Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Acta Materialia 132 (2017) 598e610

Contents lists available at ScienceDirect

Acta Materialia
journal homepage: www.elsevier.com/locate/actamat

Full length article

Direct measurement of critical resolved shear stress of prismatic and


basal slip in polycrystalline Ti using high energy X-ray diffraction
microscopy
L. Wang a, b, *, Z. Zheng a, H. Phukan b, P. Kenesei c, J.-S. Park c, J. Lind d, e, R.M. Suter d,
T.R. Bieler b
a
School of Materials Science and Engineering, Shanghai Jiao Tong University, China
b
Department of Chemical Engineering and Materials Science, Michigan State University, USA
c
Advanced Photon Source, Argonne National Laboratory, USA
d
Department of Physics, Carnegie Mellon University, USA
e
Lawrence Livermore National Laboratory, USA

a r t i c l e i n f o a b s t r a c t

Article history: Knowledge of the critical resolved shear stress (CRSS) values of different slip modes is important for
Received 10 October 2016 accurately modeling plastic deformation of hexagonal materials. Here, we demonstrate that CRSS can be
Received in revised form directly measured with an in-situ high energy X-ray diffraction microscopy (HEDM) experiment. A
13 February 2017
commercially pure Ti tensile specimen was deformed up to 2.6% strain. In-situ far-field HEDM experi-
Accepted 5 May 2017
Available online 7 May 2017
ments were carried out to track the evolution of crystallographic orientations, centers of masses, and
stress states of 1153 grains in a material volume of 1.1 mm  1 mm  1 mm. Predominant prismatic slip
was identified in 18 grains, where the orientation change occurred primarily by rotation around the c-
Keywords:
Critical resolved shear stress
axis during specimen deformation. By analyzing the resolved shear stress on individual slip systems, the
Titanium estimated CRSS for prismatic slip is 96 ± 18 MPa. Predominant basal slip was identified in 22 other grains,
Crystal plasticity where the orientation change occurred primarily by tilting the c-axis about an axis in the basal plane.
High-energy X-ray diffraction The estimated CRSS for basal slip is 127 ± 33 MPa. The ratio of CRSSbasal/CRSSprismatic is in the range of 1.7
In situ tension test e2.1. From indirect assessment, the CRSS for pyramidal 〈cþa〉 slip is likely greater than 240 MPa. Grain
size and free surface effects on the CRSS value in different grains are also examined.
© 2017 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

1. Introduction prismatic slip is the easiest mode to be activated [35]. It is difficult


to assess the CRSS of individual slip modes by deforming a large
Titanium and its alloys are widely used in aerospace and single crystalline material though, as multiple slip and twinning
biomedical industries. For manufacturers, material strength under activities can occur simultaneously during deformation and thus
complicated loading conditions is a key consideration for product complicate the analysis. As a result, CRSS of individual slip modes in
design. At the grain scale, material yielding corresponds to the Ti are more often determined by three alternative methods.
activation of slip systems in individual grains. Critical resolved The first method is to statistically count the observations of slip
shear stress (CRSS) for different slip modes is therefore important lines from different slip modes in a deformed polycrystalline
information for material design. As a hexagonal material, Ti has aggregate. Grain orientations are measured by electron back-
four major slip modes; namely, f0001g < 1210 > basal slip, scattered diffraction (EBSD) or transmission electron microscopy
f1010g < 1210 > prismatic slip, f1011g < 1210 > pyramidal <a> (TEM) in order to identify the activated slip systems by trace
slip, and f1011g < 2113 > pyramidal 〈cþa〉 slip [1,2]. Among them, analysis and to estimate their Schmid factors based upon a global
stress. Applying certain statistical assumptions, the relative CRSS
ratios between different slip modes can be identified. Table 1 shows
the obtained CRSS ratios in the literature, determined by slip line
* Corresponding author. School of Materials Science and Engineering, Shanghai trace analysis [68]. This method, however, usually does not pro-
Jiao Tong University, 800 Dongchuan Road, Shanghai, 200240, China.
vide the absolute CRSS values unless in-grain slip activity can be
E-mail address: leyunwang@sjtu.edu.cn (L. Wang).

http://dx.doi.org/10.1016/j.actamat.2017.05.015
1359-6454/© 2017 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
L. Wang et al. / Acta Materialia 132 (2017) 598e610 599

Table 1
A literature survey of CRSS values and ratios with respect to prismatic slip (in parentheses) of different slip and twin modes in pure Ti. The oxygen concentration and average
grain size from each reference are also shown if they are reported.

Prismatic Basal pyr <a> pyr 〈cþa〉 T1 twin O (wt%) Grain size (mm) Ref.

N/A N/A N/A N/A (<13) N/A 0.10 50 [6]


N/A N/A (3.6) N/A (25.4) N/A (22.5) N/A 0.25 115 [7]
120 MPa 182 MPa (1.5) 149 MPa (1.2) 240 MPa (2.0) N/A 0.16 40 [8]
37 MPa 49 MPa (1.3) N/A 197 MPa (5.3) 213 MPa (5.8) <0.01 N/A [10]
30 MPa 150 MPa (5.0) N/A 120 MPa (4.0) 125 MPa (4.2) <0.01 30 [11]
60 MPa 120 MPa (2.0) N/A 180 MPa (3.0) 125 MPa (2.1) 0.17 80 [12,13]
80 MPa 90 MPa (1.1) 110 MPa (1.4) 260 MPa (3.3) 220 MPa (2.8) 0.06 25 [14]
98 MPa N/A N/A 224 MPa (2.3) 136 MPa (1.4) <0.001 20 [15]
68 MPa 175 MPa (2.6) 120 MPa (1.8) 250 MPa (3.7) 230 MPa (3.4) 0.12 50 [16]
90 MPa 180 MPa (2.0) 140 MPa (1.6) 210 MPa (2.3) N/A 0.11 9 [17]
181 MPa 209 MPa (1.2) N/A 474 MPa (2.6) N/A 0.07 N/A [18],a
110 MPa N/A N/A N/A N/A 0.07 N/A [19],b
96 ± 18 MPa 127 ± 33 MPa (1.7e2.1) N/A 240 MPa (2.4) 225 MPa (2.3) 0.17 100 this work
a
Without considering sample size effect.
b
After considering sample size effect.

