.06210 fN

-1.06200

f-

1.06190 r-

,
0.1

I
0.2

63 MPa
1
1
0.3 0.4
sin 2 ",

I
0.5

0.6

0.7

Fig. B.3. Oscillatory "d" vs. sin 21jJ


behaviour in the 222 reflection from an
<x-brass sample loaded in situ on the
diffractometer (From Noyan and
Cohen [10])

diffracting crystallites. [This equation assumes, a priori, that inhomogeneous


micro-strain distributions (due to an inhomogeneous distribution of plastic
deformation) are negligible.]
The solution proposed for Eq. (B.3) simplifies the system even further; it is
assumed that the Reuss limit describes the stress state in the material.
Thus, all interaction stresses and strains are zero. Then the terms S~3ij are
calculated in the Reuss limit and aij are determined from a non-linear leastsquares solution. One important result of this analysis is that it predicts no
oscillations for "hoo" or "hhh" type reflections (for cubic materials) since the
S~3ij' calculated for such reflections, are equivalent to the isotropic elastic
constants [5F. Thus these authors concluded that, if one switches to an hOO or
hhh reflection after observing oscillations in "d" vs. sin 2 \jJ in an hkl reflection, one
should obtain regular "d" vs. sin 2 \jJ behavior which can then be analyzed by the
methods described in Sects. 5.3, 5.4. The theoretical prediction of linear hOO and
hhh reflections is not always satisfied in practice (Fig. B.3 ). This is so since the
Reuss model, as discussed in Sect. 3.10, does not satisfy compatibility conditions,
This can be simply illustrated for hoo reflections by considering Eq. (3.72) which describes the
relationship between single-crystal elastic constants and the average value (1 + v) IE:
1+V)
h2k2+k212+12h2
( - E =Sllll-Sll22- 3S0 r , r=
2 2 2 2 . It can be seen that for hoo reflecR
(h +k +1 )
tions r=o, thus there is no orientation dependence for (1 :v) R. For these reflections the
crystal behaves like an isotropic material. A similar derivation for hhh is given in [6].

Appendix B

256

and would cause void formation at boundaries between grains of differing


crystallographic orientations along the load axis (Sect. 3.4, Fig. 3.7). Thus, the
interaction terms in non-random inhomogeneous solids cannot be neglected in
the real world.

B.4 Methods of Peiter and Lode


These authors [7,8] suggest that oscillations in "d" vs. sin 2 'P arise because of the
presence of stress gradients in the surface layers. For example, consider the stress
proflle with depth shown in Fig. (B.4); since x-rays penetrate to different depths
at each 'P-tilt (Sect. 5.9), the net average stress affecting "d,/' will change sign
during the measurement, causing oscillations in "d" vs. sin 2 'P. In their solution for
the stresses in such cases, Peiter and Lode proposed two methods: the <I>-integral
method and the 'P-integral method. In the <I>-integral method, the strain (I:: ~ 3 ) 4"1' is
expressed as a trigonometric polynomial in <1>:

(1::~3>"''I' = ~ Bo + Bicos <I> + Al sin <I> + B2coS 2 <1> + A2sin 2 <1> + ... .

(B.4)

The Fourier coefficients A;,B; are then evaluated by integrating Eq. (B.4) over
3600 in <1>:
1t.

Ak
B =
k

360,

J0 (1::33 \'1'.

sin k<l>
kA.. d<l>.
cos ~

(B.5)

These coefficients are linear functions of the strains in the surface coordinate
system [these equations can be obtained by comparing Eq. 5.4 with Eq. (B.4)].
Thus, it is possible to determine the strains from the Fourier coefficients, and the
stresses from the appropriate form of Hooke's law.
The 'P-integ~al method is similar to the <I>-integral with one exception. The
strain (1::~3> is expressed as a polynomial in terms of'P and integrated over
'P from -45 to 45 in order to obtain the coefficients from which the strains are
obtained.
These methods are not widely used. The <I>-integral technique is a very time
consuming procedure and is rarely used. It does yield values in excellent agreement
with the standard methods (discussed in Chap. 5) for regular stress/strain

Or7----------~~--------~~
'-_/

"

" 0

Fig. B.4. Residual stress distribution


with depth that may cause oscillations in "d" vs. sin 2 1jJ

References

257

distributions and has the added advantage that for a given tp-tilt, variation of <I>
does not change the penetration depth t, so that the same x-ray average is seen at
all <I>-rotations [9, 10J. The coefficient matrices A~k,B~k for the tp-integral method
are usually ill conditioned, which magnifies experimental errors, causing large
errors in the final stress values [11]. In these methods, however, causes of
oscillations other than macro-stress gradients in the z-direction are not taken into
account. These methods should not be used for analysis of oscillatory data unless
it is determined independently that the oscillations are indeed caused by a macrostres gradient. The idea that a stress gradient similar to that shown in Fig. B.4
causes oscillations in "d" vs. sin 2 tp is quite valid and such a gradient can occur in
practice due to a peculiar load distribution during deformation processing.

