ECCI Based Characterization of Dislocation Shear in Polycrystalline Arrays During Heterogeneous Deformation in Commercially Pure Titanium

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 28

ECCI based characterization of dislocation

shear in polycrystalline arrays during


heterogeneous deformation in
commercially pure titanium
Songyang Han, Philip Eisenlohr, Martin A. Crimp

Chemical Engineering and Materials Science, Michigan State University, East Lansing, MI 48824, USA

Abstract
A new electron channeling contrast imaging (ECCI) based approach for the characterization of
heterogeneous plastic deformation across multiple grains in a polycrystal is presented. The
nature of the shear accommodation at grain boundaries has been assessed by evaluating the
spatial distribution of dislocations and slip system activities near the boundaries and relating
these observations to standard slip transfer criteria and the global state of stress. The direction
of slip band propagation, and consequently the location of dislocation slip band nucleation, was
determined by characterizing the slip band widths across individual grains. This assessment
leads to a better understanding of the factors involved in polycrystalline deformation, including
the grain boundary shear accommodation associated with both dislocation pile-ups and grain
boundary dislocation nucleation. Because ECCI allows study of dislocation structures over large
surface areas, and thus larger overall volumes than typically are accessible by other techniques,
the overall approach allows the dislocation distribution across individual grains to be understood
in terms of the broader polycrystalline constraints.
Key words

ECCI, heterogeneous plastic deformation, boundary shear accommodation, slip band


propagation direction

1. Introduction
Understanding heterogeneous deformation of polycrystalline materials is critical to elucidating
the development of overall plasticity and damage accumulation in the material. One of the
critical aspects of understanding heterogeneous deformation is the role of grain boundaries in
maintaining polycrystalline compatibility, i.e. arbitrary grain shape change, which, as
demonstrated by Von Mises, requires a minimum of 5 independent deformation modes [1]. At
the dislocation scale, this compatibility requires the ability of grain boundaries to accommodate
shear in one grain with shear in its neighboring grain. Failure to do so leads to stress
concentrations and potential damage nucleation. Shear accommodation at grain boundaries has
been studied by a large number of researchers and has led to a number of criteria that describe
the geometric compatibility of the deformation systems associated with shear accommodation
at grain boundaries. These criteria take into account the alignment of the slip planes, slip
directions, and slip plane intersection lines with the grain boundaries in different ways (Fig 1),
and include the Livingston and Chalmers factor N [2], Shen et al.’s criteria [3] that introduced the
factor M in combination with N, the LRB factor [4, 5], and the more convenient geometric
compatibility factor m’ developed by Luster and Morris [6]. Generally, these criteria have been
found to correlate to some degree with experimental observations of shear transfer [7-10].
Nevertheless, these criteria have been developed in the context of dislocation shear in one grain
approaching a boundary that must accommodate that strain, while in practice nucleation of the
same dislocations directly from boundaries [11] will also require accommodation. Furthermore,
understanding heterogeneous deformation requires a broader understanding of the mesoscale
deformation and cannot simply be offered by these single boundary criteria. Consequently, a
better understanding of how the interrelated deformation activity across mesoscale
polycrystalline arrays occurs is desirable.

For the most part, these well-established slip transfer mechanisms have been studied by
characterizing the nature of dislocations through contrast analysis using transmission electron
microscopy (TEM) or high-resolution TEM [2, 4-6, 12, 13], but TEM is typically destructive and
limited to small areas. Another approach that relates the geometry of shear accommodations to
strain accumulations in the vicinity of grain boundaries has been recently developed using high-
resolution electron backscatter diffraction (HR-EBSD) [14, 15]. However, this approach is not yet
capable of fully characterizing the dislocation structure, as it only reveals the geometrically
necessary dislocations (GNDs) and not all of the dislocations involved in the deformation
process, i.e. the statistically stored dislocations (SSDs) [16]. Alternatively, because electron
channeling contrast imaging (ECCI) [17-21] offers the ability to study the near-surface
dislocations across a bulk sample, it is often non-destructive since in many cases sample
preparation is carried out prior to treatments such as deformation. Furthermore, because it is
not limited to thin foil regions, it often allows comprehensive characterization of dislocation
morphologies across comparably large examination areas. This technique also allows the nature
of dislocations to be determined in terms of the Burgers vector b and line direction u, using g · b
= 0 and g · b × u = 0 invisibility criterion and line trace analysis, by establishing specific channeling
conditions g with electron channeling patterns (ECPs), typically for very large grains or single
crystals, or selected area channeling patterns (SACPs) for polycrystalline samples [19, 21-27].
With its broader length scale, ECCI thus becomes a perfect tool for linking dislocation activity
with deformation evolution across polycrystalline arrays.