monitored from an in-situ test inside a scanning electron micro- condition are recorded using an area detector as the sample is
scope (SEM) [8]. As the stress state in each grain is not necessarily rotated with respect to the X-ray beam. These diffraction peaks are
equal to the macroscopic applied stress [9], this introduces vari- then indexed to obtain the crystallographic orientation, centers of
ability into the conditions in which slip occurs. mass (COMs), and average stress state of individual grains in the
The second method is based on crystal plasticity modeling. A volume illuminated by the X-ray [2329]. When this analysis
polycrystalline specimen with measured initial texture is deformed process is repeated for different deformation steps while deforming
by a mechanical test, and the deformation process is simulated a sample (i.e. in situ test), microstructure evolution in individual
using a crystal plasticity model. By comparing the stress-strain grains can be characterized. In situ far-field HEDM has been used to
curves, texture evolution, and deformation details within individ- investigate the micromechanical behavior of polycrystalline mate-
ual grains between experimental observation and model predic- rials such as twin nucleation events in individual grains in poly-
tion, an optimal set of CRSS values for different slip modes can be crystalline Mg, Ti, and Zr alloys [9,3034].
inferred. Table 1 shows the CRSS values of different slip modes in Ti In the present work, this method is used to analyze grains that
reported in the literature, determined by crystal plasticity modeling deformed by a predominant dislocation slip system. In a far-field
[1017]. Most of the models include f1012g < 1011 > (T1 twin- HEDM experiment the activated slip system(s) in a grain can be
ning) as an additional deformation mode, whose CRSS values are analyzed from the shape of its diffraction peaks [3538]. Using a
also shown in Table 1. Significant variation of the CRSS values is forward modeling approach: the shape of the diffraction peaks in
evident in the table. Part of the variation can be attributed to the selected grains are simulated assuming activation of one or more
different oxygen concentration and grain size of the materials. slip systems, then the simulation result is compared with the
However, sometimes even for materials with similar oxygen con- measured diffraction peaks to identify the most likely slip system(s)
centration and grain size, the reported CRSS values can be quite that can account for the observation. Here, a different approach is
different (e.g. Refs. [11,15]). This difference reflects intrinsically used to identify the active slip systems in a larger number of grains
different constitutive frameworks used in those models, including based on the continuous rotation of the grain orientation during
the number of slip and twin modes that are allowed to be activated. deformation. In particular, prismatic slip causes a grain to gradually
The third method uses micromechanical testing of small single rotate around its 〈0001〉 axis, while basal slip causes a grain to
crystal specimens prepared by focused ion beams (FIB). The gradually tilt its 〈0001〉 axis about an axis residing in the basal
advantage of this method is that a crystal orientation favorable for a plane. With knowledge of the average stress tensor in each grain
specific slip mode can be pre-selected and machined for testing, (which can be significantly different from the global stress tensor
which makes the measurement more controllable. Table 1 shows [9,34]), the resolved shear stress (RSS) on the corresponding acti-
the measured CRSS of a commercially pure Ti material from vated slip system can be directly calculated and identified as the
microcantilever bending experiments performed by Gong and CRSS value. This analysis is performed for many grains that show
Wilkinson [18,19]. This method, however, requires an advanced predominant prismatic slip or predominant basal slip to obtain
SEM and experienced operators to conduct sample preparation and average CRSS values for the two slip modes. By comparing the RSS
subsequent micromechanical testing. Effects due to surface damage of different slip modes when one slip mode is activated and the
and the small sample volume also introduce some variability in the others are not, we can also estimate the CRSS ratio between basal
data [20]. slip and prismatic slip. The influence of the grain size and the free
In this present study, a fourth method to measure CRSS of surface on the CRSS value in different grains is also examined. This
different slip modes in hexagonal materials is presented. It is based analysis neglects the expected inhomogeneous stress distributions
on in-situ mechanical testing of a polycrystalline specimen and the within grains, which contributes to the variations in apparent CRSS
use of far-field high energy X-ray diffraction microscopy (HEDM) to values among different grains.
track the deformation history of individual grains. Far-field HEDM,
also known as three-dimensional X-ray diffraction (3DXRD), was 2. Experimental
first developed at Risø National Laboratory in Denmark by Poulsen
et al. and implemented at the European Synchrotron Radiation The in situ far-field HEDM experiment was conducted at the
Facility (ESRF) [2124]. A high energy monochromatic X-ray beam Advanced Photon Source (APS) beamline 1-ID-E as illustrated in
is used to illuminate the entire cross section of a polycrystalline Fig. 1. A tensile sample of Grade 1 commercially pure Ti, with a cross
sample. Diffraction peaks from all grains that satisfy the diffraction sectional area of 1  1 mm2 and gauge length of 5 mm, was
600 L. Wang et al. / Acta Materialia 132 (2017) 598e610

implemented using the FABLE software package jointly developed


by ESRF and Risø National Laboratory [23,40]. For each far-field
HEDM measurement, diffraction patterns at each u step were
processed by the PeakSearch program in FABLE to identify all peaks
above a certain intensity threshold. The Transformation program
transformed the peak positions from the Cartesian coordinate on
the detector into a 2q-h diagram. The 2q-h diagram at the unde-
formed state was used to determine the instrument parameters
including sample-to-detector distance, detector tilt, and direct
beam position on the detector. With an optimal set of instrument
parameters, diffraction peaks can be converted into reciprocal lat-
tice vectors (i.e. g-vectors) in the 3D space. Finally, the GrainSpotter
program was used to index grains and determine their crystallo-
graphic orientations based on these reciprocal lattice vectors. The
minimum threshold to recognize a grain is that at least 40 peaks
were assigned to that grain when the macroscopic strain is less
than 0.3%, and at least 30 peaks were assigned to that grain when
the macroscopic strain is above 0.3%. In most of the recognized
grains, the number of identified peaks was often far greater than
the threshold value. For example, the average number of identified
peaks for the recognized grains in layer 6 is 56 at load step 1 (ε ¼ 0),
54 at load step 9 (ε ¼ 0.31%), and 50 at load step 29 (ε ¼ 2.6%), as
shown in Table 3. An independent Matlab code [31] was addition-
ally used to obtain COMs and the elastic strain tensor of each
indexed grain using the position of all of the diffraction peaks
through least squares fitting. The average uncertainty (i.e. error bar)
for COM position (DX, DY) and elastic strain components (Deij) can
be calculated from the covariance matrix of the estimates for the
Fig. 1. Far-field HEDM setup at APS 1-ID along with the coordinate system used for least squares fit, and these values are shown in Table 3. (DX, DY) is
data analysis. The sample was deformed in tension with 30 incremental steps, followed
in the range of 2030 mm, which is less than half of the average
by 7 steps of unloading, as shown in the stress-strain curve.
grain size. Deij is around 3  104. As a rough estimation using the
bulk modulus (110 GPa) and the shear modulus (44 GPa) of Ti, this
mounted in a specially designed load frame [39]. The chemical translates to uncertainty of about 33 MPa for normal stress com-
composition of the material is shown in Table 2. The material ponents and 12 MPa for shear stress components. At load step 1, the
contained nearly equiaxed grains with average grain size of average magnitude of (DeXX, DeZZ, DeZX) is comparable to the
~100 mm, as determined from previous EBSD analysis [12]. A strong average elastic strain components (eXX, eZZ, eZX). After the tensile
{0001} texture close to the tensile axis was present in the sample. A test began, the magnitude of (eXX, eZZ, eZX) rapidly increased,
100 mm tall by 1.5 mm wide monochromatic X-ray beam diminishing the impact of (DeXX, DeZZ, DeXY). At load step 9
(Energy ¼ 65.4 keV, l ¼ 0.0189 nm) illuminated the entire cross (ε ¼ 0.31%), absolute values of eXX (13  104) and eZZ (22  104)
section of the sample as it was rotated with respect to the X-ray already became 4 and 7 times of DeXX and DeZZ, respectively.
beam. Following the principles of far-field HEDM [23], the sample Because this is the stage when slip activity began, the fitting un-
was rotated about the tensile axis (Z) over a [-70 , 70 ] range and certainty for the elastic strain components is estimated to intro-
diffraction patterns were recorded on an amorphous Si area de- duce an uncertainty no greater than ±20% on the measured CRSS
tector placed approximately 1.0 m away from the sample as u values. The elastic strain tensors were transformed to stress tensors
swept through 1 intervals. The sample was deformed incremen- using Hooke’s Law with following elastic constants:
tally by tension with 30 steps, followed by 7 steps of unloading. The C11 ¼ 162.4 GPa, C33 ¼ 181.6 GPa, C44 ¼ 47.2 GPa, C12 ¼ 92 GPa,
stress-strain curve, derived from the load-displacement data C13 ¼ 69 GPa [31]. For a particular grain, this stress tensor, rather
assuming uniform macro-scale deformation, is shown in Fig. 1. At than the global uniaxial stress, was used to compute the resolved
each deformation step, a far-field HEDM measurement was per- shear stresses for different slip systems. The above procedures were
formed while the displacement of the load frame crosshead was performed for all 37 deformation steps. To study the evolution of
held constant during the u scan. To establish a large data set for resolved shear stress, a key point is to track the same grains among
statistical analysis, the measurement was performed in 11 adjacent different deformation steps. This is achieved by considering both
layers along the sample gauge. A digital image correlation (DIC) the crystallographic orientation and COM information: a grain in
camera was used to measure macroscopic strain and to ensure that steps N and Nþ1 are recognized as the same grain if the difference
approximately the same volume was illuminated at all 37 defor- of grain orientation is less than 1 and the distance between their
mation steps, which is critical for tracking individual grains through COM is less than 100 mm. Due to the small strain step size (less than
the test. 0.1% per step) and relatively large grain size, grain tracking was
Grain indexing from the collected diffraction patterns was straightforward over all deformation steps. Among the 11 layers
that correspond to a material volume of 1.1 mm  1 mm  1 mm, a
total of 1153 grains were tracked and analyzed for their orientation,
Table 2 rotation, and stress state evolution during sample loading. Grains
Chemical composition of the material under investigation (wt%). that show continuous and smooth changes in orientation are cho-
Element O Fe Al Cu C Ni S Cr N Ti
sen for estimating the CRSS of slip modes.
After the mechanical test, the sample was removed from the load
Fraction 0.17 0.049 0.017 0.017 0.015 0.013 0.011 0.011 0.004 bal
frame, and a near-field HEDM scan was performed near the central
L. Wang et al. / Acta Materialia 132 (2017) 598e610 601

