p897 Chap01

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 52

Chapter 1

Diffraction Analysis of Defects: State of the Art


Rozaliya I. Barabash and Gene E. Ice
Materials Science and Technology Division,
Oak Ridge National Laboratory,
Oak Ridge, TN 37831, USA

Materialsproperties are largely determined by the self-organization and collective


behavior of defects. A number of X-ray methods including Laue microdiffraction,
rocking curve measurements and reciprocal space mapping can be used to
characterize mesoscale dislocation structures and elastic strain. This chapter
describes how dislocations and other common defects cause local strain eld
uctuations and alter the structure factors of the cells in which they occur, as well
as in neighboring cells. These changes alter the diffraction conditions in reciprocal
space, and result in characteristic features in the diffracted intensity. These
diffraction features can be used to model the local strain and dislocation structure
of advanced materials. Different dislocation arrangements are considered such
as: randomly distributed statistically stored dislocations (SSDs) and geometrically
necessary dislocations (GNDs); dislocation loops; dislocation dipoles; dislocation
boundaries including incidental dislocation boundaries (IDBs) and geometrically
necessary boundaries (GNBs). The impact of twins is also discussed.

1.1. Defect Classication in the Kinematic Approximation


Defects break the long-range symmetry of a crystal lattice and displace
atoms from their ideal positions. To understand how defects impact on
diffraction, it is necessary to address four questions:
1. How do defects distort the lattice?
2. Are the defects ordered and/or correlated?

R. I. Barabash & G. E. Ice

3. Do they change the structure factors of their unit cells and surrounding
unit cells?
4. Do they interact with external elds?
As is well known, the scattering from atoms of an ideal perfect crystal
add constructively under the BraggLaue conditions, which results in
-function-like intensity maxima localized at reciprocal lattice sites of
the crystal. Real crystals always contain defects, which perturb the ideal
periodic arrangement of the atoms and change local structure factors. The
qualitative nature of diffracted intensity depends on whether it is possible to
relate the atomic positions to an average lattice. If the defects cause atomic
displacements, which are localized in the immediate vicinity of the defects,
then it is possible to relate a single periodic lattice to the crystal; typically,
if the mean square uctuations of the atomic displacements remain nite
and do not exceed the interatomic distance, then it is possible to relate an
average periodic lattice to the crystal with defects. These defects are called
defects of the rst kind (Krivoglaz, 1969, 1996).
Some defects distort the lattice in a manner that makes it impossible to
reference atomic positions to a single periodic lattice. However, it is often
still possible to relate slightly different periodic lattices to relatively large
parts of the crystal (for example, subgrains, dislocation cells, etc.). Such
defects are called defects of the second kind (Krivoglaz, 1969, 1996).
For the analysis of the diffraction, an Ewald sphere representation
is usually used (Fig. 1.1). In this representation, k 2 and k 1 are the
wave vectors along the diffracted and incident beam. In the kinematic
approximation, complex scattering amplitudes from the electron probability
density are integrated with the corresponding phases to determine the
scattering intensity, I(Q), for diffraction vector or momentum transfer, Q.
In electronic units, intensity of scattering by a monochromatic beam is
described by the equation (Krivoglaz, 1969, 1996):
2



iQr 

I(Q) =  (r)e dr  ,

where Q = k2 k1

(1.1)

Here, (r) is a total electron probability density at position vector r.


Integration in Eq. 1.1 is performed over the whole scattering volume. For an
innite ideal crystal, the intensity distribution as a function of momentum

Diffraction Analysis of Defects: State of the Art

Fig. 1.1. Ewald representation of the diffraction experiment: Ghkl is a reciprocal lattice
vector, Q is a diffraction vector or momentum transfer and 2 is a scattering angle between
the incident k 1 and scattered k 2 radiation wave vectors with wavelength . The origin of
the reference reciprocal lattice frame, 0, is located at the end of the incident k 1 vector. The
origin of the Ewald sphere with a radius |k 1 | = |k 2 | = 2/ is positioned at E. The ends
of the momentum transfer, Q, and wave vectors k 1 and k 2 , must be on the surface of the
Ewald sphere. If Q = Ghkl , the Ewald sphere passes through one more reciprocal lattice
site and BraggLaue diffraction will occur in the direction of the k 2 vector. If the condition
Q = Ghkl is not satised, as shown in the gure, the BraggLaue diffraction will not occur
for this reection, instead the diffuse scattering will originate from the region of increased
intensity around the (hkl) reciprocal lattice point.

transfer will consist of the -function-like peaks positioned at the reciprocal


lattice sites.
For crystalline materials, it is convenient to write Eq. 1.1 in terms of the
scattering from unit cells that ll the real-space crystal lattice with repeated
structures. The complex scattering amplitudes from each unit cell called
the structure factor, F are summed to determine the total scattering from
the crystal.
2



0
eiQRs F
where F =
f eiQr
(1.2)
I(Q) =
S

Here, the index s labels the unit cell sites with positions R 0s ; are the atoms
in each unit cell with relative positions r ; F is the structure factor; and fn
is the atomic scattering factor of the nth atom of a unit cell.

R. I. Barabash & G. E. Ice

Fig. 1.2. Sketch of crystal lattice plane with one point defect in the position t = 6. Only
c6 = 1, for all other lattice sites ct = 0.

Defects typically redistribute the scattering intensity expected from a


perfect crystal. Defects displace surrounding atoms (unit cells) and change
the scattering factor of the unit cells in which they are embedded. To describe
the distribution of defects, we adopt the random numbers ct (Krivoglaz,
1969, 1996) such that:
ct = 1, for a defect at position t.
ct = 0, for positions without defects.

(1.3)

In the example given in Fig. 1.2, only c6 = 1 and at all other lattice
sites, ct = 0. In the ideal crystal the lattice site positions without any defects
are characterized by the position vector R0S , and positions of defects by the
vector Rt . Each unit cell will be identied by its number, s. If the unit cell
consists of several atoms then the position of atoms inside the unit cell will
be numbered by the index, . The defect positioned on the lattice site, t,
creates a partial displacement, ust , and partial rotation, mst , of the sth unit
cell. The total displacement, us , and/or rotation, ms , of the sth unit cell
are determined from the sum of all displacements/rotations due to all the
defects:


ct ust , and ms =
ct mst
(1.4)
us =
t

Diffraction Analysis of Defects: State of the Art

Similarly, the structure factor, Fs , for sth unit cell in the presence of defects
can be written as:

Fs = F +
ct st
(1.5)
t

Here, F is an average structure factor of the crystal without defects, and st


is a partial structure factor change of the sth cell caused by a defect in the t
location.
As briey aforementioned, in the kinematic approximation a qualitative
classication of defects with respect to their impact on diffraction has
been suggested by Krivoglaz (1969, 1996). All structural defects can be
formally classied into two kinds depending on the exponent value of the
corresponding structural DebyeWaller factor (DWF), Exp[2M]. If the
exponent, M, is nite then the defects are of the rst kind. If M then
they are defects of the second kind (Fig. 1.3).

Fig. 1.3. Impact of the rst and second defect kinds on diffraction: (a, b) -function-like
shape of the intensity positioned in the reciprocal lattice site; (c) peaks of diffuse scattering
and shifted weak -function-like maximum caused by defects of the rst kind; and (d) broad
peak formed in the crystals by defects of the second kind.

R. I. Barabash & G. E. Ice

1. Defects of the rst kind do not change the -function-like shape of the
intensity distribution, I 0 , of regular Bragg reections. They can shift the
intensity peak position, change the integral intensity of the peak due to
the change in average structure factor and introduce diffuse scattering.
The peak intensity is multiplied by a structural DWF. Examples of such
defects are point defects, small dislocation loops and precipitates.
2. Defects of the second kind cause asymmetric broadening of the peak;
their intensity distribution does not have a -function-like shape even
for an innite crystal. Examples of such defects are dislocations,
disclinations/disconnections, disclination dipoles, stacking faults, twins,
boundaries, long cylinder-like chains of impurities, etc.
Generalizing Eq. 1.2 to include displacement/rotations of the unit
cells and changes in the structure factors, the diffracted intensity can be
written as:
 iQ(R 0 R0 )
S
S  eT
e
(1.6)
I = |F |2
s,s

where T is a correlation function. In crystals without defects the correlation


function T = 0. The correlation function is distinct for different kinds of
defects.
In many diffraction experiments the diffracted intensity is also averaged
along certain directions or surfaces. For example, the Debye pattern from
a polycrystalline material averages over the textured crystal orientations of
the polycrystalline grains that intercept the beam (Krivoglaz, 1969, 1996;
Warren, 1990). Single-crystal Laue patterns integrate intensity along the
radial direction of the diffraction vector, Q (Barabash et al., 2003a; Ice and
Barabash, 2007). The integrated intensity in a rocking curve averages along
one of the directions transverse to reciprocal lattice vector, Ghkl (Ice and
Barabash, 2007).
Usually, the intensity in the radial (along the diffraction vector, Q) and
transverse (perpendicular to the diffraction vector) directions in reciprocal
space have distinct responses to distortions caused by crystal defects.
The transverse plane is most sensitive to orientation changes of the
unit cells. This is why diffraction measurements in the transverse plane
give information about the so-called orientation space. The intensity

Diffraction Analysis of Defects: State of the Art

distribution along the radial direction is sensitive to changes that arise


in the lattice parameter from uctuations of strain or chemistry. As a
result, intensity in the radial direction depends on the dilatational strain
and is sensitive to the total dislocation density. Different parameters of
the defect arrangements can be determined from measurements of the
orientation space and radial intensity. The relations between the above
intensity distributions and ensembles of different defects are described in
the next section.
1.2. Classication of Dislocations Structures
Crystals usually contain a large number of dislocations. Dislocations can
form during crystal growth, plastic deformation, materials processing,
etc. During plastic deformation a complicated hierarchical dislocation
ensemble usually forms containing different kinds of straight and/or curved
dislocations, dislocation loops, dipoles, tangles, etc. There is a complicated
correlation in the dislocation arrangement within this ensemble. The most
important kinds of dislocation structures are described below.
Straight dislocations belong to the second kind of defects. Intensity
scattered by crystals with straight dislocations is characterized by broad
peaks. Straight dislocations differ by their type and (Burgers vector and
line direction) and by the number t of the lattice site intercepted by
the dislocation line in the perpendicular plane. The dislocation density is
usually determined as a number of positions intercepted by dislocation
lines per unit area in the plane perpendicular to dislocation lines. Each
tth dislocation line creates long-range partial displacement, ust , and
partial rotation, mst , of the sth unit cell, compared to the crystal without
dislocations. Examples of typical lattice rotations and strain elds in the
vicinity of edge dislocation are shown in Fig. 1.4.
Dislocation strain elds measured with different beam size are averaged
within the area probed by the beam and very small <50 nm probes
are required for reliable characterization of strain elds near individual
dislocations. Each dislocation is characterized by its line direction, ,
and Burgers vector, b. For screw dislocations, b; for edge dislocations,
b. Mixed dislocations have both screw and edge components. The
displacement elds of individual dislocations are described in detail in
Teodosiu (1982) and Nabarro (1987).

R. I. Barabash & G. E. Ice

Fig. 1.4. (a) Rotations in the vicinity of the edge dislocation with Burgers vector
b = 0.2 nm; and (b) strain eld in the vicinity of the edge dislocation measured along
the X-axes using the beamsize of 50 nm (1) and 300 nm (2).