Because of their more complex dislocation behaviors compared to cubic metals, understanding
heterogeneous deformation is of critical interest for structural hexagonal metals, including
titanium and its alloys. The dislocation slip in hexagonal (α) titanium is typically dominated by
<a> slip on prism and basal planes, but additional activation of <c + a> type slip on a number of
planes occurs [28, 29], because it is necessary to activate slip in the c-direction to satisfy the von
Mises criterion. This suggests that a number of dislocation slip systems may be involved in the
overall deformation of grains, but other slip systems may be readily activated in association with
grain boundary shear accommodation.

The present study outlines an approach for characterizing how the plasticity of a polycrystalline
array develops due to dislocation nucleation, slip, and shear accommodation across grains and at
grain boundaries. This characterization uses ECCI to investigate the details of the dislocations
involved in grain boundary shear accommodation and the nature of the slip bands associated
with heterogeneous deformation. An important aspect of this work is that it not only studies
grain boundary shear accommodation in relation to slip propagating into and piling up at grain
boundaries, but also considers the shear accommodation necessary in neighboring grains when
dislocations nucleate from an unknown source at a boundary, which is reported as grain
boundary ledge sources at interfaces [11]. The results indicate that at some grain boundaries
shear accommodation is necessitated by easy nucleation and slip in soft grains on one or both
sides of a boundary, even if the accommodation is not geometrically advantageous, while slip
transfer at other boundaries is dominated by good geometric compatibility despite low resolved
stresses and high critical resolved shear stresses for the accommodating system.