B.5 Use of High Multiplicity Peaks


In the discussions on elastic constants (Sects. 3.9 - 3.11) it was shown that
isotropic elastic constants are obtained if single crystal elastic constants are
integrated over all possible directions. The normal to high multiplicity planes, i.e.,
those planes with a large number of equivalents in a unit cell (Sect. 4.1 0), may
have a higher probability of being oriented in a more-or-Iess random manner in a
textured or plastically deformed material. In such a case, the average interaction
effects will integrate to a constant value or to zero as discussed in Sect. 5.13. Thus,
regular "d" vs. sin 2 tp data may be obtained from such reflections. Recent work by
Hauk et al. [10], for example, shows such behavior for the 732-651 peak (Mo
radiation was used in analysis) in textured steel where a 211 reflection (examined
with Cr radiation) exhibited oscillations. Such reflections, coupled with texture
"analysis may be used to determine the average strain tensor if regular "d" vs. sin 2 tp
behavior is observed. In interpreting such data, the effect of the changing
penetration depth (5.4 vs. 15.4 microns respectively) must also be taken into
account. Analysis Of such high-angle peaks may be especially beneficial with
rotating anode or synchrotron sources since such peaks usually have very low
diffracted intensities.

References

2
3
4
5
6

R. Marion, X-ray Stress Analysis of Plastically Defonned Metals, Ph. D. Thesis,


Northwestern University, Evanston, II. (1972)
T. Shiraiwa and Y. Sakamoto, Proc. 13th Japan Congress on Mat. Res., The Soc. Mat. Sci.,
Japan, Kyoto 25 (1970)
V. Hauk, Residual Stress and Stress Relaxation, E. Kula and V. Weiss, Eds., Plenum Press,
New York, 117, 1982
H. Dolle and V. Hauk, Z. Metallkde., 69, 410 (1978)
H. Dolle, J. Appl. Cryst., 12, 489 (1979)
H. Dolle and J.B. Cohen, Met. Trans. A, llA, 831 (1980)

258
7
8
9
10
11

Appendix B
W. Lode and A. Peiter, Hiirterei-Tech. Mitt., 32, 235 (1977)
W. Lode and A. Peiter: Metall, 35, 578 (1981)
C.N.I. Wagner, B. Eigenmann and M.S. Boldrick, preprint of The Phi-integral method for
X-ray Residual Stress Measurements, Dept. of Mat. Sci. & Eng., University of California,
L.A, CA 90024 (1986)
B. Eigenmann, Computer-aided X-ray Residual Stress Analysis in High Strength Aluminum
Alloys for Production Control during the Manufacturing Process, Master's Thesis,
Universitiit Karlsruhe (TH), West Germany
C.N.I. Wagner and M.S. Boldrick, The Psi-differential and Integral Methods for Residual
Stress Measurements by X-ray Difraction, Dept. of Mat. Sci. & Eng., University of
California, L.A, CA 90024 (Paper presented in the 1983 Denver X-ray Conference)

Appendix C: Fourier Analysis

If a function, f( x), is periodic with period (a), has no more than a fmite number
of discontinuities in a fmite interval ofx, and the following integral is finite (these
restrictions are satisfied by most functions encountered in natural phenomena):
a/2

-a/2

If(x)ldx,

then f( x) can be decomposed into an infinite Fourier series of sine and cosine
terms:
co
21tnx
co B . 21tnx
f(x) =Ao+2 n"f 1 ~cos-a- +2 n"f 1 nsm-a-

(Col)

where Ao, and all the coefficients An and Bn must be determined.


One of the important features of this Fourier series is the orthogonality of the
individual terms, i.e., the fact that different terms do not interact. We can see this
from the following integrals:

+Ja / 2 cos--cos-21tnx
21tmx d
{O
X=
-8/2

if n =4= m,
.

a/2 if n=m,

21tnx
21tnx
cos--sin--dx=O,
-a/2
a
a

+ 8/2

(C.2b)

. 21tnx . 21tmx
+Ja / 2 sm--sm-d X= {O

-8/2

(C.2a)

ifn=4=m,
.

a/2 if n=m.

(C.2c)

This orthogonality of the trigonometric functions suggests the way to determine


the coefficients ofEq. (C.t ). For the given function f( x), substituting Eq. (C.1 )
into Eqs. (C.2a - c) we have:
21tffix
f(x)cos--dx =Ama ,
-8/2
a

( C.3a)

21tmx
f(x)sin--dx =Bma,
-a/2
a

(C.3b)

+a/2

+a/2

+a/2

-a/2

f(x)dx

=Aoa.

(C.3c)

The last relation tells us that the area of the function in one period is proportional
to the first coefficient, Ao. Furthermore, if the function is even (i.e., symmetrical
about the x = 0 axis) , then from Eq. (C.3b) all Bn = 0 and a cosine series results. If

260

Appendix C: Fourier Analysis

f(x) is odd, or inverted through the origin (i.e., antisymmetric), then all the
Au = 0 from Eq. (C.3a) and a sine series results. For a general function, f( x), both
terms are necessary.
We can write our Fourier series in a slightly different way if we note from
Eqs. (e.3a,b) that ~-n)=Au and B<-n)= -Bn. Then,
f(x) =

+ 00

21tnx
+ 00
21tnx
Aucos-- + L Bnsin- - .
-00
a
-00
a

(e.4 )

The restrictions we have placed on the coefficients maintain a general series for
f(x). For example, when n= -4 the sine term is negative, and if Bn did not
change sign in Eq. (e.4) this sine term would cancel the term for n = + 4 and
similarly for all n, -n; only the cosine series would remain. Note that Bo =0, so it
may be included in this compact notation.
We can write the series in a still more compact way if we define a complex
coefficient:
(C.5)

Cn=Au~iBn ,

f(x) =

+00

-00

C ne- 2"inxa.