1.2.1. Edge dislocations


A theoretical description of a crystal with a single dislocation running along
the cylinder axes was performed by Wilkens (1970). An edge dislocation
shown in Fig. 1.4 with its dislocation line along the Z-axis and extra plane
of atoms along the YZ plane creates plane strain deformation in the XY
plane with components:


b
xy
1 y
tan
+
ux =
2
x 2(1 )(x2 + y2 )


b
(x2 y2 )
1 2
2
2
ln(x + y ) +
uy =
2 2(1 )
4(1 )(x2 + y2 )

(1.7)

Here is Poissons ratio. Each dislocation creates displacive and rotational


elds in the surrounding media. In each unit cell atoms are displaced relative
to each other, moreover the center of each unit cell is rotated and displaced
relative undistorted position. Partial displacement and rotation elds from

Diffraction Analysis of Defects: State of the Art

Fig. 1.5. (a) Single screw dislocation and (b) statistically stored screw dislocations in
the cylinder in real space; (c) sketch of the disk-shape intensity distribution in reciprocal
space, and relative orientation of the diffraction vector, Q, and Burgers vector, b; and
(d) dependence of the full width at half maximum (FWHM) on the angle , between the
Burgers vector and the diffraction vector.

individual dislocations overlap and form a superimposed rotational and


displacement eld described by Eq. 1.4.
1.2.2. A single screw dislocation in a small cylindrical crystal
A sketch of a single screw dislocation along the cylinder Z-axes is shown in
Fig. 1.5a. In an elastic isotropic media, the displacements around the screw
dislocation are described by the following equation:


b
ys
ys
arctg
arctg
(1.8)
ust =
2
(xs xt )
(xs L2 /4xt )
Here, L is the diameter of a cylinder; xs and ys are coordinates of the sth
lattice site; xt is the distance of the dislocation line from the cylinder axes.
The second term in the brackets relates to the image forces at the cylinder
surface. Wilkens (1970) showed that a single screw dislocation results in

10

R. I. Barabash & G. E. Ice

the formation of high-intensity rings of scattering localized in the plane,


(Ghkl b) = l, where l is an integer. The effective radius of the high-intensity
ring is related to the reection broadening approximately l and increases
with (hkl) of the reection.
The width of the intensity distribution for a single dislocation, l ,
depends linearly on l:
l = 0 (1 + 1.30l),
where 0 is a diffraction line width without dislocations.
The intensity distribution caused by a single dislocation in the cylinder
strongly depends on the boundary conditions at the crystal surface. The
approximation of a single dislocation in the cylinder can be used for crystals
with small dislocation densities, n, up to 106 cm2 . In real crystals the
displacements and rotational elds caused by dislocations interpenetrate
and overlap, and the overall displacements/rotations are found using a
statistical description of dislocation arrangements (Eqs 1.1, 1.2, and 1.4).
1.3. The Inuence of Correlations on Total Dislocation Energy
Correlations between dislocations result in the formation of a selforganized hierarchical structure. During crystal growth, plastic deformation
or materials processing net plastic deformation takes place. The kind and
degree of hierarchical structure depends on the type, magnitude and speed
of the net plastic deformation. Two factors inuence the correlation process
in the dislocation ensemble (Krivoglaz, 1996):
1. An energetic factor, which favors arrangements that decrease the total
elastic energy of the crystal.
2. A kinetic factor, which creates the conditions for dislocation nucleation, multiplication and movement under complicated superimposed
elds from external stresses, other dislocations and obstacles, which pin
the dislocations and block their movement.
Inhomogeneous dislocation arrangements are often approximately
described as dislocation cell-wall structures (Mughrabi, 1983; Jakobsen
et al., 2006; Levine et al., 2006; Mughrabi and Ungar, 2006). In other
cases, inhomogeneous dislocation arrangements are conveniently described

Diffraction Analysis of Defects: State of the Art

11

Fig. 1.6. Correlated dislocation arrangements within boundaries: (a) incidental dislocation
boundaries of the rst type (without misorientation); and (b) small angle boundaries with
alternating lattice rotations.

as misoriented cell blocks. Further evolution of the dislocation ensemble


occurs during annealing of the deformed materials. During annealing, dislocation density and elastic energy decrease, and dislocations often regroup
within the dislocation boundaries (Fig. 1.6), etc. A phenomenological
description of diffraction by deformed materials is often based on the idea
that the crystal consists of crystallites with distinct boundaries with some
random strain distribution within the crystallites. This approach is used for a
harmonic analysis of size and strain in the crystal (Warren and Averbach,
1950, 1952).
Dislocations can form different arrangements in the crystal, such as
statistically stored random dislocations, unpaired or GNDs, dislocation
cells and walls, dislocation loops, dipoles, multipoles, disclinations or
disconnections, IDBs, GNBs and combinations of these defects in various
hierarchical structures.
SSDs cause weakening and broadening of the reections. Broadening
of different reections depends on the so-called contrast factor and
the dislocation density, n. SSDs almost equally affect radial direction
of the diffraction vector and the orientation space.
GNDs, GNBs and dislocation walls mainly broaden orientation space.
Small dislocation loops, dislocation dipoles, pairs and multipoles result
in diffuse scattering.
To relate deformation to specic dislocation arrangements and to
determine the main parameters of the dislocation structure (density,
correlation, GNDs density tensor, etc.) the diffraction for simplied models

12

R. I. Barabash & G. E. Ice

of the main types of dislocation arrangements were considered (Wilkens,


1970; Wilkens et al., 1987; Klimanek, 1993; Krivoglaz, 1996), as described
below.
In highly correlated arrangements the dislocations screen each others
displacement elds. Distortions from individual dislocations can be almost
neglected outside the so-called correlation radius, rc (Wilkens, 1970;
Wilkens et al., 1987; Klimanek, 1993; Krivoglaz, 1996).
The elastic energy, E, of a crystal with randomly distributed dislocations is given by:
 
L
Gnb2
Ln
(1.9)
E
4
b
where G is the shear modulus, and L is the crystal size. For a correlated
dislocation arrangement the crystal size should be substituted by correlation
radius, rc , reducing the energy of the crystal to:
E

Gnb2
c
Ln
4
b

(1.10)

Because of the energy penalty for random dislocations, materials with


random dislocations are thermodynamically unstable and tend to reorganize. However, reorganization depends both on thermodynamic and kinetic
factors. That is why complete screening may not take place. Moreover, not
only pair correlations between dislocations, but also triple and higher-order
correlations become important (Krivoglaz, 1996; Groma, 1998; Szekely
et al, 2001; Schaer et al., 2005; Groma and Szekely, 2006; Kaganer
and Sabelfeld, 2010, 2011). These higher-order correlations can increase
the experimentally observed screening radius, rc . When correlation is
mainly inuenced by the dislocations nucleation and movement rather
than by energy reduction, the groupings of the same sign dislocations cause
antiscreening. Grouping of p dislocations with the same sign increases their
resulting Burgers vector p times.
As a result, correlated dislocation arrangements can be more stable
than the random ones. This leads to the formation of cell-wall structures
(Mughrabi, 1983; Mughrabi and Ungar, 2002; Jakobsen et al., 2006; Levine
et al., 2006) that are organized with a high dislocation density within the
cell walls and a low dislocation density within the cell interiors (Fig. 1.6).

Diffraction Analysis of Defects: State of the Art

13

1.4. Multiple Length Scales in the Evolution of Dislocations


Ensemble
Recent studies have established that unit events of plastic deformation and
the evolution of a dislocation ensemble can take place at different length
scales (Romanov and Vladimirov, 1992). Plastic events taking place at
different length scales occur through different mechanisms and are usually
interrelated. Different models corresponding to distinct mechanisms should
be applied at different length scales. Three length scales are usually
distinguished:
1. Nanoscale: focused on unit microscopic events at the length scale of a
single dislocation at the typical length scale, lu = (130) nm.
2. Mesoscale: where classical, macroscale and nanoscale meet, lmeso =
(0.120)m.
3. Macroscale: at the length scale of the size of structural elements such as
grain or cell block, lmacro = 200m.
The existence of a plasticity length scale has been intensively discussed
in the literature (Stolken and Evans, 1998; Gao et al., 1999; Hutchinson
and Evans, 2000; Needleman, 2000; Uchic et al., 2004; Bei et al.,
2005, 2008a, 2008b; Greer et al., 2005, 2006; Volkert and Lilleodden,
2006; Maas et al., 2007; Zhu and Karihaloo, 2008; Thilly et al., 2009;
Chakravarthy and Curtin, 2011). While the description at the nanoand macroscale is already well developed, the mesoscale and structural
length scales relate to the heterogeneity of plastic deformation and are
still lacking the complete analytical description. Materials physics has
embraced the importance of mesoscale size on materials properties and
typically the smaller the characteristic sample dimension, the stronger the
size-dependent response (Nix and Gao, 1998; Greer et al., 2005, 2006).
Exactly at the mesoscale the qualitative change in materials behavior
takes place. This is one of the motivations for the drive toward materials
studies at mesoscale. However, mechanisms and possible implications of
dimensionality on materials properties are often poorly understood. In
mesoscale physics, this situation is rapidly changing, with the ability to
study deformation with high-resolution electron and X-ray microprobes
and with the recent demonstration of unusual strength in small-dimensioned

14

R. I. Barabash & G. E. Ice

nano- and mesoscale pillars. For example, the inuence of length scale on
deformation was elegantly demonstrated during uniaxial compression of
micron and nanosize gold pillars by Uchic et al. (2004) and Maas et al.
(2007). They argue that size effects are caused by dislocation starvation
hardening, with dislocations leaving the crystal more quickly than they
multiply and leading to the requirement of continual dislocation nucleation
during the course of deformation. This is supported by electron microscopy
studies of deformation that nd fundamentally different behavior in
materials below 500 nm (dislocation free). One explanation is the need
to include a new length scale for plasticity the distance a dislocation
travels before it creates another one or leaves the sample (Pantleon, 2002).
Deformation of crystals smaller than this characteristic size is expected to
be dominated by dislocation starvation while deformation in crystals with
larger dimensions should exhibit conventional dislocation plasticity (Bei
et al., 2005, 2006). One of the unsolved problems is an understanding of
interactions of the plastic events at different length scales. Other problems
relate to the regularities of the collective behavior of the dislocation
ensembles. The new experimental methods described in this book are
specically helpful for studies of materials behavior at the mesoscale.
1.5. Anisotropy of Scattering by Dislocation Ensembles
The scattered intensity distribution due to the presence of dislocations
is sensitive to the kind of dislocations present in the crystal, to their
architecture and correlations (Krivoglaz, 1969, 1996; Krivoglaz et al., 1970;
Wilkens, 1970; Barabash et al., 1976; Mughrabi, 1983; Wilkens et al., 1987;
Ryaboshapka et al., 1990; Klimanek, 1993; Ryaboshapka, 1993; Ungar
et al., 1989, 1999). Around each reciprocal lattice point corresponding to
a reciprocal lattice vector site, Ghkl , the scattered intensity is a function of
the diffraction vector, |Q| = |k2 k1 | = 4/Sin, or of the deviation
vector, q = Q Ghkl . Here, k 1 , k 2 are wave vectors of the incident and the
scattered waves with the directions dened by the unit vectors, k 2 and k 1
(Fig. 1.1). In the presence of dislocations, two directions in reciprocal space
are usually distinguished:
1. The radial direction (q  ) parallel to Ghkl . I(q  )
= I(2) describes the
shape of the radial intensity distribution of the reection, and depends

Diffraction Analysis of Defects: State of the Art

15

mainly on the translation part (strain) of the long-range displacement


elds u = u(R) caused by lattice defects.
2. q transverse to Ghkl . I(q ) depends essentially on the misorientation
part (lattice rotations) of the displacement elds u = u(R) caused by
lattice defects (orientation space).
Characterization of orientation space may be performed by white-beam
techniques (Ice and Larson, 2002; Larson et al., 2002; Barabash et al.,
2003a; Levine, 2006), rocking curve analysis (Hirsch, 1956; Wilkens
et al., 1987; May et al., 1995; Breuer et al., 2000; Barabash, 2001),
reciprocal space mapping (Barabash et al., 1990) and/or orientation
mapping (Laurisden et al., 2000; Margulies et al., 2001; Jakobsen et al.,
2006). Generally, the radial intensity distributions, I(2), and orientation
space distributions, I(q ), have quite different characters correlating with
distinct parameters of the dislocation ensemble.
The intensity distribution, I(q), of X-ray scattering due to dislocations
can be obtained from the expression:
 iq(R 0 R 0 )
S
S  exp [T(R 0 , )]
e
(1.11)
I = F2
s
s,s