2. Experiment Procedure
Commercially pure titanium was cut into a small bar approximately 30 mm x 2 mm x 3 mm. This
sample was ground through a standard series of steps down to SiC 4000 grit, and then polished
with five-to-one ratio of 0.05 μm colloidal SiO2 suspension and 30% hydrogen peroxide for 15
minutes, followed by ultrasonic cleaning in alcohol then deionized water. This process removed
approximately 1 micron of material, based on the area reduction of Vickers indentations.
Backscattered electron imaging prior to deformation revealed that most of the grains were
effectively stress-free, but that a small number of grains displayed significant bend contour
contrast, suggesting presence of residual lattice strain. A few of these grains are evident in
Figure 2 a and b, one in the upper right and one near the center left, but are not the focus of this
study. The sample was deformed to approximately 1.2 % tensile surface strain using a four-point
bending stage (all images have the surface tensile stress axis horizontal). After characterizing the
slip bands that developed during the deformation, the residual surface topography from the
deformation was gently removed by re-polishing with the same colloidal SiO2 solution for 10
seconds, removed approximately 0.01 microns of surface material. The surface was then
electropolished using a solution of 10 wt% magnesium perchlorate in 90 wt% methanol solution
using a potential of 24 V at -30 oC, in order to remove residual near-surface hydrides that formed
during the colloidal SiO2 polishing. The surface removal rate during electropolishing was
determined based on the electropolishing current, the electropolished area, the charge
difference of 4 between the atomic and ionic states (i.e. four electrons per ionized atom), and
the atomic volume of hexagonal titanium of 0.0176 nm3. Post-deformation and post-
electropolishing backscattered electron images are shown in Figure 2 a & b. The original grains
in the microstructure are still evident after the electropolishing, but the new cross-section
reveals some modification in the appearance of the grains. Prior to deformation and following
electropolishing, the grain orientations were mapped by electron back scattered diffraction
(EBSD) using a TESCAN MIRA III field emission gun (FEG) scanning electron microscope (SEM)
equipped with an EDAX EBSD system using a 30 kV accelerating voltage, a 148 µA probe emission
current, a 20.0 nm spot size, and an 18 mm working distance and a specimen current of 3.1 nA
with the sample tilted to 70° (varying step sizes were used). EBSD maps collected before and
after electropolishing are shown in Figure 3 a & b. Any noticeable differences in grain
orientations in the EBSD maps appear systematic among the grains and can be attributed to
slightly different sample mounting orientations in the two EBSD mapping scans. This has been
confirmed by comparing the Euler angles for each grain before and after electropolishing, which
revealed a systematic rotation of ~4° about Z and ~1° about X across the entire region of the
study. The active slip planes were determined by standard slip trace analysis [30-32] based on
the EBSD determined grain orientations1. The geometrical compatibility factor, m’, between
grains and the Schmid factors, m, of the various slip systems were determined based on the
observed active slip systems and the EBSD measured grain orientations and were calculated by
inputting the Euler angles from EBSD measurements in each grain into an in-house developed
matlab code. The distribution of dislocations in the slip bands and the dislocation-boundary
interactions near grain boundaries were characterized using electron channeling contrast
imaging (ECCI), which facilitated identification of line directions and Burgers vectors through
trace analysis and g ∙ b = 0 invisibility experiments respectively. Specific channeling conditions
were established using SACPs facilitated by beam rocking mode in TESCAN MIRA III SEM. ECCI
was carried out at under the same microscope operating conditions, but with a working distance
of ~10 mm (depending on the stage tilt of -10° to 10° and rotation to get the desired diffraction
condition). Each ECC image of 768 x 768 pixels was acquired in about 10 min using a scan speed
of 1 k pixel/s.

3. Results
This study will focus on the slip behavior in a microstructure patch shown in Figure 2 a, with an
emphasis on the slip behavior of three grains in the center of the patch. The slip through these
grains is linked through the slip bands that propagate through four grains. The slip bands were
quite distinct in the as-deformed material, but some of the slip bands became less distinct as
they approached grain boundaries (noted with a white arrow in Fig 2 a), suggesting either
decreasing slip band height or spreading of the height over a larger area. While in many grains
the slip bands were very straight/planar (i.e. in grain 2 and 3 in Figure 2 a), in other grains, the

1
In titanium, with the exception of basal slip, there is one unique slip direction for each slip plane, so that the
determination of the slip plane by EBSD based analysis allows identification of the dislocation slip direction.
bands were wavier (for example in grain 1), suggesting some cross-slip may have occurred.
Regardless of the overall nature of the slip bands, the slip bands in neighboring grains meet at
common points at grain boundaries between grains 1 and 2 and between 3 and 4, as is clearly
observed in Figure 2. In many cases, these intersections appear to be associated with the more
distinct slip bands. In contrast, most of the slip bands interacting with the boundary between
grains 2 and 3 do not intersect at common points on the boundary (noted by the red arrow in Fig
2 a), but some evidence of intersection exists. Regardless, the general propagation of the slip
bands across these four grains suggests that shear in one grain is accommodated by shear in a
neighboring grain, i.e. some degree of slip compatibility has occurred across these grain
boundaries.
3.1 Electron backscatter diffraction (EBSD) and slip trace analysis
EBSD based slip trace analysis [31] has been used to identify the predominant slip traces
observed in grains 1, 2, and 3 (Fig 4). In grains 2 and 3, slip bands on a single system were
predominant, with some of these bands appearing more active than others, and no slip bands
from other slip systems were observed. In contrast, in grain 1, the slip traces were much wavier.
Because of the waviness of the slip in grain 1, it was impossible to define a slip plane. In grains 2
and 3, the slip bands were very close to the traces for multiple slip planes, i.e. (11̅00) and (1̅101)
in grain 2 and (101̅0) and (2̅112) in grain 3. Therefore, it was not possible to fully attribute the
slip bands to specific systems solely on the basis of slip trace analysis. The waviness of the slip
bands in grain 1 makes trace analysis more difficult and four different potential slip planes were
identified (Fig 4 a).
3.2 Electron Channeling Contrast Imaging (ECCI)
Because the slip directions cannot be fully determined from the trace analysis, ECCI was used to
determine the Burgers vector of the dislocations in the slip bands. This analysis was carried out
on the sample after electropolishing by imaging under different electron channeling conditions.
This ECCI analysis also facilitated a detailed assessment of dislocation distributions in the slip
bands of interest in grains 1, 2, and 3. Figure 5 a shows an ECC image of the slip bands in grain 3,
along with a series of five higher magnification ECC images taken at different channeling
conditions2. The channeling images show individual dislocations, which in this case appear as
dots, short lines, and long lines, depending on their line orientations, dispersed along the slip
band. There are very few dislocations outside of this densely populated slip band. From the
strong visibility of the dislocations for g= [2̅112], [1̅101̅], [1̅100], and [12̅12] and weak contrast
for g= [1̅013] it can be concluded that the Burgers vector of the dislocations in the slip band (as
well as the other scattered dislocations) is [12̅10]. This Burgers vector is consistent with the
(101̅0) plane that was one of the two possible planes identified in the EBSD based slip trace