Table 3
Average error bars for grain orientation (Dq), COM position (DX, DY), and elastic strain components (DeXX, DeZZ, DeXY) at three load steps. For comparison, average values of
elastic strain components (eXX, eZZ, eXY) at these load steps are also shown.

Load step Peaks per grain Average grain error bars Average elastic strain

Dq ( ) DX (mm) DY (mm) DeXX (104) DeZZ (104) DeZX (104) eXX (104) eZZ (104) eZX (104)

1 (ε ¼ 0) 56 0.30 22 31 3.3 3.2 2.3 3.8 1.2 0.2


9 (ε ¼ 0.31%) 54 0.31 23 32 3.4 3.2 2.3 13 22 2.5
29 (ε ¼ 2.6%) 50 0.44 28 37 3.9 3.7 2.7 14 31 3.0

part of the far-field measurement volume. Near-field HEDM is able 3.1. Grains with predominant prismatic slip
to yield spatially resolved 3D grain maps with much higher reso-
lution (~2 mm) than that provided by the COM [4144]. The X-ray During the tensile test, grain orientations gradually changed as a
beam for the near-field scan was 4 mm tall by 1.5 mm wide with an result of deformation. For the same grain, the difference in crys-
energy of 64.3 keV. The near-field X-ray area detector tallographic orientation between step 1 and each subsequent step,
(3 mm  3 mm with 2048  2048 pixels) was placed at distances of after excluding the effect of symmetry, is calculated as the disori-
L ¼ 5.2 and 7.2 mm from the rotation axis. The sample was rotated entation value at that step. The angular difference in the grain
over a 180 range, and one diffraction image was collected at each 〈0001〉 direction between step 1 and each subsequent step is
du ¼ 1 step for each L distance. By applying a forward modeling calculated as the c-axis misalignment at that step. Mathematically,
strategy [41,43], the crystal orientation of each voxel in the scanned the c-axis misalignment of a grain should not exceed its disorien-
volume can be determined, which eventually provides a spatially tation at any step.
resolved grain map in the measured slice. The near-field scan was Fig. 3 shows the development of disorientation and c-axis
performed for 100 consecutive slices along the sample gauge length, misalignment with respect to the initial grain orientation (before
which covered a material volume of 400 mm  1 mm  1 mm. Data deformation) for Grains 90, 95, 99, and 101. The disorientation and
reconstruction was performed for the top slice that resides in the 6th c-axis misalignment increased by almost the same amount in each
layer of the far-field scan. grain during early elastic deformation up to ~0.3% macroscopic
strain; this is likely the result of a slight rotation of the whole
sample (<0.5 ) prior to yielding. After that, the disorientation
continued to increase (up to 2.5 ) with strain, but concurrently the
3. Results c-axis misalignment from its initial orientation shows relatively
small changes. This indicates that the majority of the grain rotation
Fig. 2 shows the {0001} pole figure and the COM map for the 104 was around the c-axis during the plastic stage, which is evidence of
indexed grains in layer 6 (the middle layer) before deformation. The prismatic slip activity. For each grain, the orientation difference
{0001} texture is evident as most grains have their {0001} pole between the last loading step and a step near the yield point
close to the Z direction (tensile axis). The grain COM map is (marked by the two arrows) is also annotated in each figure as qrot,
consistent with the sample cross section dimension expressed in a rotational angle-axis convention [45]. The rotational
(1 mm  1 mm) and the estimated grain size (~100 mm). A previous axis is defined with respect to the crystal orthogonal coordinate
examination of the nucleation and evolution of a T1 twin nucle- system (x1, x2, x3) illustrated in the large hexagonal unit cell in
ation event in Grain 1 at 1.6% strain is reported in Ref. [32]. For most Fig. 3. For all 4 grains, the rotational axis has a very large component
other grains, dislocation slip was the dominant deformation along x3, which confirms the dominance of prismatic slip activity.
mechanism, which is the focus of this paper. Fig. 2 highlights 17 From the {0001} pole figure in Fig. 2(a), these 4 grains are all soft
grains that were tracked in most of the 30 loading steps. They are grains whose c-axis is almost perpendicular to the tensile direction
identified with color coded markers that indicate the predominant Z, making it easy for prismatic slip to be activated. Fig. 3 also plots
slip activity in each of them, based upon the analysis strategy the evolution of the von Mises stress calculated from the stress
described next. tensor:

Fig. 2. (a) {0001} pole figure of the 104 indexed grains in layer 6 before deformation. (b) Grain COM map on the cross section plane. The 17 analyzed grains are marked based on the
identified slip activity in each of them.
602 L. Wang et al. / Acta Materialia 132 (2017) 598e610

sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 
ðs11  s22 Þ2 þ ðs22  s33 Þ2 þ ðs33  s11 Þ2 þ 6 s212 þ s223 þ s231
sVM ¼
2