Here, F is the average structure factor of the matrix atoms,  = R0s R0s
is the undistorted distance vector between the lattice cells s and s and the
correlation function, T , is:
 
T =
c
[1 exp (iQuss t )]
(1.12)

Here, labels different types of dislocations differentiated by their


Burgers vector and dislocation line directions, c is a dimensionless
quantity that indicates the fraction of lattice sites intercepted with -type
dislocations, and relative displacements, uss t , are the difference between
static displacements of the s and s lattice cells caused by the dislocation
in the tth location. The correlation function, T , depends on the distance
vector, , between the s and s cells and their relative displacements, uss t .
It differs according to the architecture of the dislocation arrangement.
The relative displacements of two scattering lattice cells, s and s , can be
written as (Krivoglaz, 1969; Krivoglaz et al., 1970; Barabash et al., 1976,

16

R. I. Barabash & G. E. Ice

2003a; Ryaboshapka, 1993):


1
uit ujt (R ij )uit + (R ij )2 uit +
2
The correlation function, T , can be expanded with respect to small
displacements. In general, for an arbitrary distribution of paired and
unpaired dislocations, the correlation function, T , has both imaginary and
real parts, T = T1 + T2 . Linear and quadratic terms of the correlation
function in the rst approximation can be written as the following:

c (Rij )(Ghkl uit );
T1 = i
ti

T2 =

c [1 cos[(Rij )(Ghkl uit )]]

(1.13)

ti

The rst term, T1 , is imaginary and describes lattice rotations due to an


unpaired portion of the dislocations population within the probed region.
The real part of the correlation function, T2 , represents the uctuating strain
elds from randomly distributed individual dislocations. It is independent
of whether dislocations are paired or unpaired. As discussed below, if
all dislocations are paired then the imaginary part is equal to zero and
the correlation function, T , is real. If there are unpaired dislocations the
correlation function, T , contains both real and imaginary parts. These two
limiting cases give rise to the two distinct intensity distributions.
1.6. SSDs
At the initial stage of plastic deformation, statistically stored (dipolar)
individual dislocations with equal densities of + and random
dislocations are formed within a crystal. For equal densities of + and
dislocations, there is no macroscopic lattice rotations or bending of
the crystal lattice, T1 = 0 and the correlation function has only real part.
The SSDs broaden reections due to random local uctuations in the unit
cell orientations and d spacing that tend to cancel out over long length
scales. A sketch of the cylinder with multiple screw SSDs parallel to the
cylinder axes is shown in Fig. 1.5b. To calculate the intensity distribution
from the cylinder with multiple dislocations, it is necessary to know T

Diffraction Analysis of Defects: State of the Art

17

within the range of equal to the average distance between the nearest
dislocations. In that case the product Quss t in Eq. 1.11 for the correlation
function can be written as follows:

Quss t
Qj Rj ij Rss Rss
(1.14)
= 0.5
i,j

Here, R st is a distance vector between the dislocation lines and scattering


cell s, and ij are components of the distortion tensor in the local coordinate
system, related to the -type dislocation.
The correlation function for a paired dislocation population has only a
real part T2 (Krivoglaz, 1969; Krivoglaz et al., 1970; Barabash et al., 1976,
2003a; Ryaboshapka, 1993):
T2 =

n(Qb)2 2
L
XY ln
8
|Qb| XY

(1.15)

Here, XY is a projection of the distance vector, , on the XY plane.


For random statistically stored screw dislocations with parallel dislocation lines (Fig. 1.5b) the intensity distribution in the XY plane is Gaussian,
and diffracted intensity is concentrated within the disks perpendicular to
the cylinder Z-axes (Krivoglaz, 1969, 1996; Krivoglaz et al., 1970):


(qx2 + qy2 )
83 N 2
,
f (qz ) exp
I(Q) =
2
2
where 2 = 0.5n(Qb)2 l;

l1

(1.16)

The full width at half maximum (FWHM) of this intensity distribution is

proportional to |Q|, n and Cos (Q b). For Q perpendicular to the


Burgers vector, b, the FWHM of the intensity distribution tends to zero
(Figs 1.6c and d).
1.7. Scattering by IDBs with Fluctuating Paired Dislocations
Stochastic formation of IDBs has been considered by Nabarro (1994) and
Pantleon (2002), amongst others. At early stages of plastic deformation,
glide dislocations trap each other statistically and form thick, loose cell
walls with randomly distributed dislocations within such a wall with almost
equal numbers of + and dislocations. Due to the random positions of

18

R. I. Barabash & G. E. Ice

the dislocations within the wall (with a certain dislocation concentration, c)


and due to the random distances between walls (spread around some mean
distance, Di ) the correlation function, T , depends on the displacement
elds produced by all individual dislocations within the walls and can be
written as:
TIDB =

82 Lh
1
(Q b)2 ni j ij ln
8
(Qb)2

(1.17)

The contrast factor, = (bk , Q, k , pk ), depends on the orientation of


the diffraction vector, plane of the boundary, directions of dislocation lines
and Burgers vectors and their types (edge or screw) for all dislocations
grouped into the walls. Such thick, loose cell walls with irregular dislocation
positions within the wall still create long-range displacement elds and
internal stresses in the cell (Mughrabi, 1983) (Fig. 1.6a). This character
of the strain elds qualitatively determines the distribution of the scattered
intensity. The central part of the reciprocal space volume near a lattice
point is typically well described by a Gaussian distribution with distinct
widths in two directions. If there are only dislocations parallel to Z-axis
grouped in the wall (both + and dislocations) the intensity
distribution near a reciprocal lattice point has a disk shape in the plane
perpendicular to this axis. Along the Z-axis the intensity distribution does
not change compared to the non-deformed crystal. The FWHM in the
XY plane depends on the orientation of the Burgers vector relative to
the diffraction vector,
Q, and the square root
of total dislocation density:
FWHMIDB = (Qb) 0.5nl. Here, l ln( nlL), and L is the effective
size of a coherently (or semi-coherently) scattering fragment. Compared
to the randomly distributed straight individual dislocations the intensity
distribution caused
by walls with very small misorientations is contracted
by a factor = (l 0.5 ln(D/ h)/ l) depending on the ratio D/h between
the average distance between the walls, D, and the average distance between
dislocations in the wall, h. Physically this contraction is due to a reduction
of stored energy with dislocations grouping. The FWHM of the intensity
distribution depends on the total density of dislocations ni grouped within
the wall. For thick, loose walls (IDBs) (Mughrabi, 1983; Ungar et al.,
1984; Wilkens et al., 1987), the width of the wall, W, is usually larger than
the characteristic distance between dislocations within the wall, h, so that

Diffraction Analysis of Defects: State of the Art

19

condition W > h is valid (Fig. 1.6a). This relatively large width of the
IDB causes additional broadening at the tails of the intensity distribution.
Usually, the contribution of the transition near wall regions (within the
width of the wall) is signicant if the distance, D, between the walls and
width of the wall, W, is of the same order D W, and D
h 5. If the
D
opposite conditions are valid, D W , and h > 5, the inuence of the
wall width on scattering is much smaller.
1.8. GNDs
After plastic deformation, geometrically necessary (polar) dislocations as
well as geometrically necessary boundaries may be formed in a crystal.
These dislocations cause not only random deformation, but also strongly
correlated long-range rotations within the crystal, grain or subgrains. The
polar portion of dislocation content is due to the incompatibility of plastic
deformation and to the curvature of the corresponding crystal lattice. It
appears to maintain the continuity of the crystal lattice. That is why at the
nanoscale volume level of the individual dislocations, all dislocations are
geometrically necessary (Nabarro, 2001; Arsenlis et al., 2003). GNDs
are scale-dependent, while SSDs are not. Thats why it should always be
specied over which so-called representative volume element (RVE) the
GNDs density is measured. For example, for micro-Laue diffraction from
a single crystal the RVE is equal to the size of the probed region. If several
grains of the polycrystalline material are probed, there will be distinct RVEs
for each grain. For GNDs and GNBs the distortion tensor, ij , is related to
the GNDs density, n+ , and slip system (direction, , and Burgers vector,
b). Severe deformation causing grain subdivision (Jakobsen et al., 2006)
can change the RVEs within the grain.
GNDs, unpaired or polar dislocations, may be typically arranged in the
following ways (Fig. 1.7):
Unpaired random GNDs.
Unpaired geometrically necessary walls/boundaries.
Unpaired dislocation walls with random dislocations between the walls.
X-ray scattering near Ghkl by crystals with unpaired dislocations exhibits
distinct intensity distributions in radial and transverse directions.

20

R. I. Barabash & G. E. Ice

Fig. 1.7. (a) Geometrically necessary (unpaired) dislocations create macroscopic lattice
rotations of the crystal; (b) mutual orientation of the dislocation line, , Burgers vector, b, and
momentum transfer of the reection, Ghkl ; (c) reections from the bent lattice in reciprocal
space turn into intensity streaks with and being the streak long and short axis correspondingly; and (d) sketch of the dislocation network hierarchy formed by GNDs and IDBs.

1.8.1. Radial direction (q  Ghkl )


For a GNDs density, n+
, belonging to the -slip system, the correlation
function has both real and imaginary terms:

T1 = iC1
n+
(R i b )([Q R ij ] );

T2 = C2 (Qb)2 ln L

(1.18)

Here, C1 , C2 are the contrast factors for imaginary and real parts of the
correlation function, L is the size of the probed subgrain (or the cut-off
radius, whichever is smaller) and is the orientation factor for each
dislocation system. T1 is linear with respect to the density of GNDs, n+
and goes to zero when n+ = 0. From Eq. 1.18, it follows that GNDs have
the same scattering intensity in the radial direction as SSDs, because in the
radial direction the cross product [Q R ij ] = 0 and T1 vanishes.
A random GND distribution results in a Gaussian shape of the intensity
prole in the radial direction with the FWHMrad 2 depending only on

Diffraction Analysis of Defects: State of the Art

the contrast factor C2 and the total dislocations density, n:

2 C2 n

21

(1.19)

If unpaired dislocations are grouped within walls, the intensity prole has a
Lorentzian shape with 2 depending on the average distance, D, between
walls.
In the intermediate case of dislocations distributed both in tilt walls
and throughout the crystal, the radial intensity prole corresponds to a
convolution of a Laue function for coherently scattering regions between
walls and a Gaussian function corresponding to the effect of the randomly
distributed dislocation walls (Krivoglaz et al., 1970; Barabash, 2001; Balzar
et al., 2004; Tomota et al., 2004; Huang et al., 2008, 2010).
1.8.2. Orientation space (qGhkl )
For GNDs (T1 = 0), the scattered intensity distribution becomes much
broader in orientation space and is described by a more complicated
function. For crystals with a random distribution of GNDs, the FWHM
q , the transverse intensity distribution scales linearly with the density of
GNDs, n+ , and the length of the probed region, L:
q Ln+

(1.20)

Due to the dependence of T1 on the orientation of activated GNDs relative


to the diffraction vector (T1 ([Q R ij ] ) (Eq. 1.18), the edge GNDs
with lines parallel to the diffraction vector do not inuence the scattering
intensity in orientation space.
For crystals with tilt dislocation walls, T1 is proportional to the number
of excess dislocations inside each wall and to the total number of GNBs.
For GNDs (Fig. 1.7a), the intensity distribution relates the GND-induced
distortion tensor, ij , to the GNDs density, n+ , within the probed area and
their slip system (dislocation direction, , and Burgers vector, b). The mean
deformation tensor can be written in terms of the antisymmetric Levi-Civita
tensor of third rank, lm , and a GND density tensor of the second rank, .


m
lm
(1.21)
=
xl

22

R. I. Barabash & G. E. Ice

Here the Levi-Civita tensor has elements 123 = 231 = 312 = 1; 213 =
132 = 321 = 1; and all other ijk = 0. The index, , in species
the crystallographic direction of the dislocation line, and indicates the
Burgers vector direction. Based on Nyes denition (1953) for a particular
type of GND, = n+ b . For example, an edge GND slip system
with dislocation density, n+ , a line direction parallel to the [lm] plane
normal, and Burgers vector [lm], the GNDs density tensor may be
written as:
+ 

n ll
n+ lm
n+ l

(1.22)
= n+ ml n+ mm n+ m
n+ l

n+ m

n+ n

Examples of deformation tensors are:


(a) Bending of the crystal by the set of edge GNDs (Fig. 1.7a), = [001],
b = [100] results in the only one non-zero component: zx = n+ b.
From Eqs 1.21 and 1.22 it follows that there are only two non-zero
components of the mean deformation tensor: xy = yx = n+ bx.
This distortion eld is antisymmetric and represents a pure rotation
about the Z-axis that increases with displacement in X (Fig. 1.7a).
The GND-induced overall macroscopic rotation of the lattice within
the probed region in real space results in intensity streaks in reciprocal
space with and being the streak long and short axes, respectively
(Figs 1.7b and c).
(b) Torsion caused by screw GNDs with the density of nGN
Scr , with their
Burgers vector running along Z-axis, is described by antisymmetric
b
GN
tensor with non-zero components: Wxy = 2b znGN
Scr , Wxz = 2 ynScr ,
Wyz = 2b xnGN
Scr , Wmk = Wkm .
1.9. Multiple GND Slip Systems: GND Density Tensors
and Elastic Strain Tensors
1.9.1. Multiple GND slip systems
Under certain deformation conditions multiple slip occurs. For example,
deformation of copper single crystals of <001>- and <111>- orientation

Diffraction Analysis of Defects: State of the Art

23

involves multiple slip (Mughrabi, 1983; Ungar et al., 1984; Wilkens et al.,
1987; Mughrabi and Obst, 2005; Larson et al., 2007, 2008). When different
dislocation systems are simultaneously activated in the probed region, the
correlation functions for radial direction and orientation space, Eqs 1.18
and 1.20, are sums over all activated slip systems, . The individual type
of GND or GNB is identied by their slip plane, N , Burgers vector, b,
and the dislocation line direction, . Under multiple slip the correlation
function, T , depends on the dislocation density tensor, ij , and effective
strain gradient tensor, lmk . The element of a GNDs density tensor
using Nyes notations (Nye, 1953; Arsenlis and Parks, 1999) can be
written as:
ik = i bk (r)

(1.23)

Following this approach (Gao et al., 1999; Huang et al., 2000) in the
framework of strain gradient plasticity, the total GNDs and GNBs density
tensor, ij , relates to the strain gradient tensor, lmk , with the following
equation:
ilm lmk = ik

(1.24)

where ilm is the antisymmetric Levi-Civita tensor (Eq. 1.21). Under


multiple slip, the intensity distribution around each reciprocal lattice point
depends on the GNDs density tensor and effective strain gradient. For
multiple slip systems the dislocation density tensor, ik , gives the sum of the
Burgers vectors of all dislocations whose Burgers vector and line direction
are directed parallel to the xk - and xi -axes (k, i = 1, 2, 3), respectively.
In regions with GNDs activated in multiple slip systems inducing bending
and/or torsion, deformation of the lattice depends both on the screw and
edge portions of GNDs. The geometrically necessary (GN)-related linear
part of the correlation function in orientation space may be written as
 is a distance vector between two diffracting unit
T1 = iAR ss  , where Rss
cells and A is a vector depending on the type of GNDs activated in the
diffracting volume:

n+
Aedge =
Edge (R s b )([Q R ij ] )

24

R. I. Barabash & G. E. Ice

AScrew =

n+
Screw {RS [Q Rij ] 2( RS )( [Q Rij ])}

T1 = i(AScrew + Aedge ) Rss

(1.25)

Here, is a summation over different GND types. It is essential that vector


(AScrew + Aedge ) is orthogonal to Q. For this reason the presence of GNDs
in the probed region does not affect the radial intensity distribution and only
affects the orientation space. Orientation space is much more sensitive to the
presence of GNDs than to the same amount of SSDs or even total dislocation

density. If the GNDs density, n+ , satises the condition n+ L > 0.1 n,


with L being the length of the probed volume along the Burgers vector
direction, GNDs broaden orientation space. For example, this condition
holds if L = 10 m, n = 1012 cm2 , n+ = 109 cm2 . This means that
even when the polar portion of the dislocation population is only about 0.1%
of the total dislocation population, n+ = 0.001n, the intensity distribution
in the orientation space, q , becomes very broad and depends mainly on
GND density and length of the probed region:
q bQLn+

(1.26)

In the radial direction, intensity distribution remains narrow and depends


only on total dislocation density, n. That is why GNDs cannot be detected
by powder diffraction. This estimate agrees with modeling (Arsenlis et al.,
2003) that demonstrates that small (0.01 and less) GND portions in the
total dislocation population induces essential lattice curvature. This is also
why GNDs cause broadening of the rocking curves and streaking of the
Laue spots (see Section 1.15).
1.9.2. Local elastic strain
At the early stages of loading, the lattice can be strained only elastically.
With single crystal or single grain method the local strain tensor of the
unit cell can be measured directly. For example, consider a unit cell with
lattice parameters {aI } and {I } in the Cartesian coordinate system {uI }.
Assume a1 vector of the unstrained unit cell is parallel to the u1 axis, a2
vector is parallel to the u1 u2 plane, and u3 is perpendicular to the u1 u2

Diffraction Analysis of Defects: State of the Art

25

plane (Patterson, 1959; Rollett, 1965; Busing and Levy, 1967; Chung and
Ice, 1999). Any vector in the Cartesian coordinates vu can be transformed
to a vector with unit cell coordinates v by vu = Av, where:

a1

A = 0

a2 cos 3
a2 sin 3

a3 cos 2

a3 sin 2 cos 1

(1.27)

1/b3

Here the {bi } and {i } are the sets of reciprocal lattice parameters and angles
for the unit cell, respectively.
A position of any cell in a crystal can change either due to rigid body
translation of the crystal, distortion of the crystal lattice or both of these
(Noyan and Cohen, 1987). Comparison of the measured unit cell parameters
with the undistorted unit cell ones allows the determination of the strain
of the unit cell. The matrix, AMeas , converts a measured position vector v
into the measured crystal Cartesian coordinates. Let A0 be the matrix for
an unstrained unit cell that converts v into the Cartesian reference frame of
the measured (strained) crystal. Let the origins of these two unit cell vectors
coincide. The transformation matrix, T, relates the matrices AMeas and A0
and maps, A0 = T AMeas . Both distortion and rotation are included in the
transformation matrix. For unstrained crystals T = I. The strain tensor, ij ,
in the crystal coordinate reference system can be written as:
ij = (T ij + T ji )/2 = I ij

(1.28)

Equation 1.28 takes into account that the strain tensor is symmetric,
ij = ji , and removes the anti-symmetric rotation terms (Noyan and Cohen,
1987). The strain tensor components in the crystal reference frame can
be rewritten in sample reference coordinates if the rotation matrix, R, is
known:
Sample

ij

= R ij R 1

The strain tensor components (Eq. 1.28) contain both a hydrostatic strain
(dilatation) and a shear (deviatoric) strain term. These two components can
be written separately subtracting the isostatic strain, /3 or the mean strain

26

R. I. Barabash & G. E. Ice

components as follows:


12

11
3


ij =
22
21
3

31
32
ij = ij + D
ij

13

23
+ 0


33
0
3

0

3
0


3
(1.29)

where = 11 + 22 + 33
Deviatoric strain tensor components can be measured with micro-Laue
diffraction using polychromatic radiation (see Section 1.15), while dilatational strain is measured with monochromatic radiation in different diffraction methods (Chapters 4, 5 and 6).
Non-linear least-square tting of the peaks is typically used to
determine the local crystallographic orientation and strain. Stephen (1999)
proposes a phenomenological model of strain-induced anisotropic peak
broadening in powder diffraction. However, his method is not valid for
strain measurements in orientation space.
1.10. GNBs
At a higher structural level, dislocations can organize into strongly correlated arrangements including walls and sub-boundaries. In real crystals
individual GNDs tend to group into walls to reduce the stored energy
(Krivoglaz, 1969, 1996; Mughrabi, 1983; Ungar et al., 1984; Nabarro,
1987; Hansen, 2001; Pantleon, 2002). Consider the structure with geometrically necessary randomly distributed tilt dislocation boundaries (GNBs)
(Fig. 1.6b). Each wall provides a rotation between two neighboring mosaic
blocks around the line of the wall (Fig. 1.6b). The unit vector parallel
to the rotation axis of each wall coincides with the unit vector, along
the dislocation lines in the case of a tilt boundary. We consider pure tilt
boundaries formed by equidistant edge GNDs (so-called thin walls).
These boundaries in the innite crystal do not produce long-range elastic
strain but generate subgrain rotations. The remaining stresses in the GNB
region are periodic with a period, b = b csc /2. These stresses are only
appreciable at distances within b from the GNB (Nabarro, 1987). If h

Diffraction Analysis of Defects: State of the Art

27

is the distance between these dislocations in the wall, one can consider
the boundary as a single defect producing the local rotation eld, . The
misorientation angle across the boundary is dened by the equation:

b
= 2 sin
h
2

(1.30)

where hb for small angle GNBs. If GNBs are not correlated and the
number of + and GNBs probed by the beam is equal, the broadening
of the transverse intensity distribution increases with the misorientation
angle through each GNB and with the total number of GNBs probed by the
beam.
1.11. Cumulative GNBs Misorientation
To characterize this model quantitatively, we dene the average distance,
D, between GNBs, and write the number of GNBs per unit length as D1 .
The total density of GNDs grouped in the GNBs is denoted by:
1
(1.31)
Dh
We further dene an X-axis perpendicular to the plane of the wall, and
a Z-axis parallel to the direction of dislocation lines in the wall. For this
coordinate system (Figs 1.7a, b, c), two non-zero components of the mean
deformation tensor are equal: xy = yx = x/D. This means that the
deformation tensor results in pure rotations about the Z-axis. The rotation
increases with displacement x. The number of GNBs per unit length and the
length of the probed region along the Burgers vector direction determines
the total measured rotation.
Generally, real crystals contain a hierarchy of dislocation structures.
Some of the GNDs are distributed randomly, and the rest may form
different non-random arrangements: IDBs and GNBs. It has been found that
dislocation boundaries evolve within a regular pattern of grain subdivision
on two scales. The smaller scale is related to usual cell boundaries socalled IDBs. The larger scale is related to long and continuous so-called
GNBs. Usually, GNBs separate volume elements which deform by different
slip system modes with different strain amplitudes (Hughes et al., 1997;
Hughes and Hansen, 2000; Hansen, 2001; Pantleon, 2002). Typically there
n+ =

28

R. I. Barabash & G. E. Ice


Disloca
Line
t

Intensity
Disk

n
Dipole L
d
Shoulder

Rst

Ghkl

rd
Real Space

Reciprocal
Space

(a)

(b)

(c)

(d)

Fig. 1.8. (a) Randomly distributed dislocation dipoles on real space; (b) disks of intensity
formed at each reciprocal lattice point are oriented perpendicular to the direction; (c)
polarized dislocation walls composed by grouping of dislocation dipoles within the walls;
(d) dislocation multipoles formed by correlated groupings of three parallel dislocations.

are many IDBs separating ordinary dislocation cells between two cell-block
boundaries formed by GNBs (Fig. 1.7d).
1.12. Dislocation Pairs and Dipoles
Under certain conditions, plastic deformation results in the formation of a
large number of pairs of parallel straight dislocations with opposite Burgers
vectors (Krivoglaz, 1996; Kaganer and Sabelfeld, 2010). These dislocation
pairs are termed dislocation dipoles. Within each dipole, dislocations are
located at a relatively small distance, the dipole shoulder, Ld , compared
to the distance between the dipoles, rd , so that Ld  rd (Fig. 1.8a).
The center point between the dislocations within the dipole in the plane
perpendicular to their lines is referred to as t. The presence of dislocation
dipoles essentially changes the diffracted intensity distribution. In the sth
unit cell, located at large distance, Rst , compared to the dipole shoulder,
Rst Ld , the distortions from opposite sign dislocations in the dipole
mutually compensate each other, and the resulting displacement of the
sth cell from its ideal position linearly decreases with distance from the
dipole:
Qust =