2
Because the slip line contrast comes predominantly from the strain fields from dislocations, the slip lines in the
electropolished sample cannot be observed in secondary electron mode or generalized backscattered imaging that is
not under a proper channelling condition. The signal to noise is different in different channelling condition due to
different BSE yield under each condition.
analysis, and not consistent with the other possible plane, (2̅112). Analysis of the dislocations in
other slip bands in this grain revealed the same dislocation types with similarly well-defined band
morphologies. The same approach has been applied to the slip bands in grains 1 and 2, both in
the centers of the grains and at the slip band intersections with the boundaries. The active slip
bands in grain 1 were predominantly identified as [112̅0] (11̅00) (Fig 6 a), with the same
dislocation slip band types in grain 2 (Fig 6 b). These analyses also confirm that the dislocations
in the well-defined slip bands in the centers of the grains have the same Burgers vectors as those
involved in the intersections of the bands with the boundaries.

The dislocation distribution morphologies in the slip bands appeared modified near the grain
boundaries. Figure 73 shows the intersection of a slip band in grain 3 at the boundary with grain
2. While the dislocations in grain 3 (right side of the Fig 7 a) have the same Burgers vector as
those in the center of the grain (Fig 5), they are less densely packed and appear to be limited to a
smaller slip band width than in the center of the grain. Contrasting this is the dislocation
morphology on the grain 2 side of this boundary in Figure 7 b, which shows a high density of
dislocations spreading along the boundary. Figure 7 c shows both of the grains in reasonable EC
contrast. It can be observed that there is a distance between the intersection points of each slip
line and the grain boundary, suggesting a more complex mechanism of strain transfer rather
than the direct transmission of dislocations. It appears that the dislocations in grain 3 have been
generated at a dislocation source at/near the boundary as a result of a stress build-up (pile-up
stress) (Fig 7 c). This assertion is grounded in the fact that the dislocations near the grain
boundary are confined to a very narrow slip band, which broadens as the band extends in to the
grain. It is reasonable to expect dislocations moving from any source to spread out around other
dislocations and other obstacles via cross-slip, resulting in broadening of the overall slip band as
the distance from the source increases. This is consistent with recent work by Murr [11], who
used TEM to study the dislocation emission from grain boundary ledges and showed very narrow
slip bands propagating from these grain boundary dislocation sources.
Once activated at the boundary between grains 2 and 3, the dislocations in grain 3 have
propagated away from the boundary on essentially a single plane, but the slip band broadens
towards the center of the grain as cross-slip around different (unknown) barriers has occurred,
resulting in the observed more tangled, broader slip band (Fig 5). Even more slip band
broadening is observed towards the boundary with the next grain. This is consistent with the
argument that localized strain by a blocked slip band at the boundary can be relieved by diffuse
distribution of dislocations along grain boundaries [14]. This broadening of the slip band as it
propagates across grain 3 is illustrated in Figure 8, which shows the overall slip band morphology
in Figure 8 a (electropolished so that most of the contrast is generated from the dislocation