(e.6a)

Negative values of n correspond to clockwise rotations on the complex plane,


whereas positive values are counterclockwise rotations. Replacing C n by
Eq. (e.S) and recalling that eiO = cos e+ i sin e,
f(x) =

+ 00 (
21tnx
L Aucos--00
a

. 21tnx )
+ Bnsm-

. +00 (
. 21tnx
21tnx)
L Ausm-- Bncos-- .
-00
a
a

-1

(e.6b)

The second sum must be zero to correspond to Eq. (C.4), which holds for a real
function. This condition is satisfied since ~ _ n) = Au and B( - n) = - Bn ( i.e.,
C(-n)=C:, where the asterisk denotes "complex conjugate"). Using the orthogonality relations, Eqs. (e.2a - c ) , the reader may show that the complex Fourier
coefficients C m can be obtained from
aCm=

+a/2

f(x)e<+2nimx /a)dx.

(C.7)

-a/2

Consider two functions f1 (x) and f2 (x) each with period a, multiplied together
and averaged over this period,

-1 +a/2
J fdx)f2(x)dx = -1 +a/2
J
a

-a/2

+ 00

-a/2

n=-oo

1 +00
= - L

(Cdn

n=-oo

+00

n=

-00

}
(Cdne(-2ninx/a) f2(x)dx
+a/2

-a/2

f 2 (x)e(-2ninX/a)dx

(Cdn(C 2 ):,

(e.S)

Appendix C: Fourier Analysis

261

Using the Fourier series representations of these functions, we have shown that
the average value ofthe product of the functions is the sum of the products of the
coefficients. Suppose the two functions are identical. Then we are taking the
average value of [f( x)] 2 over the period. This result may be written as
1

+~2

-~2

- J

(f(x) )2dx=

+00

L IC nl2

n=-oo
+00
n=

(C.9)

-00

According to Eq. (C.9), which is known as Parse val's theorem, the sum of the
squares of the Fourier coefficients will give us the average "intensity" of the
function. We can see this result directly from the series itself; look at Eq. (C.4),
and consider the average value of the square of one term, say, A2cos 21t2x/a. The
average value of the square is 1/2(A2) 2. But there are +n and -n values, so the
average of the pair of these is (A 2 ) 2. The same is true for all terms. The sum of the
squares of the coefficients is merely the sum of the average "intensity" in the unit
cell of each wave in the series, and this is the same average we obtained by squaring
the total series to arrive at Parseval's theorem. This is so because the terms do not
interact.
Ifwe plot the coefficients An and Bn as a function ofn/a, such a plot will consist
of a series of discrete lines spaced l/a apart. As the period a increases, these lines
move closer together, and as the period approaches the limit of infinity, a
continuous curve of coefficients develops. This limiting case of a Fourier series will
allow us to represent non periodic functions as combinations of orthogonal plane
waves. Rewriting the series, Eq. (C.l ), and using the definitions of the
coefficients,
1 +a/2
00 1
21tnx +a/2
21tn~
f(x)= - J f(~)d~+ 2 L -cos- J f(~)cos--d~
a -a/2
n=l a
a -a/2
a
00 t
21tnx +a/2
21tnll
+2 L -sin-- J f(ll)sin--dll
n=la
a -a/2
a
1 +a/2
2 00 +a/2
( X -11 )
=- J f(~)dll+- L
J
f(ll)cos21tn
dll,
a -a/2
a n=l -a/2
a

( C.10)

where 11 is a dummy integration variable. Let the period a approach infinity; the
first term becomes zero, and if the function still satisfies the conditions for a
Fourier series, the integral is finite. If we further state that when l/a approaches
zero as n approaches infinity, n/a approaches zero, then n/a can be considered as a
continuous variable, which we shall call s. With these conditions, Eq. (C.1 0)
becomes
00
+00
f(x) =2 J ds J f(ll)cos21ts(x-ll)dll.
(C.1l )
o
-00
This important result is known as Fourier's integral theorem.

262

Appendix C: Fourier Analysis

Expanding the cosine of Eq. (C.1i),


co

f( x) = 2 Jds
o

co

-co

f( 11) {cos 2nsll cos 21tsx + sin 2nsll sin 21tsx} dll

co

co

-co

= 2 Jcos 21tsxds

f ( 11) cos 21tSlldll

co

co

-co

+ 2 Jsin 2nsxds

f( 11) sin 2nslldll

( C.12)

Representing the cosine and sine integrals of f( 11) by a (s) and b (s),
co

a ( s) =

f( 11) cos 21tSlldll ,

(C.13a)

J f(ll)sin2nslldll,

(C.13b)

-co
co

b(s) =

-co

Eq. (C.i2) befomes


co

f ( x) = 2 J {a ( s ) cos 2nsx + b ( s ) sin 2nsx} ds .


o

Multiplying and dividing by [a 2 (s) + b 2 (s)] 1 / 2,


- co 2
b2 )
f (x)-2![a
(s)+
(s]

1/2 {

a (s)

[a2(s)+b2(s)]1/2cos2nsx

b(s)
.
}
+ [a2(s) + b2(s)] l/2sm2nsx ds.
Now, letting Rs= [a 2 (s) +b 2 (s)]1/2 and tanO s = -b(s)ja(s),
co

JRscos{2nsx+<I>s}ds

f(x)=2
=

co

JRs {exp [i ( 21tsx + <l>s) ] + exp [ o

i ( 2nsx + <l>s) ] } ds .