For each grain, when the disorientation and the c-axis 42) ¼ (4.2 , 70.3 , 42.3 )) when the macroscopic strain was 0.14%,
misalignment started to separate, the von Mises stress concurrently and the instantaneous tprism was 105 MPa tprism at this moment can
began to plateau. This shows a strong correlation between the be taken as the estimated CRSS for prismatic slip in Grain 90. The
activation of prismatic slip and grain yielding. instantaneous tbasal and t〈cþa〉 as well as the von Mises stress (sVM)
From the stress tensor, the resolved shear stress (RSS) on any are also indicated in Fig. 4. Note that tprism is lower than both tbasal
slip system can be calculated. Fig. 4 shows the RSS on the most and t〈cþa〉 when prismatic slip was activated, indicating that the
stressed prismatic slip system (tprism), basal slip system (tbasal), and prismatic slip mode has a lower CRSS than either basal slip and
pyramidal 〈cþa〉 slip system (t〈cþa〉) in Grains 90, 95, 99, and 101 as pyramidal 〈cþa〉 slip. The same calculations were performed for
a function of the macroscopic strain. Evolution of tprism is generally Grains 95, 99, and 101, in which the CRSS for prismatic slip are
smooth: it increased in early deformation prior to yield, then sta- found to be 90 MPa, 112 MPa, and 94 MPa, respectively.
bilized around a certain value for the rest of the loading steps. This After searching for similar grain rotation behavior in all eleven
is typical flow behavior for a grain deforming by a single dislocation layers, predominant prismatic slip was identified in a total of 18
slip, where no hardening takes place. For each grain, it is assumed grains (out of 1153 grains in the population). This number is small
that the strain increment when its disorientation deviated from the relative to the total number of grains for two reasons: (1) the
corresponding c-axis misalignment (see Fig. 3) marks the activation texture of the sample is generally unfavorable for prismatic slip; (2)
of the most stressed prismatic slip system (in all cases, that system only those grains whose orientation rotation apparently resulted
has the highest global Schmid factor assuming uniaxial tension from prismatic slip were counted, while more ambiguous cases
along Z). For example, prismatic slip ð0110Þ½2110 (Schmid were excluded. Fig. 5 shows the distribution of CRSS for prismatic
factor ¼ 0.44) was activated in Grain 90 (Bunge Euler angles (41, F, slip in these 18 grains. The average and standard deviation CRSS for

Fig. 3. Evolution of the disorientation, c-axis misalignment, and the von Mises stress in four grains that show prismatic slip activity. The initial orientation for each grain is
represented by a hexagonal cell in these images, with the tensile axis pointing out of the page. Orientation difference between the last loading step and a step at earlier deformation
(marked by the two arrows), expressed in a rotational angle-axis convention, are also shown (qrot). The rotational axis is defined in the crystal orthogonal coordinate system (x1, x2,
x3) illustrated in the hexagonal unit cell on the right.
L. Wang et al. / Acta Materialia 132 (2017) 598e610 603

Fig. 4. Evolution of tprism, tbasal, and t〈cþa〉 with strain in Grains 90, 95, 99, and 101. For each grain, instantaneous von Mises stress (sVM), tprism, tbasal, and t〈cþa〉 when prismatic slip
was activated are shown.

< 1010 > directions. For example, activation of ð0001Þ½1210 would


cause lattice rotation around ½1010, which corresponds to the di-
rection of [0.866, 0.5, 0] in the orthogonal crystal coordinate system
(x1, x2, x3). The observed rotation axis for Grain 2, Grain 29, and
Grain 69 are generally close to this type of theoretical direction.
From the observed rotation axis in Grain 2 (41, F, 42 ¼ 81.6 , 7.7,
39.4 ), it is inferred that ð0001Þ½1120 basal slip with a global
Schmid factor of 0.13 was activated. Similar calculations are per-
formed for Grains 29 and 69, where the identified active basal slip
systems had global Schmid factor of 0.35 and 0.19, respectively. For
Grain 5 (41, F, 42 ¼ 28.3 , 18.3 , 43.9 ), the rotation axis is close to
½2110, which could be achieved by the simultaneous operation of
two basal slip systems ð0001Þ½1210 and ð0001Þ½1120 (global
Schmid factors of 0.20 and 0.29, respectively).
Fig. 7 shows the RSS on the most stressed prismatic slip system
(tprism), basal slip system (tbasal), and pyramidal 〈cþa〉 slip system
(t〈cþa〉) in Grains 2, 5, 29, and 69 as a function of the macroscopic
Fig. 5. Distribution of the CRSS value for prismatic slip identified in 18 grains and basal
strain. For Grain 2, Grain 29, and Grain 69, the activated basal slip
slip identified in 22 grains.
systems inferred from lattice rotation are in fact the most stressed
basal slip system (i.e. highest RSS). For Grain 5, ð0001Þ½1120 is the
prismatic slip is 96 ± 18 MPa, with most instances falling in the most stressed basal slip system. In all four cases, tprism is lower than
70110 MPa range. the lowest CRSS value of 69 MPa identified above for prismatic slip
to be activated. Because the overall disorientation as well as the c-
axis misalignment changed continuously through the test, it is not
3.2. Grains with predominant basal slip
obvious when basal slip was first activated in as clear of a way as
that shown in Fig. 3 for prismatic slip (i.e. deviation of disorienta-
Fig. 6 shows the lattice rotation behavior and von Mises stress
tion from c-axis misalignment). As an alternative criterion, it is
evolution for Grains 2, 5, 29, and 69 in the same way as in Fig. 3. In
assumed that basal slip was activated when the von Mises stress
these grains, the increase of c-axis misalignment was almost equal
(sVM) began to plateau. Under this assumption, the CRSS values for
to the increase of disorientation throughout the loading. Hence, the
basal slip were calculated for each grain and shown in Fig. 7 along
grain rotation was mainly achieved by gradually tilting the c-axis,
with the instantaneous tbasal, t〈cþa〉, and sVM. The CRSS for basal slip
as illustrated by the large hexagonal unit cell in Fig. 6. This behavior
is apparently higher than the values determined for prismatic slip.
most likely results from basal slip, whose activation only tilts the c-
Grain 29 is a good example to show that use of the local stress
axis of the crystal without causing any rotation around the c-axis.
tensor to obtain the RSS is a more reliable parameter to predict slip
For each grain, the orientation change between the last loading step
activity than the global Schmid factor: when the basal slip was
and a step near the yield point (marked by the two arrows) are also
activated, tprism and tbasal were 61 MPa and 135 MPa, but the global
shown in the figure as qrot. Theoretically, activation of a single basal
Schmid factor for the two slip systems were equal (~0.35). From the
slip system should make the crystal rotate around one of the
604 L. Wang et al. / Acta Materialia 132 (2017) 598e610

Fig. 6. Evolution of the disorientation, c-axis misalignment, and the von Mises stress in four grains that show predominant basal slip activity. The increase in c-axis misalignment
was almost equal to the increase of disorientation throughout the loading. This suggests that the grain rotation was mainly achieved by gradually tilting the c-axis, as illustrated for
the specific example of activation of one basal slip resulting in rotation about < 101 0 > rotation axis. The disorientation between the last loading step and a step at early
deformation (marked by the two arrows) are also shown (qrot).