3
Because of the different orientations of the grains on either side of each boundary, the channeling conditions
necessary for imaging the dislocations can typically only be established on one side of the boundary at a time. As a
result, in many of the images, the dislocations are in good contrast on one side of the boundary, but the other side
often displays extreme black or white contrast.
strain in the slip bands at a specific channeling condition) and the broadening at positions b, c,
and d. This particular band increases in width from approximately 0.12 µm at position b, to
around 0.5 µm at position c in the middle of the grain, and approximately 2.6 µm near the
intersection with the boundary with grain 4 (position d)4. Another interesting feature of the
dislocations in the slip bands as a function of position can be noted. Based on the sense of the
black/white contrast being the same for the vast majority of the dislocations in grain 3 near the
boundary with grain 2, it can be concluded that these majority dislocations in the slip band have
the same sign Burgers vector (over 85% in Fig 8 b). However, as the slip band propagates, larger
fractions of the dislocations display opposite contrast sense, indicating there is a larger fraction
of dislocations with opposite Burgers vectors present in the slip band in the center of the grain
(around 35% dislocations with opposite Burgers vector shown in Fig 8 c) and even larger
fractions of these opposite Burgers vector dislocations near the intersection with grain 4 (around
75% dislocations with opposite Burgers vector shown in Fig 8 d). The reasons for this change in
fraction of opposite Burgers vectors may be a result of more dislocation loop formation
occurring in association with the increase in dislocation numbers in the slip bands (i.e. more
cross-slip and dislocation bowing may be occurring as the dislocations propagate further along
the band). Similar band broadening has been observed in other grains, suggesting that it is
possible to determine in which direction a slip band propagated and consequently, the direction
in which strain transfer occurred at a given boundary. While it is difficult to determine the slip
propagation direction in grain 1 based on slip band broadening, the matter in which the slip
bands have branched as they move away from the boundary with grain 2 (Fig 4 a) suggests that
the slip bands in grain 1 nucleated at the boundary with grain 2. It appears that the slip bands in
grain 1 nucleated from the highly strained areas on the boundary indicated by the bright areas
on the boundary (highlighted in Fig 9 b) with a width of approximately 0.3 µm. The details of the
dislocations in the branching slip bands are shown in Figure 9 c, suggesting the width of one of
the dominant slip bands nucleating from the boundary is increased to around 2 µm. As these
bands propagated further into the grain, the dislocations continue to spread into less distinct
bands as shown in Figure 9 d, with a width of around 4.5 µm. This analysis suggests that the
dislocations in both grain 1 and 2 have nucleated at the boundary between them and
propagated away from the boundary.

4. Discussion
Shear transfer between grains as envisioned by the standard slip transfer criteria [2-6], may
involve a number of interrelated processes that reflect the resistance of the boundary to
incorporation of incoming dislocations, including the potential for reactions between the
incoming dislocations and grain boundary dislocations, the possibility of the motion of existing or