From the definitions,


a( -s) =a(s),

<I>-s= -<I>s

b( -s) = -b(s),

Rs=R-S'

so we can rewrite the integral for f( x) in two parts:


f( x) =

co

JRsexp ( o

co

2nisx) exp ( - i<l>sl ds

JRsexp ( o

21tisx) exp ( - i<l>s) d ( - s) .

(C.13c)

Appendix C: Fourier Analysis

263

Thus,
f( x ) =

00

JRsexp ( o

21tisx) exp ( - i<l>s) ds

Rsexp ( - 21tisx ) exp ( - i<l>s ) ds ,

-00

00

Rsexp ( - 21tisx) exp ( - i<l>s) ds .

-00

Finally, letting
F ( s) = Rsexp ( - i<l>s) = a ( s) + ib ( s) ,
f ( x) =

00

F ( s ) e - 2lt;SXds .

(C.14 )
(C.15 )

-00

This is the more common expression of Fourier's integral theorem. Note from
Eqs. (C.13a-c) that when f( x) is real and even, b( s) =0 for all s, while for f( x)
real and odd, a (s ) = 0 for all s.
Substituting Eqs. (C.13a - c) in Eq. (C.14), we obtain
F(s) =

+00

f(x)cos21tsxdx+i

-00

+00

f(x)sin21tsxdx,

-00

or
F ( s) =

+00

J f ( x ) e + 2lt;SXdx .

( C.16)

-00

The function F (s) is known as the Fourier transform off( x). Note the similarity
between Eqs. (C.15) and (C.16). The two functions F (s) and f( x) related by
these equations are known as a "Fourier pair", and the variables s and x are
known as a "conjugate pair". We can abbreviate these two equations as:
Tf(x) =F(s)

(C.17a)

and:
T-1F(s) =f(x) ,

(C.17b)

where T and T- 1 are read as "transform of' and "inverse transform of',
respectively. We can think of Eqs. (C.15) and (C.16) in terms of two spaces: a
Fourier transform space, s, for F (s), and a "real" space, x, where our function
f( x) is plotted. From the nature of the Fourier transform, we can see certain basic
properties:
F(O) =

+00

-00

f(x)dx, or f(O) =

+00

F(s)ds.

(C.18 )

-00

The integral of a function in one space equals its transform at the origin of the
other space. Other important properties to note, which the reader can readily

Appendix C: Fourier Analysis

264

F(S)

fIx)

-II

-I

+11

x-

+I

5-

Fig. C.l. A schematic representation of ( a) a rectangular function


f ( x ) and ( b) its Fourier transform F(s)

prove, are
Tcf(x) =cF(s) ,

(C.19a)

TLCifi(x) = LciFi(s) ,

(C.19b)

(C.19c)
where c is a C6nstant. An important example occurs for c = -1 in (C.19c):
Tf( -x) =F*(s) .

( C.19d)

Finally, the transforms of nth derivatives of f( x) may be expressed in terms of


F(s) by
(C.1ge)
As an example of a transform, consider the function plotted in Fig. C.l a,
f(x)=

{ 1 for Ixl ~ 1/2,


o (orlxl>1/2.

Its transform is then


F (s) =

+00

f(x)e+2nisxdx=

-00

+1/2

-1/2

e+2nisxdx,

= ( enis _ e - niS) /2nis = sin ns/ns ,

( C.20a)

and
F(O) =

+ 1/2

J' f(x)dx=l.

-1/2

The transform is plotted in Fig. C.1 b. A similar function is f( x) = 1 for


Ixl~t/2,f(x) =0 for Ixl>t/2. The transform of this function is
F ( s) = sin nts/ns .

(C.20b)

Appendix C: Fourier Analysis

265

Now, if t is large the tranform is sharp, and if t is small the transform is broad.
Thus the transform has a reciprocal nature with respect to real space. This is also
true of the Fourier coefficients of a function; they fall off with n more rapidly if the
function is broad than if it is sharp.
A very useful treatise on Fourier mathematics is "An Introduction to Fourier
Analysis", by R. D. Stuart, Methuen Co., Ltd., London (1961).

Appendix D: Location of Useful Information in


"International Tables for Crystallography" 1

1)

X-ray Wavelength

Vol. IV, pp. 6-43

2)

Filters for X-rays

Vol. III, pp. 75,76

3)

X-ray Atomic Scattering Factors

Vol. IV, pp. 72-146

4)

Dispersion Corrections for


X-ray Atomic Scattering Factors

Vol. IV, pp. 149,150

5)

Neutr6n Atomic Scattering Factors

Vol. IV, pp. 270,271

6)

X-ray Absorption Coefficients

Vol. IV, pp. 47-66

7)

Neutron Absorption Coefficients

V01. III, p. 197

1 "International Tables for Crystallography", Vol. I (1952), Vol. II (1959), Vol. III (1962)
(K. Lonsdale, ed.), Vol. IV (1974) (J. A. Ibers and w. C. Hamilton, eds.). Kynoch Press,
Birmingham, England.