QbLd

2Rdt

R st Ld

(1.32)

Here the function depends on the specic type of the dislocation dipole.
The type of the dislocation dipole is characterized by the unit vector parallel

Diffraction Analysis of Defects: State of the Art

29

to the dislocation lines in the pair, , and vector, Ld n, connecting the


dislocations of the pair perpendicular to their line direction . The character
of the displacement eld around the dislocation dipole is similar to the eld
from a point defect in the two-dimensional (2D) crystal. If concentration
of the dipoles is not very high so that static DebyeWaller factor exponent,
2M  1, the dipoles affect the scattered intensity as rst-kind defects do:
they decrease the intensity and cause diffuse scattering, but do not broaden
the Bragg reections. The 2D distortion elds around dislocation dipoles
cause the formation of intensity disks perpendicular to the direction at
all reciprocal lattice points (Fig. 1.8b). Different types of dipoles result
in differently oriented intensity disks. Dislocation dipoles can group into
the walls. Models of so-called polarized dislocation walls composed
of dislocation dipoles with opposite sign of dislocations on each side
of the wall (Fig. 1.8c) were proposed by Mughrabi and Ungar (2002),
and Kocks et al. (1980) to explain the compatible plastic deformation
of soft and hard regions in the composite model of plastic deformation
(Mughrabi, 1983). Higher-order correlation between parallel dislocations
can result in the formation of dislocation multipoles (Fig. 1.8d). Dislocation
dipoles/multipoles and polarized dislocation congurations result in the
asymmetric intensity distribution along the radial direction (Groma et al.,
1988; Ungar et al., 1989; Krivoglaz, 1996).

1.13. Dislocation Loops


Dislocation loops appear under two different kinds of external elds:
(i) after radiation as a result of vacancy condensation in the plane of the disk
and its further collapse and (ii) plastic deformation (especially severe plastic
deformation). Small and large dislocation loops diffract X-rays differently.
Diffraction by dislocation loops essentially depends on their radius, R0 ,
and on the number of loops per unit volume, NL .
1.13.1. Small dislocation loops
Distortion elds around a small loop depend on its Burgers vector, b, radius
of the loop, R0 , and normal to the plane of the loop, n (Fig. 1.9a). At large
distances from the loop, Rst R0 , the displacement eld from the loop

30

R. I. Barabash & G. E. Ice

Fig. 1.9. (a) Dislocation loops in real space; possible shapes of diffuse scattering in
reciprocal space from small loops; (b) single drop; (c) double drop.

follows Coulombs law:


ust =

bR20
R2st

(1.33)

where the orientation factor, |C|, depends on the mutual orientation between
the Burgers vector, loop plane normal and radius vector to the specic lattice
point R st . The strongly distorted volume around the loop is of the order of
R20 . Close to the reciprocal lattice point for q  R 1QB , strong so-called
0
Huang scattering is observed (Barabash et al., 2012), given by:
IH = NL |f |2 e2M

b2 R40 Q2
H
2 q2

(1.34)

The factor H depends only on the relative orientation of the loops plane and
Burgers vector to the diffraction vector and direction in reciprocal space. It
characterizes the angular dependence of the intensity. For example, for facecentered cubic (FCC) crystals with dislocation loops in the {111} planes
and <110> Burgers vectors, factor H for (h00) type reections is equal to
(Trinkaus, 1972; Krivoglaz, 1996):


2
c12
c12
2
H=
for q running along [100] direction.
+3 2
1+2
9
c11
c11
H=

2
3

for q running along [010] or [011] direction.

(1.35)

Scattered intensity by crystals with small dislocation loops has an oval


shape around reciprocal lattice points (Fig. 1.9b). Iso-intensity lines do

Diffraction Analysis of Defects: State of the Art

31

not pass through reciprocal lattice points but surround them. For some
specic types of dislocation loops for example, prismatic loops with b
<100> the scattered intensity has a double-drop shape around (h00) and
(hh0) reections at small q (Fig. 1.9c). However, even for these reections,
at larger q the intensity shape changes to a single drop (Fig. 1.9b).
Scattered diffuse intensity strongly depends on the loop radius as R40 . Each
dislocation loop increases diffuse intensity (R0 /d)4 times more than a
single vacancy or interstitial. That is why grouping of vacancies/interstitials
into NL loops increases the diffuse intensity. The sign of the product b n
determines the direction of the shift of the intensity compared to the ideal
Bragg position but does not inuence the amplitude of the diffuse intensity.
The shift of the intensity maximum is proportionate to NL R20 . That is why
measuring the shift of the diffuse maxima NL R20 , static DebyeWaller
factor M NL R20 and the integrated diffuse intensity NL R40 allows one
to determine the type of the loop interstitial or vacancy and separately
identify the density of loops (number of loops per unit volume), NL , and
their radius, R0 .
For large hkl Miller indices of the reection there is an interval of
q where the main input to intensity is coming from the coupling of
interference from the two strongly distorted regions at large distances from
the loop and the diffuse intensity follows the dependence ID 1/q4
with oscillating terms. Oscillations are most pronounced for the product,
ID q4 . If R0  rL and Qb 1, the diffracted intensity averaged over the
orientation of different crystallites (for example, in powder diffraction) has
a Lorentz shape. Full width at half maximum, 2, and shift of the maxima,
0 ), for prismatic interstitial loops in {111} planes with isotropic
2(m m
elastic moduli are described by the following dependencies (Barabash and
Krivoglaz, 1978, 1982; Krivoglaz, 1996):
2
= 2.9(h00),
NL bR20 tg
0)
2(m m
= 0.86(h00),
NL bR20 tg

5.01(hh0),

5.2(hhh)
(1.36)

0.92(hh0),

0.87(hhh)

At the tails of the line prole the intensity decreases as 1/q3 .

32

R. I. Barabash & G. E. Ice

Numerical calculations of diffuse scattering from interstitial dislocation


loops on {111} planes in aluminum (Larson and Schmatz, 1980) show
large asymmetries for parallel and antiparallel q directions compared to the
diffraction vector. Experimental and computational studies of vacancy and
interstitial loops in irradiated copper (Larson andYoung, 1987) demonstrate
the possibility of obtaining the detailed information about the kind of loops
and their size distribution with diffuse scattering analysis.
1.13.2. Large dislocation loops
When dislocation loop radius, R0 , and the number of loops per unit volume,
NL , become large, the static DebyeWaller factor exponent 2M 1, and
R0 rL . This can happen because dislocation loops are located in different
atomic planes (Fig. 1.9a). For this case the following condition is valid:
2M (Qb)2/2 1
The diffuse intensity is coming from distorted regions of the crystal located
within the distance R0 from the dislocation loop. Within relatively
small distances from the loop (R0 ), the strain eld near the loop is
similar to the one near a straight dislocation. Diffracted intensity forms
a broad peak typical for the second kind of defects. The shape of the broad
intensity distribution depends on the ratio between the loop radius and
the average distance between the loops, rL (Fig. 1.9a). The shift of the
intensity maximum is much smaller than its width. Integral half width,
2qint , is described by the equation:

2
2qint = 0.88bQ NL R0
(1.37)
The distortions caused by the density, NL , of large dislocation loops with
R0 loop radius are equivalent to the distortions from dislocation density,
n, being of the order of 3NL R0 . Parameter depends on the type
of dislocation loop. Averaging over possible crystallite orientation with
large dislocation loops, for example for powder diffraction, results in the
following equation for integral half width:
2
=
D

1
Nl R0 (Qb)2
2

(1.38)

Diffraction Analysis of Defects: State of the Art

33

It is important to note that variation of the half width due to the presence
of large dislocation loops is similar to that for random straight dislocations
with slightly different coefcients due to correlation between different parts
of the dislocation loop. At the tails, the diffracted intensity decreases as
ID q2 , similar to the case of straight dislocations.
1.14. Stacking Faults, Twins and Partial Dislocations
Stacking faults, twins and partial dislocations are often formed during
plastic deformation and/or crystal growth. They change the order of the
atomic planes and cause an abrupt phase change at certain lattice planes
resulting in a broadening of those (hkl) reections for which the phase
change is not multiple of 2. They inuence diffraction similar to antiphase
boundaries separating ordered domains in ordered alloys. Stacking faults
displace atoms at certain vectors and rotate parts of the crystal at certain
angles (Fig. 1.10). For example, in FCC crystals the {111} most closely
packed planes are also the dislocation slip planes and the twinning planes.
Let us indicate the possible positions of the closely packed planes as
A, B, C. In FCC crystals the two neighboring planes cannot have same
indices like AA. The undeformed sequence of layers for FCC structures is
ABCABCABC. . . Twin formation corresponds to the rotation by 180 in
{111} plane, or mirror image of the lattice planes sequence in {111} plane.

Fig. 1.10. (a, b) Sketch of the two different orientations of twins in real space; (c) rodlike intensity distribution corresponding to twin 1 in reciprocal space; (d) multiple rod-like
intensity distribution corresponding to simultaneous presence of twins 1 and 2.

34

R. I. Barabash & G. E. Ice

For example, in the sequence ABCAB:C:BACBAC the mirror symmetry


is dened by two colons. The displacement vector can be written as
u = d 1 +d63 2d 2 . This vector is located in the {111} planes and d1 , d2 and
d3 are running along the edges of the FCC cubic cell. If the two s and
s scattering cells are located on different sides of the stacking fault or
twin, they will be displaced relative to each other by usign(R0s R0s ). At
a relatively small concentration, c, of stacking faults or twins, they cause
the formation of rods in reciprocal space with -like intensity distribution
in the plane parallel to the stacking fault plane, (q1 ). Along the axes
perpendicular to the defect plane, the intensity distribution broadens and
relates to the Lorenz function (Warren, 1990; Krivoglaz, 1996):


1cos
2
(q
)
1
D
1
c
Nf
I(Q) = 82
;
=
2
2
1
2
(q + D Sin) + D (1 Cos)
D
d
(1.39)
where D is the average distance between the stacking faults, d is the distance
between the possible neighboring stacking faults position perpendicular to
their planes, = Q u. Integral width, q, of this intensity prole and
maximum shift, qm , relate to the following equations:
1

(1 Cos); qm = Sin.
(1.40)
D
D
They depend on the average distance, D, between the stacking faults and
on the (hkl) type of the reection.
For example, in FCC crystals with stacking faults parallel to {111}
planes:

d1 + d3 2d2
hi h2
; = (h1 + h2 h2 )/3 =
u=
6
3
q=

(1.41)
In FCC lattices constructive intensity interference takes place only for hi

being all odd or all even. That is why for i hi = 3m, where m is an
integer, parameter 2 and stacking faults do not inuence diffracted

intensity of these reections. If i hi = 3m 1, then differs from 2

by 2
i hi = 3m + 1,
3 and rod-like intensity maxima will appear. For
the reection maximum will be shifted away from the zero point of the

Diffraction Analysis of Defects: State of the Art

35


reciprocal lattice, and for i hi = 3m 1, the maximum will be shifted
towards it. If there are several types of stacking faults, then several intensity
rods perpendicular to the stacking faults planes will appear near reections
with:

hi = 3m 1
(1.42)
i

In the powder diffraction method the intensity is averaged over


the crystallite orientations. Intensity maxima for {222} and {200} will
be shifted in the opposite direction for FCC crystals. This signature can be
used to differentiate stacking faults from other dislocation congurations.
More details can be found in Warren (1990) and Krivoglaz (1996).
1.15. GNDs and Strain Imaging with Micro-Laue Diffraction
at Mesoscale
Micro-Laue diffraction is best suited for probing deviatoric strain tensor
components, the rotational part of plastic deformation, and in particular the
GNDs density tensor, the types of activated GNDs and related curvature
tensor in the probed region (Sections 1.8 and 1.9). Several experimental
setups and instrumentation for micro-Laue diffraction are described in detail
in Chapters 2 to 5. Below we describe the specics of the micro-Laue
image formation and ways to extract quantitative information about GND
characteristics related to the rotational deformation part. The description
of the code for 3D depth-resolved reconstruction analysis can be found in
Chapter 10.
1.15.1. Imaging GNDs with micro-Laue
diffractionmisorientation vector
In Laue diffraction, the incident beam and scattered beam directions dene
a line in reciprocal space (Ice and Larson, 2002; Larson et al., 2002).The
radial position along this line is determined by the wavelength of the
scattered radiation. For example, reciprocal lattice points (00h), (002h),
(003h), etc., are scattered towards the same pixel on a charge-coupled
device (CCD) but lie at different positions radially in reciprocal space. To
analyze the white-beam intensity distribution from a deformed grain, we