Fig. 7. Evolution of tprism, tbasal, and t〈cþa〉 with strain in Grains 2, 5, 29, and 69. For each grain, instantaneous von Mises stress (sVM), tprism, tbasal, and t〈cþa〉 when basal slip was
activated are shown.
L. Wang et al. / Acta Materialia 132 (2017) 598e610 605

global Schmid factor, one would expect prismatic slip to be acti- 3.3. Grains with weak or complicated slip activity
vated; but from the measured RSS, basal slip was clearly favored
over prismatic slip in this grain, consistent with the observation. While the 8 grains analyzed above appear to be dominated by
Though t〈cþa〉 reached 245 MPa, 243 MPa, and 242 MPa in Grains the activation of a single slip system, this is not always the case for
2, 5, and 69 by the time basal slip was activated, it appears that most other grains. Fig. 8 shows the development of disorientation
pyramidal 〈cþa〉 slip was not activated in a dominant way. Had and c-axis misalignment with strain for nine grains, in which it is
pyramidal 〈cþa〉> dominated, the absolute value of the x3 difficult to identify a dominant slip system with confidence. For
component of the rotation axis would be close to 0.23, a value at Grains 4, 36, and 47, the lattice rotation was relatively small, sug-
least two times greater than the observed absolute value of the x3 gesting weak slip activity. Such a change in the orientation may
component in all four grains. This suggests that the CRSS of pyra- have arisen from rigid body motion imposed by deformation in
midal 〈cþa〉 slip is likely no less than 240 MPa. On the other hand, neighboring grains. For Grains 12, 14, and 91, the c-axis misalign-
the fact that the x3 component is non-zero suggests that pyramidal ment and the disorientation change in an increase-decrease-
〈cþa〉 might have occurred as secondary slip. Grain 2 and Grain 69 increase manner. This complicated path of lattice rotation may
had relatively high ratio of t〈cþa〉/tbasal (~2.3 and ~2.1) whereas this have arisen from simultaneous operation of two slip systems at
ratio in Grain 5 and Grain 29 were ~1.6 and ~0.8. Activation of py- varying rates whose associated lattice rotation were partially
ramidal 〈cþa〉 slip in addition to basal slip might be that reason that canceled out from each other. For Grains 60, 61, and 65, the
the observed rotation axes show a relatively large x3 component in disorientation was higher than the c-axis misalignment during
Grain 2 and Grain 69 compared with Grain 5 and Grain 29. early deformation, but then the difference gradually shrank during
After searching for similar grain rotation behavior in all the later deformation. This scenario suggests a change in the dominant
eleven layers, predominant basal slip was identified in a total of 22 slip system with increasing strain in these grains. Fig. 9 shows the
grains. The distribution of CRSS for basal slip in these 22 grains is evolution of tprism, tbasal, and t〈cþa〉 in the same nine grains. Grains
shown in Fig. 5. The average CRSS for basal slip is 127 ± 33 MPa. 12, 14, 91, 60, 61 and 65 all had tbasal (RSS on the most stressed basal

Fig. 8. Evolution of the disorientation, c-axis misalignment, and the von Mises stress for Grains 4, 36, 47, 12, 14, 91, 60, 61, and 65. The initial orientation of each grain is represented
by a hexagonal cell in these images, with the tensile axis pointing out of the page.
606 L. Wang et al. / Acta Materialia 132 (2017) 598e610

Fig. 9. Evolution of tprism, tbasal, and t〈cþa〉 with strain in Grains 4, 36, 47, 12, 14, 91, 60, 61, and 65.

slip system) exceeding 140 MPa after grain yielding. Given the The two values were very similar throughout the loading. Upon
previously determined average CRSS for basal slip (127 MPa), it is unloading, residual stresses emerged again in many grains even
likely that basal slip was activated in these grains. The RSS on the after the macroscopic stress on the sample was fully removed. The
second most stressed basal slip system also exceeded 120 MPa in average von Mises stress was 160 MPa at the final unload step. This
each of the 6 grains at some point. The RSS on the most stressed post-deformation residual stress can be attributed to the “internal
prismatic slip system exceeded 70 MPa in Grains 12, 14, and 91. stresses” between neighboring grains that developed during
When multiple slip systems were activated in a grain at different deformation and unloading.
strains and in different regions, a complex path of grain rotation can
be expected. 3.5. Grain map from the near-field HEDM scan

3.4. Grain-level residual stress and stress heterogeneity After the in situ far-field HEDM, a near-field HEDM scan was
performed in approximately the same region of the sample. Near-
Fig. 10(aec) show the von Mises stress (sVM) development at field HEDM provides a non-destructive planar mapping of the
macroscopic strains of 0, 0.31%, and 2.6% in all grains in layer 6. detailed arrangement of grains in the microstructure similar to
Prior to the deformation, sVM varies from 50 to 200 MPa in different the more familiar electron back-scatter diffraction (EBSD) scans.
grains (Fig. 10(a)). This indicates the existence of residual stress in The original purpose for the near-field HEDM was to further
the initial state, which was probably due to the anisotropic thermal examine twin nucleation from the in situ far-field HEDM mea-
contraction of Ti during cooling after a recrystallization anneal. surement. (Among the 1153 grains, twin nucleation was identified
After the tensile test started, sVM in those grains increased and in 19 grains by far-field HEDM.) The near-field HEDM data allow
became more and more heterogeneous among different grains the exploration of the morphology and location of the nucleated
(Fig. 10(b and c)). This grain-level stress heterogeneity is the result twins with respect to other microstructually-relevant features
of the heterogeneous deformation paths among different grains. such as grain boundaries and the spatial arrangement of hard
Fig. 10(d) compares the average von Mises stress among all grains in versus soft grains, but this is beyond the scope of the present
layer 6 with the macroscopic tensile stress applied on the sample. paper.
L. Wang et al. / Acta Materialia 132 (2017) 598e610 607

Fig. 10. Von Mises stress (sVM) in different grains in layer 6 when the macroscopic strain was (a) at 0, (b) at 0.31%, and (c) at 2.6%. The values of sVM (in MPa) are shown next to the
COM of each grain, and the COM is colored according to this value. (d) Comparison of the average von Mises stress among all grains in layer 6 with the macroscopic tensile stress
applied on the sample. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 11. (a) Reconstructed grain map from the top slice of the near-field HEDM scan. Each grain is colored according to the angle between its c-axis and the Z direction: toward red if
that angle is close to 0; toward blue if that angle is close to 90 . (b) {0001} pole figure generated from the grain orientations. Grains 1, 2, 4, 5, 12, 14, 47, 61, 65, 69, 90, 91, 95, and 101
are marked in both the grain map and the pole figure. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 11(a) shows the grain map from the top slice of the near- is colored towards blue (soft-oriented grain). Fig. 11(b) shows the
field HEDM scan, which was found to be located in the 6th layer corresponding {0001} pole figure for the grains in the near-field
of the far-field scan. The pixel size for the reconstruction was set to map. Out of the 18 grains (including Grain 1 in which twin nucle-
be 2 mm, Each grain is colored according to the angle between its c- ation occurred) analyzed above, 14 are identified in the near-field
axis and the Z direction: if that angle is 0, the grain is colored to- grain map based on the match of both grain position and grain
wards red (hard-oriented grain); if that angle c-axis is 90 , the grain orientation. Comparing Figs. 11 and 2, the grain COMs are similar
608 L. Wang et al. / Acta Materialia 132 (2017) 598e610

between the near-field and far-field results, but there is no reason


to expect them to be the same, as the near field measurement
samples a thin slice rather than the entire volume of the grain.
Grain orientations are generally consistent, but some differences
still exist between the near-field and far-field grain orientations.
This difference is plausible and could result from sample mounting:
the near-field scan was performed after the sample was taken out of
the load frame and glued onto another stage; it is very difficult to
guarantee that the sample was remounted exactly the same way as
it was in the load cell. Grains 29, 60, 36 and 99 are absent in the
near-field scan grain map, which likely implies that these grains did
not intersect the measured slice but are present elsewhere in the
100 mm layer 6 of the far-field scan.