Appendix E: Values of Gx [Eq. (4.37)]


for Various Materials (By Dr. M. James)

Deptb of penetration calculation: Iron; Cr 211. Cr radiation: 9=78.00000; density=7.870000


g/cm 3
Element
Fe

mU/Q, crn1/g %

Mass Fraction

108.000

1.000

100.000

mu total = 849.9600
Depth ip microns

G.

1jl=0

18

26

32

39

45

0.50
0.67
0.95

3.988
6.379
17.238

3.775
6.038
16.316

3.546
5.672
15.327

3.323
5.315
14.360

3.008
4.811
12.999

2.693
4.307
11.638

0.140
0.223
0.603

0.131
0.209
0.565

0.118
0.189
0.512

0.106
0.170
0.458

Depth in Mils

0.50
0.67
0.95

0.157
0.251
0.679

0.149
0.238
0.642

Deptb of penetration calculation: Aluminum; Cr 311. Cr radiation; 0 = 60.50000; density =


2.699000 g/cm 3
Element

mU/Q, cm1/g %

Mass fraction

AI

158.000

1.000

100.000

mu total =426.4420
Depth in microns

G.
0.50
0.67
0.95

7.612
12.176
32.900

18

26

32

39

45

7.133
11.409
30.828

6.614
10.580
28.587

6.103
9.762
26.378

5.374
8.595
23.225

4.630
7.406
20.012

0.260
0.417
1.125

0.240
0.384
1.039

0.212
0.338
0.914

0.182
0.292
0.788

Depth in mils

0.50
0.67
0.95

0.300
0.479
1.295

0.281
0.449
1.214

Appendix E: Values of G. for Various Materials

268

Depth of penetration calculation: MAR - M246; Cr 220. Cr radiation; 0 = 64.20000; density =


8.440000 g/cm 3
Element

mujQ, cm1jg %

Mass fraction

Ni
Co
W
Cr
Mo

145.700
124.600
495.500
85.710
457.400

0.562
0.082
0.257
0.065
0.034

68.500
10.000
10.000
9.000
2.500

mu total = 2008.720
Depth in microns

G.

11'=0

18

26

32

39

45

0.50
0.67
0.95

1.553
2.485
6.714

1.441
2.305

1.319
2.109
5.699

1.197
1.915
5.174

1.022
1.635
4.418

0.842
1.346
3.638

0.052
0.083
0.224

0.047
0.Q75
0.204

0.040

0.033
0.053
0.143

6.i27

Depth in mils
0.061
0.098
0.264

0.50
0.67
0.95

0.057
0.091
0.245

0.064
0.174

Depth of penetration calculation; 440 stainless steel. Cr radiation; 9=78.00000;


density
7.800000g/cm3

Element
Fe
Cr
C
Mn
Mo
Si

mujQ,cm1jg %

Mass fraction

108.000
85.710
14.460
96.080
457.400
202.700

0.813
0.169
0.002
0.005
0.008
0.002

79.800
17.800
1.030
0.480
0.480
0.410

mu total=811.2000
Depth in microns

G.

11'=0

18

26

32

39

45

0.50
0.67
0.95

4.179
6.684
18.061.

3.956
6.327
17.095

3.716
5.943
16.059

3.481
5.568
15.047

3.151
5.041
13.620

2.821
4.513
12.194

0.146
0.234
0.632

0.137
0.219
0.592

0.124
0.198
0.536

0.111
0.178
0.480

Depth in mils
0.50
0.67
0.95

0.165
0.263
0.711

0.156
0.249
0.673

Appendix E: Values of Ox for Various Materials

269

Depth of penetration calculation: Ti 5Al- 2.5Sn Cu 213. Cu radiation;


density

= 4.460000 g/cm

e=70.50000;

Element

mU/Q,cm 2 /g %

Mass fraction

Ti
Al
Sn

202.400
50.230
253.300

0.911
0.028
0.061

92.500

5.000

2.500

mu total = 892.0000
Depth in microns

Ox

\jJ=0

18

26

32

39

45

0.50
0.67
0.95

3.662
5.858
15.829

3.437
5.498
14.855

3.194
5.108
13.803

2.954
4.725
12.767

2.612
4.178
11.290

2.265
3.623
9.789

0.126
0.201
0.543

0.116
0.186
0.503

0.103
0.164
0.444

0.089
0.143
0.385

Depth i,n mils


0.50
0.67
0.95

0.144
0.231
0.623

0.135
0.216
0.585

Appendix F: A Compilation of X-ray Elastic Constants!


(By Dr. M. James)

Material

Radiation

hkl

(1+v)/E
x to- 8 psi

Ref.