36

R. I. Barabash & G. E. Ice

Fig. 1.11. (a) Misorientation vector m in micro-Laue diffraction shows the difference
between the unit vector parallel to exact momentum transfer for (hkl) reection and an
arbitrary direction in its vicinity; (b) a probed volume of reciprocal space (shaded in gray)
is bounded by the short and long wavelength limits of Ewalds sphere.

introduce unit vectors in each direction of scattering, k = k/|k|. We dene


a special misorientation vector, m, near a Bragg reection, m = k k hkl
(Barabash et al., 1976, 2003a). The misorientation vector, m, shows the
difference between the unit vector parallel to the exact momentum transfer
for (hkl) reection and an arbitrary direction in its vicinity (Fig. 1.11a).
This is a crucial difference between the general intensity distribution in
reciprocal space and the intensity distribution of a scattered polychromatic
(white) beam. For example, with respect to the (1, 0, 0) Bragg reection,
the distinct reciprocal space points (1, 0, 0.1) and (2, 0, 0.2) have the
same m.
The region of high intensity, IL (m), around each vector k hkl in Laue
diffraction (Fig. 1.11) is a function of a misorientation vector. Within this
approximation, the resulting intensity distribution by the deformed crystal
in the white micro-Laue diffraction can be written as follows:
IL (m)
=A


I0 (k)I(q)dk;

q = |k hkl |m + (|k hkl |mrad + kG/|k hkl |)

(1.43)

Diffraction Analysis of Defects: State of the Art

37

Here, |k hkl | is the radius of the Ewald sphere that passes through Ghkl. ,
A is a constant, and mrad and m are components of the misorientation
vector, m, along and perpendicular to the direction of the scattered beam,
k hkl , respectively. The probed volume of reciprocal space is bounded by the
short and long wavelength limits of the Ewalds sphere, shaded in gray in
Fig. 1.11b. The misorientation vector, m, characterizes the Laue intensity,
I L (m), assuming the spectral distribution varies smoothly near the energy
corresponding to the Bragg reection. The second part of Eq. 1.43 is valid
when |k k0 |  1. In the rst approximation, the misorientation vector, m, is
perpendicular to k hkl (Fig. 1.11a).
Using GND density tensor components as input parameters, the
intensity distribution of each Laue spot can be calculated numerically from
Eq. 1.43.
There is a small 1mrad convergence in the incident beam due to
specics of micro-focusing optics (Chapter 2). This convergence does not
practically affect the length of the streak, but for the narrow direction of the
streak it should be taken into account. I L (m) should therefore be convoluted
with the experimental angular resolution function.
1.15.2. Origin and analysis of streaks at the Laue patterns
In Sections 1.8 and 1.9 it was shown that at the mesoscale GNDs
cause lattice curvature. Micro-Laue diffraction provides direct information
about the lattice curvature in the probed region. To describe the intensity
redistribution near a Laue spot due to a particular set of GNDs, the mutual
orientation of two planes are important (Figs 1.7b, c):
The plane perpendicular to the direction of dislocation line, , in real
space.
The plane perpendicular to the direction of reciprocal lattice vector,
Ghkl .
The line of intersection of these two planes naturally denes the direction
of the streak axis, :
= Ghkl /| Ghkl |

(1.44)

The second axis, , is perpendicular to the axis and to the reciprocal


lattice vector Ghkl . The intensity of a Laue spot is strongly elongated in

38

R. I. Barabash & G. E. Ice

the direction (Fig. 1.7). The FWHM in the direction FWHM m


depends on the orientation and the number of GNDs in the probed volume.
For multiple GND slip systems operating within the probed region, the
streak direction of the Laue spot depends on the directions of eigenvectors
of the dislocation density tensor.
The FWHM m in the transverse direction, , depends on the
total number of all dislocations (GNDs and statistically stored) and usually
m m . The orientation and broadening of Laue streak depends on
the density and slip systems of activated GNDs described by the GNDs
density tensor (Figs 1.7 and 1.12).
For example, for an arbitrary direction of the diffraction vector (but not
parallel to the dislocation line direction, ), and for one GND slip system
the intensity distribution of each Laue spot (1.43) simplies and can be

Fig. 1.12. Distinct streaking of the Laue spots for an FCC crystal in cubic axes with 12
different GND slip systems. Simulation performed for the back Laue geometry where the
incident beam is perpendicular to the sample surface. One frame is indexed for convenience.

Diffraction Analysis of Defects: State of the Art

39

written as (Barabash et al., 1976, 2009):


IL (m) =

AI0 (k0 )
Q /k0



1

GN

+ d2 (m )/( ) + (m ))

 d (
1




 ( GN + d2 (m )/( ) + (m ))

d1


d3 (m )
Exp 2
(1.45)
( )

Here, (x) is a cumulative probability distribution, and the width of the


Laue intensity distribution depends on the contrast factors GN and :
!

bQ
nl
GN
+ Q
2

= n b L 1 ( ) ; =
(1.46)
k0
8(1 )
Here, n+ is the GNDs density for an activated slip system; n is the
total (GN + SS) dislocation density; is a unit vector along the diffraction
vector Q; L is the size of the probed region in the direction of the Burgers
vector, b; parameter l is l 1; Q is a projection of the diffraction vector Q
along the dislocation line direction; d1 , d2 , d3 are parameters depending on
the elastic constants and on the mutual orientation between the dislocation
slip system and momentum transfer for each Laue spot. Using Eqs 1.45
and 1.46, the Laue pattern from crystals with specic GND distribution can
be modeled and compared to the experimental ones. An example modeling
for 12-edge GND slip systems typical for an FCC crystal is presented in
Fig. 1.12 for the geometry where the incident beam is perpendicular to the
sample surface. As was noted in Section 1.10, GNDs broaden orientation
space and cause streaking of the Laue spots when their portion in the total
dislocation population is more than 1%. In the simulations it was assumed
that the portion of GNDs is 5% of the total dislocation density. One
frame is indexed for convenience. It is seen that Laue intensity streaks
rotate simultaneously as the GNDs slip system changes. It is not only
the orientation of the streaks, but also the length and width of each Laue
streak which changes corresponding to their contrast factors (Eq. 1.46).
Some Laue spots are practically not streaked, while others demonstrate

40

R. I. Barabash & G. E. Ice

long streaks. Compare, for example, the (1, 1, 7), (1, 1, 7) and (1, 1, 7)
Laue spots at the indexed frame in Fig. 1.12. Multiple examples of the
streak analysis of Laue patterns can be found in Barabash et al. (2001,
2003a, 2003b, 2004, 2009), Ice and Barabash (2007), Ohashi et al. (2009)
and Wang et al. (2011).
1.15.3. Insensitivity of micro-Laue diffraction to GNDs in specic
directions
Condition 1.44 and contrast factor Eq. 1.46 dene the case when plastic
deformation will not be visible with micro-Laue diffraction: if GNDinduced overall rotation axes are parallel to the momentum transfer Ghkl ,
the corresponding Laue spots will not be affected by plastic deformation.
Insensitivity of micro-Laue diffraction to deformation in specic directions
was experimentally conrmed byYang et al., (2003) with 3D depth-resolved
measurements. The Cu micro-compression study (Chapter 5) demonstrates
an example when micro-Laue diffraction is not sensitive to specic kinds
of deformation.
A comparison between the experimental and the modeled Laue pattern
for hexagonal crystal is shown in Fig. 1.13 for the measurement geometry
described in Chapter 2 with the incident beam intercepting the sample
surface at 45 . The comparison between the experimental and the modeled
Laue patterns (Figs 1.13a, b) demonstrates that the best t with experiment

Fig. 1.13. Comparison between the (a) experimental and (b) modeled Laue pattern
demonstrates that (1101)[1213] GND slip system is activated in a deformed grain of a
Ti polycrystal; (c) sketch of the active slip system orientation in the unit cell.

Diffraction Analysis of Defects: State of the Art

41

corresponds to the (1101) [1213] GND slip system activated in a deformed


grain of a Ti polycrystal (Wang et al., 2011).
1.15.4. Why do Laue spots split?
If the angular misorientation between GNBs is sufciently large, the
reection becomes discontinuous. To understand this dependence, we
assume that along the penetration depth, L, the X-ray beam intersects
several cell blocks with average size Dcb . Each cell block contributes to
the diffraction. The number of such contributions is L/Dcb . Each boundary
produces an average misorientation, . When misorientation through the
boundary exceeds the average FWHMcb mcb of the Laue spot, the Laue
spot splits (Figs 1.14c, f). If almost all dislocation walls are unpaired, the
following criterion can be used:

K=


mcb


(1.47)

Fig. 1.14. Partial grouping of GNDs within the dislocation boundaries splits Laue spot.
Simulated results for the Laue streak and intensity prole along the streak. The portion of
dislocations grouped within the walls increases from (a, d) 0.5% to (b, e) 50% and (c, f)
90%. Different numbers of walls are modeled: (a, b, c) three walls and (d, e, f) seven walls.
Contour maps of the streak are shown on top of the streak intensity prole for each gure.

42

R. I. Barabash & G. E. Ice

If K <1, the Laue spot intensity distribution is continuous (Fig. 1.13). If


K > 1 the Laue spot splits into separate spots and the intensity prole along
the Laue streak consists of several spikes (Figs 1.14b, c, e, f).
1.15.5. Partial grouping of GNDs within boundaries
Unpaired boundary and unpaired individual dislocations can have both
similar and dissimilar effects on the Laue pattern. Both scales of dislocation
organization result in the streaking of Laue reections. The same orientation
of GNDs and GNBs correspond to the same direction of streaking.
However, GNDs and GNBs cause continuous and discontinuous intensity
distributions along the streak, respectively. The FWHM in the narrow
direction of the streak, m , is most strongly inuenced by all individual
dislocations and can, in principle, be used to separate boundary dislocations
from those located inside the cell. This can be done by ltering the incident
X-ray energy with a monochromator.
Simulations of the intensity diffracted by crystals with different
numbers of cell blocks help understand the main features of white beam
scattering from crystals with various dislocation arrangements (Fig. 1.14).
The results of simulated scattering by general dislocation structures
induced both by boundary dislocations and by dislocations in the interior
(when dislocation slip systems in the boundary differ from the randomly
distributed ones in the inner regions) are illustrated in Fig. 1.14. To simplify
the interpretation of the patterns, the total number of unpaired dislocations
was kept constant. A total dislocation density of n = 1012 cm2 was
chosen. This value is typical of highly deformed crystals. Firstly, only
0.5% was assumed grouped within the walls. The deformation elds from
random GNDs superimpose and the intensity is calculated from Eq. 1.43.
The Laue spot streaks along the -axis with a typical at-top shape
(Figs 1.14a, d). As dislocations are removed from the cell interior and added
to the dislocation walls, a correlated misorientation develops between the
neighboring parts of the crystal.
When the dislocation walls are well developed, so that the distance
between dislocations within the wall, h, is much shorter than the distance
between the walls, D, (h  D) (corresponding to the case when more
than 50% of all dislocations are grouped within the walls), the intensity
distribution becomes discontinuous (Figs 1.14c, f). Such walls produce