4
It should be noted that width measurements of slip bands should only be used to make relative comparisons within
a given band or to another band of the same slip system in the same grain using the same channelling conditions, as
variations in slip plane orientations and changes in channelling condition will both affect the projected width of the
slip bands.
newly created grain boundary dislocations, and the activation of dislocation sources that creates
new dislocations operating on new slip systems in the neighboring grain (or even in the parent
grain). Based on these well-established criteria, it is possible to have two slip systems with high
Schmid factors and low critical resolved shear stresses (for example, prism <a> slip, which in
commercially pure Ti has the lowest CRSS [28, 33-34], slightly lower than <a> slip on basal planes
[35-37], a factor of 4-10 times lower than pyramidal <c + a> slip [36-37], and a factor ranging
from 1.8-8 less than pyramidal <a> slip [34-38] on either side of a boundary that are,
nonetheless, poorly aligned so that the slip transfer criteria are low (i.e. the m’ becomes low).
This suggests that the highly active shear in one grain will not be compatible with highly active
shear in the other grain. Conversely, two slip systems may be well aligned, for example with m’
approaching 1, but if these systems have low Schmid factors and/or high critical resolved shear
stresses, dislocations are unlikely to be extensively activated in the incoming grain and
dislocations in the outgoing grain will not have a strong driving force to move away from the
boundary. While m’ appears to be associated with the activation of the different slip systems
associated with grain boundary strain accommodations, it should be pointed out that m’ does
not consider the influence of character of grain boundary (i.e. degree of tilt or twist boundary)
[5, 6], which might be expected to influence slip transfer. Nevertheless, to facilitate a discussion
on the factors that may be involved in shear transfer and slip band activity, the Schmid factor
and m’ values associated with the grains and grain boundaries examined in this study are
outlined in Table 1. As outlined above, the slip system activated in grain 1 is the [112̅0] (11̅00)
prism <a> system, which has a Schmid factor of 0.15; the cross-slip system, [1̅1̅20] (1̅101), which
has a much higher CRSS, has a Schmid factor of 0.23. The prism <a> [112̅0] (11̅00) slip system is
activated in grain 2, with 0.41 Schmid factor; the prism <a> [12̅10] (101̅0) slip system is activated
in grain 3, with 0.48 as the Schmid factor. The m’ between the observed systems at the
boundary between grains 1 and 2 is 0.83, indicating good alignment between the active systems,
while the respective m’ between the grains 2 and 3 is 0.42, evidence of relatively poor
alignment.
The active prism <a> systems in grains 2 and 3 both have high global Schmid factors and show
extensive, planar slip, but the low m’=0.42 between them suggests shear will not be easily
transferred between these grains. This is consistent with the observation that, with one
exception (noted by a red arrow in Fig 2 a), the slip bands in these two grains are not well
correlated, i.e. they do not meet at one point along the boundary. Perhaps, these two grains do
not experience extensive deformation near the boundary, consistent with some of the slip bands
in grain 2 fading near the boundary with grain 3, or perhaps other mechanisms, such as
dislocation motion in the boundary, may be assisting the strain transfer between grains to
maintain boundary integrity. Nevertheless, the activation of the observed slip systems in grains
2 and 3 does not appear to be associated with shear accommodation between these grains.
In contrast, slip bands in grains 2 and 1 line up much more precisely at their boundary than at
the boundary between grains 2 and 3. This apparent shear accommodation is consistent with
the much larger m’ value 0.83 between the active prism <a> slip systems in grains 1 and 2 than
the 0.42 between the active prism <a> systems in grains 2 and 3. That is, the need for
compatibility with slip nucleating at the boundary and gliding easily in grain 2 may result in the
activation of the dislocations observed near the grain boundary in grain 1, but these low Schmid
factor dislocations (0.15) in grain 1 may have difficulty moving away from the boundary. This
may be evidenced by the less distinct slip bands in grain 1 that do not extend across the entire
grain. The cross-slip observed in this grain might be manifested by the accommodating prism
<a> dislocations cross-slipping onto the pyramidal <a> system of higher Schmid factor (m = 0.23),
despite the significantly higher CRSS of this system [39-41]. While T1 twins and several
pyramidal <c + a> slip systems have somewhat higher global Schmid factors (less than twice)
than any of the observed <a> dislocation slip systems in grain 1, none of these have high
compatibility with the primary slip systems in grain 2. Given that the CRSS of T1 twinning and <c
+ a> pyramidal slip systems are much higher than those of the observed prism <a> slip, it is not
surprising that these T1 twins and pyramidal <c + a> slip systems are not activated based on a
combination of the global Schmid factors and critical shear stresses. Overall, it appears that the
deformation in grain 1, which in general is not well oriented for slip under the global state of
stress, deforms near the grain boundary in response to the slip compatibility necessary to
accommodate the easily activated slip in grain 2.