Aluminum
A15083-H23
A15083-H23
A15083-H23
Al2219-T87
Al 7075

Co
Cr
Co
Cu
Cr
Cr

420
311
420
511/333
311
311

14.09
12.53
12.55
12.55
11.71
11.33

[6]
[3]
[3]
[3]
[8]
[7]

ARMCO Fe
0.39 %C steel
0.73 %C steel
0.73 %C steel
Fe (4.3 %C, 0.57 %Si,
3.96 %Mo)
Fe(O.1 %C,0.82 %Mn,0.3 %Si
4.9 %Ni,0.55 %Cr,0.57 %Mo
D-6ac
200 Maraging steel
Railroad steel
HSLA 328
1045
1045
4820, carburized
4340, Rc50
410SS, Rc22
4tOSS, Rc42
422SS, Rc34
422SS, Rc39
304SS

Cr
Cr
Cr
Co

211
220
211
310

2.41
3.88
4.38
5.16

[5]
[6]
[6]
[6]

Fe

121 (Fe 3C)

4.76

[2]

Cr
Cr
Cr
Cr
Cr
Cr
Co
Cr
Cr
Cr
Cr
Cr
Cr
Cr

211
211
211
211
211
211
3tO
211
211
211
211
211
211
220

4.38
3.24
4.26
4.06
3.50
3.49
4.99
3.87
4.08
3.91
3.98
3.79
3.83
4.95

[2]
[8]
[8]
[8]
[10]
[5]
[5]
[4]
[7]
[7]
[7]
[7]
[7]
[7]

Incoloy 903
Incoloy 903
Incoloy 800

Cr
Cu
Cr

220
331
220

3.21
7.09
4.27

[7]
[8]
[7]

1 Converted from data given in:


[1]
[2]
[3]
[4]
[5]

Braski and Royster 1967


Hanabusa et al. 1969
Hilley et al. 1967
MacDonald 1966
Marion and Cohen 1977

[ 6]
[ 7]
[ 8]
[ 9]
[to]

Maucherauch 1966
Prevey 1977
Ranganathan 1976
Wooden et al. 1960
Schlosberg 1979

Errors are not generally given in these references; they are useful in comparative studies, but for
highest accuracy they should be measured.

Appendix F: A Compilation of X-ray Elastic Constants

271

Material

Radiation

hkl

Incoloy 800
Inconel6oo
Inconel6oo
Inconel 718
Inconel 718
Inconel 718
Inconel X750
Monel K500
Nickel
Nickel
Ti-6A1-4V
Ti-6A1-4V
Ti-6A1-2Sn-4Zr-2Mo
Ti-SAI -1Mo-1V

Cr
Cr
Cu
Cu
Cu
Cr
Cr
Cu
Cu
Cu
Cu
Cu
Cu
Cu

420
220
420
331
331
220
220
420
420
331
213
213
213
213

4.65
4.73
4.33
4.35
5.00
3.21
2.71
4.76
4.36
4.83
8.20
7.51
6.76
6.99

[7]
[7]
[7]
[8]
[7]
[7]
[7]
[7]
[6]
[6]
[7J
[1]
[7]
[1]

ex-Brass
~-Brass

Co
Co

400
310

10.62
13.92

[6]
[6]

Zircoloy-2

Cr

104

8.48

[9]

Copper

Co

400

9.35

[6]

Tungsten

Co

222

2.15

[6]

Uranium

Cu

116

2.56

[9]

(1+v)/E
x 10- 8 psi

Ref.

References
Baucum, W.E., and Ammons, A.M. (1973) Adv. in X-Ray Analysis 17, 371-382
Braski, D.N., and Royster, D.M. (1967) Adv. in X-Ray Analysis 10, 295-310
Cohen, J.B. (1964) Report on Tungsten Lattice Parameter Round Robin, X-Ray Subcommittee
of SAE Iron and Steel Technical Committee, Div. 4
Dolle, H., and Hauk, V. (1978) Z. Metallkunde 69, 410-417
Esquivel, A.L. (1969) Adv. in X-Ray Analysis 12, 269-298
Faninger G. (1970) J. Soc. Mat. Sci. 19,42-57
French, "tl.N. (1969) J. Amer. Cer. Soc. 52, 271- 275
Hanabusa, T., Fukura, J., and Fujiwara, H. (1969) Bull. of J.S.M.E. 12, 931 -939
Hilley, M.E., Wert, J.J., and Goodrich, R.S. (1967) Adv. in X-Ray Analysis 10,204-294
Hilley, M.E., Larson, J.A., Jatczak, C.F., and Ricklefs, R.E. (eds.) "Residual Stress Measurements by X-Ray Diffraction," SAE Information Report J784a. (1971) SAE, Pennsylvania
James, M.R. (1977) Ph.D. Thesis. Northwestern University, Evanston, Illinois
James, M.R., and Cohen, J.B. (1977) Adv. in X-Ray Analysis 20,291-308
Kelly, C.J., and Eichen, E. (1973 Adv. in X-Ray Analysis 16, 344-353
Kelly, C.J., and Short, M.A. (1970) Adv. in X-Ray Analysis 114, 377-387
Kirk, D. (1971) Strain 7, 7 - 14
Koistinen, D.P., and Marburger, R.E. (1959) Trans. ASM. 51,537-555