Diffraction Analysis of Defects: State of the Art

43

sharp rotations of the cell blocks with an abrupt rotational phase variation.
However, when the walls are not well developed as in the case of 0.5% of
the total number of dislocations grouping within the walls (Figs 1.14a, d),
the condition, h
= D, is true. Here each cell block still contains many
randomly distributed dislocations and has a large mcb . This results in
overlapping of the diffracted intensity and formation of an intensity
cloud with spikes of intensity on top of it. With the increase of the
number of boundaries the cloud can be observed even when 50% of
dislocations are grouped within the walls (Fig. 1.14e). Such intensity
clouds are observed by Levine et al. (2006) using micro-Laue diffraction
and Jakobsen et al. (2006) using monochromatic 3D X-ray diffraction
(3DXRD) (see Chapter 9). The total length of the streak, m , is large.
With an increasingly large number of walls, the scattering from each
cell block blends together. For higher-resolution measurements, a larger
number of separate spots can be detected within a streak as observed
experimentally. This simulation illustrates that observed splitting of a white
beam reection (hkl) into spots depends on the following parameters:
density of unpaired dislocations inside the wall, misorientation angle
created by the dislocation sub-boundary, number of walls in the irradiated
volume, size of a cell block, size of scattering volume and experimental
resolution function. Based on the above analysis we can make the following
conclusions:
The Laue pattern is mainly affected by the GNDs and/or GNBs(related
to correlated deformations of the lattice), if n+ L 0.1 nl. For
example, this condition is valid if L = 1 m, n = 1011 cm2 , n+ = n.
In this case m m .
When statistically stored dislocations dominate in the dislocation
ensemble, m m , the Laue spots are sharp and almost isotropic.
The intensity of the Laue spot mainly relates to the total dislocation
density.
Micro-Laue diffraction is not sensitive to lattice rotation and plastic
deformation if the axis of an overall mesoscale rotation is parallel to
momentum transfer.
When the Laue intensity splits along the streak (K > 1), the dislocation
structure must be predominantly characterized by GNBs.

44

R. I. Barabash & G. E. Ice

A detailed 3D Laue intensity distribution is necessary to adequately


determine the parameters of the dislocation arrangement.
1.16. Depth-Resolved Measurements
Depth-resolved measurements are performed using a special differential
aperture X-ray microscopy (DAXM) method. The idea of the method
is as follows: a 50 m platinum wire serves as a differential aperture
and moves parallel to the sample surface in the diffracted radiation
and shadows different regions of the Laue pattern. Using a ray-tracing
algorithm it is possible to further relate the depth and diffracted intensity.
Differential aperture X-ray microscopy measurements using polychromatic
radiation give information about depth-resolved lattice rotations and GND
density tensor based on point-to-point orientation correlations. When
monochromatic radiation is used, the dilatational strain gradients can be
recovered. The instrumental description of DAXM is given in detail in
Chapter 2, and the code for depth-resolved measurements reconstruction is
discussed in Chapter 10.
1.16.1. Imaging the depth-resolved elastic strain tensor with
DAXM
Although the directions of four non-planar Bragg reections determine
the average deviatoric strain tensor of a crystalline grain, determination
of the absolute strain tensor requires that at least one wavelength be
determined (Chung and Ice, 1999). This can be done by measuring the
energy of one reection with a solid-state detector (Rebonato et al., 1989)
with a wavelength dispersive diffracted beam analyzer (Ice, 1987) or with
an incident beam monochromator (Ice, 1987; Ice et al., 2000a, 2000b).
Although all three methods have been used, an incident beam monochromator provides better energy resolution than existing energy dispersive
detectors and provides better mechanical stability than a diffracted beam
analyzer (Sheremetyev et al., 1991).
The availability of a non-dispersive monochromator similarly allows
for detailed mapping of reciprocal space within the micro-beam probe
volume. Assuming an average sample orientation, the intensity distribution
in reciprocal space volume is directly mapped out by plotting the measured

Diffraction Analysis of Defects: State of the Art

45

detector intensity in each radial direction with the momentum transfer


determined by the scattering angle and X-ray wavelength.
1.16.2. Resolving plastic and elastic strain effects
In general, both elastic and plastic strains are present in samples. Sometimes
the strain eld is mainly plastic (as in a nanoindented Cu crystal (Barabash
et al., 2001)). Sometimes the strain eld is mostly elastic as in a silicon
crystal bent at room temperature (Larson et al., 2008) analyzed in connection with the problem of sagittal focusing of high-energy synchrotron
X-rays and of neutron imaging. As described above, polychromatic X-ray
microdiffraction directly measures the average elastic strain within a
resolved volume through the deformed unit cell parameters (Chung and
Ice, 1999). The plastic deformation tensor can also be determined, but the
connection to experiment is less direct. Typically, streaking in Laue patterns
is attributed to the presence of lattice rotations as a result of unpaired
or geometrically necessary dislocations (Barabash et al., 1976, 2001,
2003a; Ice and Barabash, 2007). However, streaking can also result from
gradients in the elastic strain tensor (Yang et al., 2003; Larson et al., 2007,
2008). In cases where both elastic and plastic deformations are present,
the average local elastic contribution must be calculated point-by-point
and used to correct the lattice rotations that are then used to calculate
the deformation tensor. The interplay between elastic and plastic strains is
particularly relevant to mesoscale deformation structure and strain-gradient
plasticity theories.
Assuming linear superposition of plastic and elastic strain elds, we
may write the total mean deformation tensor as:
p

ij = ij + ije

(1.48)

Elastic strain gradient results in both macroscopic lattice rotation of the


crystal (deviatoric strain) and a linear gradient of lattice unit cell parameters
(isostatic strain component). The elastic part can be obtained from the lattice
parameters of the unit cell in different regions of the crystal and subtracted.
With micro-Laue diffraction measurements, it is imperative to retain all
nine components of the crystal strain tensor and local grain orientation.
The stress tensor can be further determined from the strain tensor and the

46

R. I. Barabash & G. E. Ice

anisotropic single-crystal elastic constants (stiffness moduli) (Noyan and


Cohen, 1987; Chung and Ice, 1999):

Cijkl kl
(1.49)
ij =
kl

Further details of elastic strain measurements with micro-Laue diffraction


can be found in Chapter 2.
1.16.3. Depth-resolved strain gradient in NiAl/Mo composite
Here we describe an example application of depth-resolved strain measurements to recover possible near-surface strain gradients in individual submicron phases of a NiAl/Mo composite. A direct d-spacing measurement
with micro-Laue diffraction is performed using monochromatic radiation
and scanning energy within a certain energy range. Initial examination of the
composite structure is performed using polychromatic radiation. NiAl/Mo
in situ composites grow with cube-on-cube orientation along the [001]
direction. As the Laue pattern is sensitive only to orientation, the (00h)
type intensity from both the Mo bers and the NiAl matrix is diffracted
in the same direction of reciprocal space. Due to unavoidable small 0.3
misorientations between the matrix and different bers, their Laue spots
are distinct in the Laue pattern (Fig. 1.15).

Fig. 1.15. Indexed Laue pattern of the NiAl/Mo composite with growth direction along
[001] direction (left) and enlarged region with [00h]-type reections for individual Mo bers
and NiAl matrix (right).

Diffraction Analysis of Defects: State of the Art

47

Fig. 1.16. Depth-resolved near-surface strain gradients in (a) individual Mo bers and (b)
NiAl matrix.

The enlarged central (00h) Laue spot shows distinct inputs from
NiAl and different Mo bers. Depth-resolved measurements of the (006)
reection were performed with DAXM and monochromatic radiation
within the energy range 18.42018.700 keV for NiAl and 17.0917.47 keV
for Mo bers with 2eV step (Fig. 1.16) in the as-grown state. Both phases
demonstrate strong near-surface gradient of the inverse lattice parameter,
Q = 2/d. Mo bers are in compression along the growth direction and
NiAl matrix is under tension. The as grown strain in these composites is
caused by the difference between the thermal expansion coefcients and
lattice mismatch. Near the surface these as-grown strains tend to relax to a
strain-free surface layer. The detailed micromechanical analysis of the strain
gradient in composites can be found elsewhere (for example, see Bei et al.
(2008) and Barabash et al. (2010a, 2010b, 2011a, 2011b)). Interfaces put
additional constraint on the process of strain relaxation. A non-monotonic
strain gradient in NiAl matrix (Fig. 1.16b) is caused by characteristic strain
uctuations at the interfaces between the matrix and the bers (see Barabash
et al. (2011a) for details). Other examples of depth-resolved measurements
are described in Chapter 2 and Chapter 10.
1.17. Conclusions
Hierarchical mesoscale dislocation arrangements displace atoms in the
crystals and change diffraction conditions. Statistically stored dislocations
create locally uctuating overlapping strain elds, and result in the
broadening of reections along the radial direction of the momentum
transfer. Full width half maximum of the broadened distribution scales

48

R. I. Barabash & G. E. Ice

as, n. GN-dislocations additionally cause macroscopic lattice rotations


(bending, torsion etc.) in the representative volume element and form a
broad at-top intensity distribution in the orientation space which can be
probed with micro-Laue diffraction, reciprocal space mapping and rocking
curve techniques. GNDs-induced broadening of orientation space is scaledependent, while broadening from statistically stored dislocations is not.
Full width at half maximum of the GND-induced broadened intensity
streaks in orientation space linearly depends on the density of the GND,
contrast factor and size of the probed region. Streak analysis may be used to
determine the GNDs density tensor components. Dislocation boundaries
split the intensity distribution in the orientation space, Laue streaks
and rocking curves. Large dislocation loops diffract X-rays similar to
straight dislocations, while small dislocation loops and dislocation dipoles
cause diffuse scattering. Three-dimensional depth-resolved measurements
illustrate the possibility of determining elastic strain gradients and pointto-point orientation correlations.

Acknowledgments
This research was supported by the U.S. Department of Energy, Ofce
of Basic Energy Sciences, Materials Sciences and Engineering Division.
The use of the APS was supported by U.S. Department of Energy, Ofce
of Basic Energy Sciences, and the Scientic Users Facilities Division.
Measurements were performed at 34 ID-E of the Advanced Photon Source.
The authors wish to thank Dr. J. Vitek (ORNL) for carefully reading the
manuscript.

References
Arsenlis, A. and Parks, D.M. (1999). Acta Mater., 47, 15971611.
Arsenlis, A., Parks, D.M., Becker, R. and Bulatov, V.V. (2003). J. Mech. Phys.
Solids, 52, 1213.
Balzar, D., Audebrand, N., Daymond, M.R., Fitch, A., Hewat, A., Langford, J.I.,
Le Bail, A., Louer, D., Masson, O., McCowan, C.N., Popa, N.C., Stephens,
P.W. and Toby, B.H. (2004). J. Appl. Crystallogr., 37, 911924.
Barabash, R.I. (2001). Mater. Sci. Eng., A309A310, 4954.