4. Discussion

4.1. CRSS ratio between basal slip and prismatic slip

From Sections 3.1 and 3.2, the CRSS values for prismatic slip and
basal slip are in the range of 96 ± 18 MPa and 127 ± 33 MPa,
respectively, in this material. Our previous work showed that the T1
twin in Grain 1 nucleated at a RSS of 225 MPa [32]. Although the
CRSS for pyramidal 〈cþa〉 slip is not directly determined from this
study, it is probably greater than 240 MPa from the above analysis.
Fig. 12. Instantaneous tprism and tbasal for the 18 grains that show prismatic slip (red
These results are listed in the last row in Table 1 for comparison
circle) and 22 grains that show basal slip (blue square). Grains with a free surface are
with the literature values. also distinguished from grains from the interior volume. The CRSS ratio between basal
The ratio of CRSSbasal/CRSSprism is particularly important for the slip and prismatic slip is estimated to be between 1.7 and 2.1. (For interpretation of the
ductility of hexagonal metals. A ratio that is not too far from 1.0 references to colour in this figure legend, the reader is referred to the web version of
allows both prismatic and basal slip to be activated during material this article.)

deformation, which is beneficial for the material ductility. Simply


dividing the average CRSS values (127 MPa for basal slip, 96 MPa for
the CRSS in some grains in which dislocation slip was initiated in a
prismatic slip), however, is not an accurate way to estimate this
region where the local stress was substantially higher than the
ratio of CRSSbasal/CRSSprism. Instead, the values of tprism and tbasal
grain-averaged stress. Unfortunately, it is difficult to examine the
at the onset of the dominant mode in each relevant grain are
stress localization with far-field data. The third factor is the influ-
used to estimate the upper and lower bounds for the ratio of
ence of neighboring grains. Many studies have shown that dislo-
CRSSbasal/CRSSprism. This method effectively eliminates the influ-
cation slip or twinning in one grain could stimulate the activation of
ence of grain size and neighboring grains. To some extent, it is
slip or twins in the neighboring grain through slip transfer, should
equivalent to conducting slip activity analysis for 40 single crystals
the two deformation systems have a good geometric alignment
with different orientations. The result is summarized in Fig. 12. Of
with each other [5,8,44,46,47]. Again, it is not straightforward to
the 18 grains that show prismatic slip, 17 have tbasal/tprism less
assess this issue with the far-field data that cannot provide accurate
than 2.1. In other words, prismatic slip is dominant only when
grain morphology information. The grain size and free surface are
tbasal/tprism is less than 2.1. For the 22 grains that show basal slip, all
two other factors that could account for the variation of the CRSS
have tbasal/tprism greater than 1.7. In other words, basal slip is
values. These two factors are examined in detail in Fig. 13. For each
dominant only when tbasal/tprism is greater than 1.7. Thus, we es-
grain, its grain size is estimated to be the distance from its COM to
timate the ratio of CRSSbasal/CRSSprism is between 1.7 and 2.1 for this
the COM of its 3rd closest neighboring grain. The average grain size
Ti material. This range is in the middle of the reported values in the
from this estimation is 103 mm, very close to the previously
literature (see Table 1).
measured grain size (~100 mm) of this material [12]. From the grain
From the grain COM positions, it is possible to distinguish grains
COM map, each grain is classified as either a surface grain if its COM
with a free surface (i.e. surface grains) from grains in the interior
is adjacent to one of the four surface edges or an inner grain if
volume (i.e. inner grains). From Fig. 12, this has no apparent in-
otherwise. Fig. 13(a) compares the CRSS and the grain size of the 18
fluence on the CRSS ratio, but free surface effects are discussed
grains that show prismatic slip. Nine of these are surface grains and
further in the next section.
the others are inner grains. It appears that the CRSS for prismatic
slip is negatively correlated with the grain size. The average CRSS
4.2. Reasons for the variation of the CRSS values
values for prismatic slip in the nine surface grains and the nine
inner grains are 91 ± 19 MPa and 102 ± 16 MPa, respectively. A
From Fig. 5, the CRSS for both prismatic slip and basal slip have a
statistical analysis using the t-test indicates that the probability for
broad range. The variation of CRSS from grain to grain may be
the alternative hypothesis H1 ¼ {CRSSsurface grains < CRSSinner grains}
attributed to several factors. First of all, the stress tensor in each
to be valid is 0.88. Fig. 13(b) compares the CRSS and the grain size of
grain was determined from its elastic strain tensor, which is fitted
the 22 grains that show basal slip. Again, it appears that the CRSS
using all identified peaks of that grain. As discussed in Section 2, the
for basal slip to occur in a grain is negatively correlated with the
fitting procedure itself is associated with some intrinsic uncer-
grain size. The average CRSS value for basal slip in the eight surface
tainty. Second, the stress tensor obtained from the far-field mea-
grains and the fourteen inner grains are 105 ± 25 MPa and
surement is a grain-averaged value. Local stress within a grain,
139 ± 31 MPa, respectively. The same t-test indicates that the
however, often deviates from the grain-averaged stress particularly
probability for the alternative hypothesis H1 ¼ {CRSSsurface
near grain boundaries. This could result in an underestimation of
L. Wang et al. / Acta Materialia 132 (2017) 598e610 609

this paper, analysis of a material volume of 1.1 mm  1 mm  1 mm


of the far-field HEDM measurement was made to identify enough
grains with dominant prismatic slip and basal slip to estimate the
CRSS for these two slip modes. A 3D grain structure reconstruction
using the information of all 11 layers and Voronoi tessellation for
subsequent crystal plasticity simulation has been constructed, and
will be the basis for future assessments of heterogeneous defor-
mation. The CRSS values obtained from the present work provide
important input parameters, and the accurately measured grain
orientation change during loading will serve as a good benchmark
to assess the quality of the model itself.

5. Conclusions

In situ far-field HEDM was used to track the deformation history


of individual grains in a bulk Ti polycrystalline sample deformed up
to 2.6% tensile strain. Deformation history in a total of 1153 grains
were analyzed with a focus on those grains that show predominant
activity of either prismatic slip or basal slip. The following con-
clusions are reached:

1. Predominant prismatic slip was found in 18 grains, whose lattice


rotations were mostly around the c-axis. Assuming that pris-
matic slip was activated when the disorientation started to
deviate from c-axis misalignment in each grain, the average
CRSS for prismatic slip was estimated to be 96 ± 18 MPa.
2. Predominant basal slip was found in 22 grains, whose lattice
rotations were predominantly a tilting of the c-axis about an
axis in the basal plane. Assuming that basal slip was activated
when the grain-averaged von Mises stress began to plateau, the
average CRSS for basal slip was estimated to be 127 ± 33 MPa.
The CRSS for pyramidal 〈cþa〉 slip systems, for which no ex-
amples were observed, is apparently greater than 240 MPa.
3. Not all grains show simple slip activity. Instead, the majority of
the grains show either small or complicated lattice rotations.
The latter can be attributed to the activation of multiple slip
systems within a grain.
4. An ex situ post-deformation near-field HEDM scan allows
visualization of the detailed arrangement of grains in the 6th
layer of the far-field scan. This provided confirmation of the
locations and credibility of geometric COM values extracted
from far-field measurements.
5. The ratio of CRSSbasal/CRSSprism is estimated to be between 1.7
and 2.1 for this commercial purity Ti material.
6. The observed variation of CRSS values for both prismatic slip and
basal slip can be attributed to a number of factors, such as the
intrinsic fitting uncertainty of the stress tensor, in-grain stress
inhomogeneity, slip transfer from neighboring grains, grain size,
Fig. 13. Effect of grain size and free surface on the CRSS for (a) the 18 grains that show and the influence of the free surface. Examination of the last two
prismatic slip and (b) the 22 grains that show basal slip. factors shows that the CRSS for slip is negatively correlated with
the grain size. Also, grains adjacent to the free surface show
lower CRSS for slip to be activated compared with grains from
the interior of the material.
grains< CRSSinner grains} to be valid is as high as 0.99.
From Fig. 13, it is concluded that both the grain size and the free Acknowledgements
surface will affect the CRSS for slip activity. Larger grains generally
require less resolved shear stress for prismatic or basal slip to be This work was supported by NSF Materials World Network
activated, which agrees with the Hall-Petch relation. Grains adja- grants NSF-DMR-1108211 and DMR-0710570 at Michigan State
cent to the free surface also require less resolved shear stress for University. Work at Carnegie Mellon University was supported by
prismatic or basal slip to be activated compared with grains in the the U.S. Department of Energy under Award DESC0002001. Use of
interior volume. This finding suggests that the common practice of the Advanced Photon Source was supported by the United States
using the surface slip activity characterized by SEM to represent the Department of Energy, Office of Science, Office of Basic Energy
behavior of the bulk material must be taken with some caution. Sciences, under Contract No. DE-AC02-06CH11357. LW acknowl-
Experimental measurement of CRSS of different slip modes edges the support of the National Natural Sciences Foundation of
provides important parameters for crystal plasticity modeling. In China (No. 51671127). JL was partly supported by Lawrence
610 L. Wang et al. / Acta Materialia 132 (2017) 598e610