272

Appendix F: A Compilation of X-ray Elastic Constants

Kurita, M. (1977) Bull. J.S.M.E. 20, 1375 -1383


MacDonald, B.A. (1970) Adv. in X-Ray Analysis 13,487-506
Macherauch, E. (1966) Exp. Mech. 6,140-153
Macherauch, E., and Wolfstieg, U. (1977) Mater. Sci. and Eng. 30,1-13
Marion, R.H. (1972) Ph.D. Thesis. Northwestern University, Evanston, Illinois
Marion, R.H., and Cohen, J.B. (1977) Adv. in X-Ray Analysis 20, 355 - 368
Prummer, R. (1970) Proc. 6th Inter. Conf. on Nondestructive Testing, Hannover, Germany
Prummer, R., and Macherauch, E. (1965) z. Naturforschg. 20A, 1369-1370
Prummer, R., and Macherauch, E. (1966) Z. Naturforschg. 21A, 661-662
Prevey, P.S. (1977) Adv. in X-Ray Analysis 20,345-354
Ranganathan, B.N., Wert, J.J., and Clotfelter, W.N. (1976) J. Test. Eval. 4, 218-219
Rutledge, A.L., and Taylor, R.M. (1972) J. Strain Analysis 7, 1-6
Schlosberg, W.H. (1979) Ph.D. Thesis, Northwestern University, Evanston, Illinois
Singh, AX., and Balasingh, C. (1971) J. Appl. Phys. 42, 5254-5260
Society of Materials Science (1973) "Standard Method for X-Ray Stress Measurement," The
Society of Materials Science, Japan
Taira, S., Hayashi, K., and Watase, Z. (1969). Proc. of the 12th Japan Congress on Materials
Research, pp. 1 - 7
Wooden, BJ., House, E.C., Ogilvie, R.E. (1960) Adv. in X-Ray Analysis 3,3331-336
Zantopulos, H., and Jatezak, C.F. (1970) Adv. in X-Ray Analysis 14, 360-376

Subject Index

Absorption (neutrons) 114


x-rays 82
absorption coefficient 82
absorption factor (flat plate) 100
acoustic waves 4
alignment methods 203
alignment errors 196
anelastic limit 44
anisotropic distribution 13
anisotropic solids 27
anisotropic Young's modulus 36
annealed powder method 127,205
asymmetric peaks-peak location 178-181
atomic coordinates 29
atomic scattering factor 86
average strains 131, 135, 136
average stress 137
average microstress 54
averaging function 130
Barkhausen noise 4
biaxial stress analysis 122
Bragg's law 87
Bragg's law-differentiation of 165
Bravais lattice 29
Brittle materials, yielding of 39, 40
bremsstrahlung radiation 77
bulk elastic constants (selected values)
bulk modulus 27
Calculation, micros tresses 56
pseudo-macrostresses 58
Cauchy function 169 ,
centroid 167
characteristic radiation 77
choppers 112
collimators 91
compatibility conditions 38
compliance moduli (selected materials)
compliance tensor 33,
composition gradient 157
Compton scattering 85
contracted notation 33
coordinate axes-transformation 20
critical resolved shear stress 42
cross-correlation method 176
crystal monochromators 89, 90

69

35

crystal systems (definitions) 29


crystallographic directions 30
plane orientations 30, 31
Depth of penetration, x-rays 110
neutrons 114, 154, 155
dead time correction 108
Debye rings 95
Defocusing errors, omega goniometer 191
psi goniometer 200
difference method 175
differential plastic deformation 53
diffracting volume-neutrons 154
x-rays 134, 135, 144
diffraction patterns, single crystal 93
polycrystal 94
diffraction elastic constants 69
diffraction peak, selection for stress
measurements 164
diffractometer 96
diffractometers, use in stress
measurements 211
dilatation 17
direction cosines 20
displacement 17
distortion 17
distribution function 253
dummy suffix 22
divergence, collimator 91, 92
horizontal (diffractometer) 192
vertical 195
Dolle-Hauk method 254
Doublet, effect on peak shift 179, 180
Eigenvectors 23
Einstein suffix notation 22
Elastic constants 25
bulk averages 63-69
x-ray average 69-72, 149
elastic incompatibility 53
elastic limit 13, 44
elastoplastic material 40
electropolishing 206
engineering shear strain 26
equilibrium equations 37
macrostresses 51
microstresses 54

274

Subject Index

equivalent directions 31
equivalent inclusion method 58
equivalent slip systems 42
errors, beam optics 190
omega goniometer 190-199
psi goniometer 200-203
Eshelby's tensor 56, 57
Fatigue crack, associated residual stresses
fatigue response 1, 8
filters 84
fixed count method 182
fixed time method 182
flow curves 39
fluorescent radiation 83
focal point 191
force balance 23, 38
force vector 15
Fourier analysis 238,259
Fourier series 234-238
fractional coordipates 29
free suffix 22 .

layer removal 206


line broadening 89
long range order 28
Lorentz factor 99
Lorentz-polarization factor 100
Lorentzian function-modified 168
LVDT measurement of strain 44, 45
8

Indexing, tensors 15
incoherent scattering 85
inhomogeneous distribution 13
inhomogeneous materials 44
inhomogeneous plastic flow 43, 44
instrumental errors 190-205
intensity (diffracted) 88, 97
interference (constructive) 87
isotropic distribution 13
K absorption edge 84
kinematic theory 90
Kronecker's delta 26
Kroner average 72
Lame's constants 27
lattice plane spacings 31
lattice point 28
Laue camera 93, 94

205

Neerfeld average 68
neutrons, absorption of 114
diffracting volume 154
penetration depth of 115
scattering of 115
neutron flux 111
nuclear reactors 111
normal stresses 15

Gaussian function 169


grinding stresses 9
Half-value breadth 167
hardware requirements 212
hexagonal systems-indexing 31
high multiplicity peaks 257
homogeneous distribution 13
homogeneous strain distribution 130
hook correction 243
Hooke's law 13, 25, 32
Hooke's law-anisotropic materials 33
horizontal divergence-effect on focusing
hydrostatic strain 19