Diffraction Analysis of Defects: State of the Art

49

Barabash, R.I., Bei, H., Gao, Y.F. and Ice, G.E. (2011a). Scripta Mater., 64,
900903.
Barabash, R.I., Bei, H., Gao, Y.F. and Ice, G.E. (2010a). Acta Mater., 58,
67846789.
Barabash, R.I., Bei, H., Gao, Y.F., Ice, G.E. and George, E.P. (2010b). JMR, 25,
199206.
Barabash, R.I., Bei, H., Ice, G.E., Gao,Y.F. and Barabash, O.M. (2011b). JOM-US,
63, 3, 3034.
Barabash, R.I., Demin, S., Krivoglaz, M. and Chuistov, K. (1990). J. Appl.
Crystallogr., 23, 458561.
Barabash, R.I., Ice, G.E., Karapetrova, E.A. and Zschack, P. (2012). Metall. Mater.
Trans. A, 43A, 14131422.
Barabash, R.I., Ice, G.E., Kumar, M., Ilavsky, J. and Belak, J. (2009). Int. J.
Plasticity, 25, 20812093.
Barabash, R.I., Ice, G.E., Larson, B.C., Pharr, G.M., Chung, K.-S. and Yang, W.
(2001). Appl. Phys. Lett., 79, 749.
Barabash, R.I., Ice, G.E., Tamura, N., Valek, B.C., Bravman, J.C., Spolenak, R.
and Patel, J.R. (2003b). J. Appl. Phys., 93, 9, 57015706.
Barabash, R.I., Ice, G.E., Tamura, N., Valek, B.C., Bravman, J.C., Spolenak, R.
and Patel, J.R. (2004). Microelectron. Eng., 75, 1, 2430.
Barabash, R.I., Ice, G.E. and Walker, F. (2003a). J. Appl. Phys., 93, 3, 14571464.
Barabash, R.I. and Krivoglaz, M.A. (1978). Fiz. Met. Metalloved., 45, 718.
Barabash, R.I. and Krivoglaz, M.A. (1982). Metalloz. Nov. Tech., 4, 310.
Barabash, R.I., Krivoglaz, M.A. and Ryaboshapka, K.P. (1976). Fiz. Met.
Metalloved., 41, 3343.
Bei, H., Barabash, R.I., Ice, G.E., Liu, W., Tischler, J. and George, E.P. (2008a).
Appl. Phys. Lett., 93, 071904.
Bei, H. and George, E.P. (2005). Acta Mater., 53, 6977.
Bei, H., Shim, S., George, E.P. and Pharr, G.M. (2008b). Acta Mater., 56,
47624770.
Breuer, D., Klimanek, P. and Pantleon, W. (2000). J. Appl. Crystallogr., 33,
12841204.
Busing, W.R. and Levy, H.A. (1966). Acta Crystallogr. A, 22, 457464.
Chakravarthy, S.S. and Curtin, W.A. (2011). PNAS, 38, 1571615720.
Chung, J.-S. and Ice, G.E. (1999). J. Appl. Phys., 86, 5249.
Gao, H., Huang, Y., Nix, W.D. and Hutchinson, J.W. (1999). J. Mech. Phys. Solids,
47, 12391263.
Greer, J.R., Oliver, W.C. and Nix, W.D. (2005). Acta Mater., 53, 1821.
Greer, J.R., Oliver, W.C. and Nix, W.D. (2006). Acta Mater., 54, 1705.

50

R. I. Barabash & G. E. Ice

Groma, I. (1998). Phys. Rev. B, 57, 75357542.


Groma, I. and Szekely, F. (2006). Acta Mater., 54, 753757.
Groma, I., Ungar, T. and Wilkens, M. (1988). J. Appl. Crystallogr., 21, 4753.
Hansen, N. (2001). Metall. Mater. Trans. A, 32A, 29172935.
Hirsch, P.B. (1956). Progress in Metal Physics, Oxford: Pergamon Press, p. 236.
Huang, E.W., Barabash, R.I., Ice, G.E., Clausen, B., Liu, Y.L., Kai, J.J., Woods,
K.P. and Liaw, P.K. (2010). Int. J. Plasicity, 26, 11241137.
Huang, E.W., Barabash, R.I., Jia, N., Wang, Y.D., Ice, G.E., Clausen, B., Horton, J.
and Liaw, P.K. (2008). Metall. Mater. Trans. A, 39A, 30793088.
Huang, Y., Gao, H., Nix, W. and Hutchinson, J. (2000). J. Mech. Phys. Solids, 48,
1, 99128.
Hughes, D.A. and Hansen, N. (2000). Acta Mater., 48, 29853004.
Hughes, D.A., Liu, Q., Chrzan, D.C. and Hansen, N. (1997). Acta Mater., 45,
105112.
Hutchinson, J.W. and Evans, A.G. (2000). Acta Mater., 48, 125135.
Ice, G.E. (1987). Nucl. Instrum. Meth. B, 245, 397.
Ice, G.E. and Barabash, R.I. (2007). White Beam MicroDiffraction and Dislocations Gradients, in Nabarro, F.R.N. and Hirth, J.P. (eds), Dislocations in
Solids, Elsevier, Vol. 13, Chapter 79, pp. 500601.
Ice, G.E., Chung, J.-S., Lowe, W., Williams, E. and Edelman, J. (2000a). Rev. Sci.
Instrum., 71, 2001.
Ice, G.E., Chung, J.-S., Tischler, J.Z., Lunt, A. and Assoud, L. (2000b). Rev. Sci.
Instrum., 71, 2635.
Ice, G.E. and Larson, B.C. (2002). Adv. Eng. Mater., 2, 10, 643646.
Jakobsen, B., Poulsen, H.F., Lienert, U., Almer, J., Shastri, S.D., Sorensen, H. O.,
Gundlach, C. and Pantleon, W. (2006). Science, 312, 889.
Kaganer, V.M. and Sabelfeld, K.K. (2010). Acta Crystallogr. A, 66, 703716.
Kaganer, V.M. and Sabelfeld, K.K. (2011). Phys. Status Solidi A, 208, 25632566.
Klimanek, P. (1993). J. Phys. IV, 3, C7, 21492154.
Kocks, U.F., Hasegawa, T. and Scattergood, R.O. (1980). Scripta Mater., 14, 449.
Krivoglaz, M.A. (1969). Theory of Scattering of X-Rays and Thermal Neutron
Scattering by Real Crystal, New York: Plenum Press.
Krivoglaz, M.A. (1996). Theory of X-Ray and Thermal Neutron Scattering by Real
Crystals, New York: Springer-Verlag.
Krivoglaz, M.A., Ryaboshapka, K.P. and Barabash, R.I. (1970). Fiz. Met.
Metalloved+, 30, 11341145.
Larson, B.C., El-Azab, A., Yang, W., Tischler, J.Z., Liu, W. and Ice, G.E. (2007).
Phil. Mag. Lett., 87, 13271347.
Larson, B.C. and Schmatz, W. (1980). Phys. Status Solidi B, 99, 267275.

Diffraction Analysis of Defects: State of the Art

51

Larson, B.C., Tischler, J.Z., El-Azab, A. and Liu, W. (2008). J. Eng. Technol., 130,
021024-110.
Larson, B.C., Yang, W., Ice, G.E., Budai, J.D. and Tischler, J.Z. (2002). Nature,
415, 887890.
Larson, B.C. and Young, F.W. (1987). Phys. Status Solidi A, 104, 273286.
Lauridsen, E.M., Jensen, D.J., Poulsen, H.F. and Lienert, U. (2000). Scripta Mater.,
43, 561.
Levine, L.E., Larson, B.C., Yang, W., Kassner, M.E., Tischler, J.Z., Delos-Reyes,
M.A., Fields, R.J. and Liu, W. (2006). Nat. Mater., 5, 619.
Maas, R., Van Petegem, S., Van Swygenhoven, H. and Derlet, P.M. (2007). Phys.
Rev. Lett., 99, 145505.
Margulies, L., Winther, G. and Poulsen, H.F. (2001). Science, 291, 23922394.
May, C., Klimanek, P. and Magerl, A. (1995). Nucl. Instrum. Meth. A, 357,
511518.
Mughrabi, H. (1983). Acta Metall. Mater., 31, 9, 13671379.
Mughrabi, H. and Obst, B. (2005). Z. Metallkd., 96, 7, 586695.
Mughrabi, H. and Ungar, T. (2002). Long-range internal stresses in deformed
single-phase materials: The composite model and its consequesces, in
Nabarro, F. R. N. and Hirth, J.P. (eds), Dislocations in Solids, Elsevier,
Chapter 60, p. 345.
Nabarro, F.R. N. (1987). Theory of Crystal Dislocations, New York: Dover
Publications.
Nabarro, F.R. N. (1994). Scripta Metall. Mater., 30, 10851987.
Nabarro, F.R. N. (2001). Mater. Sci. Eng. AStruct., 317, 12.
Needleman, A. (2000). Acta Mater., 48, 105124.
Nix, W.D. and Gao, H. (1998). J. Mech. Phys. Solids, 46, 411425.
Noyan, I.C. and Cohen, J.B. (1987). Residual Stress, New York: Springer-Verlag.
Nye, J.F. (1953). Acta Metall. Mater., 1, 153162.
Ohashi, T., Barabash, R.I., Pang, J.W. L., Ice, G.E. and Barabash, O.M. (2009).
Int. J. Plasticity, 25, 920941.
Pantleon, W. (2002). J. Mater. Res., 17, 9, 24332441.
Patterson, A.L. (1959). International Tables for X-ray Crystallography, Vol. II,
Birmingham, UK: Knoch, p. 61.
Rebonato, R., Ice, G.E., Habenschuss, A. and Bilello, J.C. (1989). Phil. Mag. Lett.,
A60, 571.
Rollett, J.S. (1965). Computational Methods in Crystallography, Oxford, UK:
Pergamon Press.
Romanov, A.E. and Vladimirov, V.I. (1992). In Nabarro, F.R. N. and Hirth, J.P.
(eds), Dislocations in Solids, Elsevier, Vol. 9, pp. 191402.

52

R. I. Barabash & G. E. Ice

Ryaboshapka, K. (1993). Physics of X-ray Scattering by Deformed Crystals, Kiev:


Naukova Dumka.
Ryaboshapka, K., Stelmashenko, N. and Grigorev, O. (1990). Mater. Sci. Eng.,
A127, 6569.
Schaer, E., Simon, K., Bernstorf, S., Hanak, P., Tichy, G., Ungar, T. and
Zehetbauer, M.J. (2005). Acta Mater., 53, 2, 315322.
Sheremetyev, I.A., Turbal, A.V., Litvinov,Y.M. and Mikhailov, M.A. (1991). Nucl.
Instrum. Meth. A, 308, 451.
Stephen, P.W. (1999). J. Appl. Crystallogr., 32, 281289.
Stolken, J.S. and Evans, A.G. (1998). Acta Mater., 46, 51095115.
Szekely, F., Groma, I. and Lendvai, J. (2001). Scripta Mater., 45, 1, 5560.
Teodosiu, C. (1982). Elastic Models of Crystal Defects, Berlin: Springer-Verlag.
Thilly, L., Van Petegem, S., Renault, P.-O., Lecouturier, F., Vidal, V., Schmidt, B.
and Van Swygenhoven, H. (2009). Acta Mater., 57, 31573169.
Tomota, Y., Tokuda, H., Adachi, Y., Wakita, M., Minakawa, N., Moriai, A. and
Morii, Y. (2004). Acta Mater., 52, 20, 57375745.
Trinkaus, H. (1972). Phys. Status Solidi B, 51, 307319.
Uchic, M.D., Dimiduk, D.M., Florando, J.N. and Nix, W.D. (2004). Science, 305,
986.
Ungar, T., Dragomir, I., Revesz, A. and Borbely, A. (1999). J. Appl. Crystallogr.,
32, 9921002.
Ungar, T., Groma, I. and Wilkens, M. (1989). J. Appl. Crystallogr., 22, 2634.
Ungar, T., Mughrabi, H., Ronnpagel, D. and Wilkens, M. (1984). Acta Metall.
Mater., 32, 333342.
Volkert, C.A. and Lilleodden, E.T. (2006). Philos. Mag., 86, 3335, 55675579.
Wang, L., Barabash, R., Yang, Y., Bieler, T.R., Crimp, M.A., Eisenlohr, P., Liu, W.
J. and Ice, G.E. (2011). Metall. Mater. Trans. A, 42, 626635.
Warren, B.E. (1990). X-Ray Diffraction, Ontario: Dover Publications, p. 381.
Warren, B.E. and Averbach, B.L. (1950). J. Appl. Phys., 21, 6, 595599.
Warren, B.E. and Averbach, B.L. (1952). J. Appl. Phys., 23, 4, 497.
Wilkens, M. (1970). Phys. Status Solidi A, 2, 359370.
Wilkens, M., Ungar, T. and Mughrabi, H. (1987). Phys. Status Solidi A, 104,
157170.
Yang, W.G., Larson, B.C., Ice, G.E., Tischler, J.Z., Budai, J.D., Chung, K.-S. and
Lowe, W.P. (2003). Appl. Phys. Lett., 82, 3856.
Zhu, H.X. and Karihaloo, B.L. (2008). Int. J. Plasticity, 24, 9911007.

You might also like