Livermore National Laboratory under contract DE-AC52- [22] H.F. Poulsen, S.F. Nielsen, E.M. Lauridsen, S. Schmidt, R.M. Suter, U. Lienert,
L. Margulies, T. Lorentzen, D.J. Jensen, Three-dimensional maps of grain
07NA27344. RMS received support through the Visiting Scientist
boundaries and the stress state of individual grains in polycrystals and
program of the X-Ray Science Division, Argonne National Labora- powders, J. Appl. Crystallogr. 34 (2001) 751e756.
tory. The authors are grateful for the support from Jonathan Almer [23] H.F. Poulsen, Three-dimensional X-ray Diffraction Microscopy, Springer, Ber-
and Ali Mashayekhi of APS Beamline 1-ID during the beamtime. We lin, 2004.
[24] H.F. Poulsen, An introduction to three-dimensional X-ray diffraction micro-
thank Armand Beaudoin for his help with the Matlab code for far- scopy, J. Appl. Crystallogr. 45 (2012) 1084e1097.
field data analysis. JL and RMS also thank Shiu Fai Li for many [25] L. Margulies, T. Lorentzen, H.F. Poulsen, T. Leffers, Strain tensor development
helpful conversations and for his implementation of the forward in a single grain in the bulk of a polycrystal under loading, Acta Mater. 50
(2002) 1771e1779.
modeling, near-field reconstruction code. [26] R.V. Martins, L. Margulies, S. Schmidt, H.F. Poulsen, T. Leffers, Simultaneous
measurement of the strain tensor of 10 individual grains embedded in an Al
References tensile sample, Mater. Sci. Eng. A 387 (2004) 84e88.
[27] J.V. Bernier, N.R. Barton, U. Lienert, M.P. Miller, Far-field high-energy diffrac-
tion microscopy: a tool for intergranular orientation and strain analysis,
[1] J.J. Fundenberger, M.J. Philippe, F. Wagner, C. Esling, Modelling and prediction
J. Strain Anal. Eng. Des. 46 (2011) 527e547.
of mechanical properties for materials with hexagonal symmetry (zinc, tita-
[28] J. Oddershede, S. Schmidt, H.F. Poulsen, H.O. Sorensen, J. Wright, W. Reimers,
nium and zirconium alloys), Acta Mater. 45 (1997) 4041e4055.
Determining grain resolved stresses in polycrystalline materials using three-
[2] X. Tan, H. Guo, H. Gu, C. Laird, N.D.H. Munroe, Cyclic deformation behavior of
dimensional X-ray diffraction, J. Appl. Crystallogr. 43 (2010) 539e549.
high-purity titanium single crystals: Part II. Microstructure and mechanism,
[29] J. Oddershede, B. Camin, S. Schmidt, L.P. Mikkelsen, H.O. Sorensen, U. Lienert,
Metall. Mater. Trans. A 29 (1998) 513e518.
H.F. Poulsen, W. Reimers, Measuring the stress field around an evolving crack
[3] M.H. Yoo, Slip, twinning, and fracture in hexagonal close-packed metals,
in tensile deformed Mg AZ31 using three-dimensional X-ray diffraction, Acta
Metall. Trans. A 12 (1981) 409e418.
Mater. 60 (2012) 3570e3580.
[4] F. Bridier, P. Villechaise, J. Mendez, Analysis of the different slip systems
[30] C.C. Aydiner, J.V. Bernier, B. Clausen, U. Lienert, C.N. Tome, D.W. Brown,
activated by tension in a a/b titanium alloy in relation with local crystallo-
Evolution of stress in individual grains and twins in a magnesium alloy
graphic orientation, Acta Mater. 53 (2005) 555e567.
aggregate, Phys. Rev. B 80 (2009) 024113.
[5] L. Wang, Y. Yang, P. Eisenlohr, T.R. Bieler, M.A. Crimp, D.E. Mason, Twin
[31] T.R. Bieler, L.Y. Wang, A.J. Beaudoin, P. Kenesei, U. Lienert, In situ character-
nucleation by slip transfer across grain boundaries in commercial purity ti-
ization of twin nucleation in pure Ti using 3D-XRD, Metall. Mater. Trans. A 45
tanium, Metall. Mater. Trans. A 41 (2010) 421e430.
(2014) 109e122.
[6] S. Zaefferer, A study of active deformation systems in titanium alloys:
[32] L. Wang, J. Lind, H. Phukan, P. Kenesei, J.-S. Park, R.M. Suter, A.J. Beaudoin,
dependence on alloy composition and correlation with deformation texture,
T.R. Bieler, Mechanical twinning and detwinning in pure Ti during loading and
Mater. Sci. Eng. A 344 (2003) 20e30.
unloading - an in situ high-energy X-ray diffraction microscopy study, Scr.
[7] H. Li, D.E. Mason, T.R. Bieler, C.J. Boehlert, M.A. Crimp, Methodology for esti-
Mater. 92 (2014) 35e38.
mating the critical resolved shear stress ratios of a-phase Ti using EBSD-based
[33] H. Abdolvand, M. Majkut, J. Oddershede, S. Schmidt, U. Lienert, B.J. Diak,
trace analysis, Acta Mater. 61 (2013) 7555e7567.
P.J. Withers, M.R. Daymond, On the deformation twinning of Mg AZ31B: a
[8] B. Barkia, V. Doquet, J.P. Couzinie , I. Guillot, E. He
ripre
, In situ monitoring of
three-dimensional synchrotron X-ray diffraction experiment and crystal
the deformation mechanisms in titanium with different oxygen contents,
plasticity finite element model, Int. J. Plast. 70 (2015) 77e97.
Mater. Sci. Eng. A 636 (2015) 91e102.
[34] H. Abdolvand, M. Majkut, J. Oddershede, J.P. Wright, M.R. Daymond, Study of
[9] J.C. Schuren, P.A. Shade, J.V. Bernier, S.F. Li, B. Blank, J. Lind, P. Kenesei,
3-D stress development in parent and twin pairs of a hexagonal close-packed
U. Lienert, R.M. Suter, T.J. Turner, D.M. Dimiduk, J. Almer, New opportunities
polycrystal: Part I e in-situ three-dimensional synchrotron X-ray diffraction
for quantitative tracking of polycrystal responses in three dimensions, Curr.
measurement, Acta Mater. 93 (2015) 246e255.
Opin. Solid State Mater. Sci. 19 (2015) 235e244.
[35] S.L. Wong, J.-S. Park, M.P. Miller, P.R. Dawson, A framework for generating
[10] A.A. Salem, S.R. Kalidindi, S.L. Semiatin, Strain hardening due to deformation
synthetic diffraction images from deforming polycrystals using crystal-based
twinning in a-titanium: constitutive relations and crystal-plasticity modeling,
finite element formulations, Comp. Mater. Sci. 77 (2013) 456e466.
Acta Mater. 53 (2005) 3495e3502.
[36] D.C. Pagan, M.P. Miller, Connecting heterogeneous single slip to diffraction
[11] X. Wu, S.R. Kalidindi, C. Necker, A.A. Salem, Prediction of crystallographic
peak evolution in high-energy monochromatic X-ray experiments, J. Appl.
texture evolution and anisotropic stressestrain curves during large plastic
Crystallogr. 47 (2014) 887e898.
strains in high purity a-titanium using a Taylor-type crystal plasticity, Acta r, G. Ribarik, G. Zilahi, R. Mulay, U. Lienert, L. Balogh, S. Agnew, Slip
[37] T. Unga
Mater. 55 (2007) 423e432.
systems and dislocation densities in individual grains of polycrystalline ag-
[12] L. Wang, R.I. Barabash, Y. Yang, T.R. Bieler, M.A. Crimp, P. Eisenlohr, W. Liu,
gregates of plastically deformed CoTi and CoZr alloys, Acta Mater. 71 (2014)
G.E. Ice, Experimental characterization and crystal plasticity modeling of
264e282.
heterogeneous deformation in polycrystalline alpha-Ti, Metall. Mater. Trans. A
[38] J. Oddershede, J.P. Wright, A. Beaudoin, G. Winther, Deformation-induced
42 (2011) 626e635.
orientation spread in individual bulk grains of an interstitial-free steel, Acta
[13] Y. Yang, L. Wang, T.R. Bieler, P. Eisenlohr, M.A. Crimp, Quantitative atomic
Mater. 85 (2015) 301e313.
force microscopy characterization and crystal plasticity finite element
[39] U. Lienert, S.F. Li, C.M. Hefferan, J. Lind, R.M. Suter, J.V. Bernier, N.R. Barton,
modeling of heterogeneous deformation in commercial purity titanium,
M.C. Brandes, M.J. Mills, M.P. Miller, B. Jakobsen, W. Pantleon, High-energy
Metall. Mater. Trans. A 42 (2011) 636e644.
diffraction microscopy at the advanced photon source, JOM 63 (2011) 70e77.
[14] J.L.W. Warwick, N.G. Jones, K.M. Rahman, D. Dye, Lattice strain evolution
[40] S. Schmidt, GrainSpotter: a fast and robust polycrystalline indexing algorithm,
during tensile and compressive loading of CP Ti, Acta Mater. 60 (2012)
J. Appl. Crystallogr. 47 (2014) 276e284.
6720e6731.
[41] R.M. Suter, D. Hennessy, C. Xiao, U. Lienert, Forward modeling method for
[15] M. Knezevic, R.A. Lebensohn, O. Cazacu, B. Revil-Baudard, G. Proust, S.C. Vogel,
microstructure reconstruction using x-ray diffraction microscopy: single-
M.E. Nixon, Modeling bending of alpha-titanium with embedded polycrystal
crystal verification, Rev. Sci. Instrum. 77 (2006) 123905.
plasticity in implicit finite elements, Mater. Sci. Eng. A 564 (2013) 116e126.
[42] C.M. Hefferan, S.F. Li, J. Lind, U. Lienert, A.D. Rollett, P. Wynblatt, R.M. Suter,
[16] D. Gloaguen, G. Oum, V. Legrand, J. Fajoui, S. Branchu, Experimental and
Statistics of high purity nickel microstructure from high energy X-ray
theoretical studies of intergranular strain in an alpha titanium alloy during
diffraction microscopy, Comput. Mat. Contin. 14 (2009) 209e219.
plastic deformation, Acta Mater. 61 (2013) 5779e5790.
[43] S.F. Li, R.M. Suter, Adaptive reconstruction method for three-dimensional
[17] K.E.K. Amouzou, T. Richeton, A. Roth, M.A. Lebyodkin, T.A. Lebedkina, Micro-
orientation imaging, J. Appl. Crystallogr. 46 (2013) 512e524.
mechanical modeling of hardening mechanisms in commercially pure a-ti-
[44] J. Lind, S.F. Li, R. Pokharel, U. Lienert, A.D. Rollett, R.M. Suter, Tensile twin
tanium in tensile condition, Int. J. Plast. 80 (2016) 222e240.
nucleation events coupled to neighboring slip observed in three dimensions,
[18] J. Gong, A.J. Wilkinson, Anisotropy in the plastic flow properties of single-
Acta Mater. 76 (2014) 213e220.
crystal a titanium determined from micro-cantilever beams, Acta Mater 57 , H.R. Wenk, Texture and Anisotropy, Cambridge Uni-
[45] U.F. Kocks, C.N. Tome
(2009) 5693e5705.
versity Press, Cambridge, 2000.
[19] J. Gong, A.J. Wilkinson, Micro-cantilever testing of 〈a〉 prismatic slip in
[46] Y. Guo, T.B. Britton, A.J. Wilkinson, Slip bandegrain boundary interactions in
commercially pure Ti, Philos. Mag. 91 (2011) 1137e1149.
commercial-purity titanium, Acta Mater. 76 (2014) 1e12.
[20] M.D. Uchic, D.M. Dimiduk, J.N. Florando, W.D. Nix, Sample dimensions influ-
[47] L. Nervo, A. King, A. Fitzner, W. Ludwig, M. Preuss, A study of deformation
ence strength and crystal plasticity, Science 305 (2004) 986e989.
twinning in a titanium alloy by X-ray diffraction contrast tomography, Acta
[21] E.M. Lauridsen, S. Schmidt, R.M. Suter, H.F. Poulsen, Tracking: a method for
Mater. 105 (2016) 417e428.
structural characterization of grains in powders or polycrystals, J. Appl.
Crystallogr. 34 (2001) 744e750.

You might also like