Machining stresses 8
macrostress 47
macrostress gradient-correction for
magic mirror 1
Marion-Cohen method 252
matrix notation 33
microstress 47, 53
Miller indices 30
Monochromators 90
mosaic structure 89,240, 241
Moseley's law 78
multiplicity 98

Offset yield stress 45


oscillation-free reflections 254
oscillations 149, 252
oscillatory "d" vs. sin plots 119, 149-153
overpeening 2
192

Parabola fitting 171


parafocusing method 132
parallel beam geometry 204
Parseval's theorem 261
peak location 166
peak shift determination 175,176
peak shift, due to doublet summation 179,
180
Pearson function 168
Peiter-Lode methods 256
penetration depth (omega goniometer) 134
(psi goniometer) 135
periodicity 28, 29
permutation tensor 39
phase transformation stresses 2
pinhole patterns 95
plane indexing 30
plane indexing (hexagonal systems) 31
plastic flow 39
point lattice 29
Poisson's ratio 26

Subject Index
Poisson's ratio (selected materials) 27
polarization factor 85
for monochromators 91
polychromatic radiation 77
portable residual stress units 213, 214
position sensitive detector 105
potential well concept 32
principal strains 25
principal stresses 24
principal tensor components 22
proportional detector 104
proportional limit 44
pseudo-macro stress 54
psi-axis offset 199
psi-position missetting 202
psi-splitting 119
pulsed beam neutron source 114
pulse-height analyzer 105
Radial stresses 50
radiographic analysis 75
ratemeter 104
regular "d" vs. sin plots 118
residual stresses (definition) 47
analysis 47
.
macrostresses 47--49
microstresses 52-54
measurement methods 4--7
stress gradients 140
stress in a gear tooth 222
stress in bolted plate assembly 216
stress in crankshaft journals 220
stress in ring assemblies 215
stress in rolling 2
stress in solid circular cylinder 218
stress in shot-peened references 223
stress in welding 2
Reuss Average 65, 71
rigid body translation 17
Roentgen 75
rotating anodes 80
Safety precautions 212
scattering 84
from planes of atoms 86
from the unit cell 87
Schmid factor 43
Schmid's law 42.
scintillation counter 107
sealed tubes 80
self-consistent method for average
elastic moduli 67
shear modulus 26
shear stresses 15
short wavelength limit 77
shrink-fitting 50

275

Simpson's rule 167


sin2 psi technique 122
single channel analyzer 104
single crystals, elastic constants of 35
plastic deformation of 41
structures of 28
single exposure method 123
slip 41
slits 91
slit corrections (peak shape) 233
software requirements 212
solid state detector 108
Soller slits 92
space group 29
specimen alignment 203
specimen bending 49
specimen curvature 193, 201
specimen displacement 196-198,200
standard deviation (peak position) 183
statistical errors 181-190
statistical error (in stress) 186, 187, 189
statistical error (in x-ray elastic constants)
189
stiffness moduli 33
stiffness moduli (selected materials) 35
Stokes' correction 235
strain (in terms of lattice spacing) 117
strain invariants 25
strain, definition 14
stress, definition 14
stress constant, two-tilt method 123
single exposure method 125
variation with sin 2 psi 165
stress gradients 140
stress in a gear tooth 222
in bolted plate assembly 216
in crankshaft journals 220
in ring assemblies 215
in shot-peened references 223
in solid circular cylinder 218
stresses, indexing 15
stress invariants 24
stress measurement instruments 213
with neutrons 154
with acoustic methods 4
with hole drilling 4
with magnetic methods 4
stress pattern' calculations 5
stress, origin 1
stress-strain relations (isotropic solid) 27
structure factor 87
surface constraint 48
synchrotron radiation 80
Taylor series 32, 207
temperature factor 103
tensors 15

276
tensor notation 20
Thomson equation 85
time of flight analysis 114
timer/scaler 104
top 15% rule 171
total diffracted intensity 109
total stress state in surface deformation 61
transformation of axes 20
transformation law (fourth rank tensors) 34
transformation law (second rank tensors) 22
transverse strains 25
triaxial stress analysis 125
true absorption 83
true elastic limit 44
true strain 40
true stress 41
twinning 44
two-tilt method 123
Undulators 81
unit cell 28
unstressed lattice spacing 126
Variance (in count rate) 182
vectors 15, 21
Vegard's law 157
vertical beam divergence 195
Voigt average 63, 70
von Laue 76

Subject Index
Warren-Averbach analysis of peak shape
238
white radiation 77
working origin 173
X-ray line broadening 230
x-ray peak shape 232
x-ray peak shape analysis 230
x-ray spectroscopy 89
x-ray strain equation 118
x-ray tubes 80
x-rays
absorption of 82
absorption coefficient of 82
Bremsstrahlung 77
characteristic radiation of 77
Compton scattering of 85
filters 84
generation of 80
inelastic scattering of 77
minimum wavelength 77
monochromator for 90
scattering factor of 86
structure factor for 87
white radiation 77
Yield strength 45
yielding (x-ray determination of) 159
Young's moduli (for selected materials) 27
Young's modulus, anisotropic materials 36

You might also like