To understand the deformation of polycrystalline materials, it is also interesting to consider the


grains in this microstructure patch as a whole, rather than as individual grains. Characterizing
the deformation evolution throughout this patch, including the accommodations at the
boundaries, will provide insight to facilitate how the entire arrangement reacts to external
stress.

ECCI analysis has been performed both along the predominant slip bands as well as on less
distinct bands, in order to investigate whether the slip bands propagate in the same direction in
all of the bands observed in a given grain. The broadening of the dislocation fields along the slip
bands revealed by ECCI suggests that the four dominant slip bands in grain 3 (Fig 10) all
propagated in the same direction, from the boundary with grain 2 toward the boundary with
grain 4, with their broadening following a similar trend from around 0.2 to 3 µm in width across
grain 3. There is, however, a less distinct band of the same prism <a> slip system that appears to
nucleate from the boundary with grain 4 and broadens in the opposite direction of the dominant
bands (Fig 10 b-d). As this band propagates towards to the upper left, it continues to broaden
and becomes indistinct in the middle of the grain. The width of such slip band starts from less
than 0.1 µm at position b to 1 µm at position c, and finally to approximately 4.5 µm at position d,
where the slip band becomes ‘physically invisible’ at lower magnifications as shown in Figure 10
a. It can also be observed that some additional dislocation slip bands nucleate from the
boundary between grains 3 and 4 on different slip systems with 0.1 µm width, but they only
propagate a short distance of 3 to 4 µm from the boundary reaching less than 1 µm in width,
which suggests these dislocations are involved in accommodating shear at the boundary rather
than the broader deformation of the grain. Overall, it appears that grain 3 deformed primarily
through the motion of [12̅10] (101̅0) prism <a> type dislocations that nucleated at the boundary
with grain 2. Nevertheless, accommodation of the deformation associated with other grains is
also necessary, in this case illustrated by the additional activation of slip systems from the
boundary with grain 4.

Likewise, the slip system observed in the dominant slip bands in grain 2 is the [112̅0] (11̅00)
prism <a> slip system, which has a relatively high Schmid factor (0.41); these slip bands show
broadening from the boundary between grain 1 and 2 towards the boundary between grains 2
and 3 (Fig 11 a), suggesting the deformation shear is nucleated from the boundary between
grains 1 and 2 and propagates towards the lower right in grain 2. The broadening of slip bands
also follows a similar trend, which starts from approximately 0.02 µm (Fig 6 a) and ends up with a
diffused, 3 µm wide band along the grain boundary with grain 3 (Fig 7 b). Contrasting the
dominant slip bands in grain 2, additional slip systems nucleating at the boundary between grain
2 and 3 (see the red square region in Fig11 a) that display weak contrast and only propagate less
than 10 µm from the grain boundary (Fig 11 b) have been observed. The morphology of these
systems is also different from that of dominant slip bands, as shown in Figure 11 c, revealing a
high density of dislocations spreading out rapidly from a point. The band width is approximately
0.2 µm close to the grain boundary, 1.2 µm in the grain interior about 3 µm away from the
boundary, and approximately 2.3 µm at around 7 µm from the nucleation area, where further
measurements of the band width beyond this area become impossible due to the resolution loss
of dislocations in large area scans. This apparent nucleation point coincides with the point
where a dominant slip band in grain 3 has nucleated (this point is noted by open arrows in Fig 11
b-d), indicating dislocations move away from the boundary from this point in both grains. Figure
11 d shows this common point with dislocations on both sides of the grain boundary in
reasonable EC contrast. This image shows very strong contrast close to the boundary in grain 2,
suggesting significant local strain associated with the shear accommodation. ECCI reveals that
the Burgers vectors of these secondary slip system dislocations in grain 2 are [21̅1̅3̅], but
because these dislocations are spread out, they do not clearly define a slip plane. This variability
in the slip plane is consistent with observations of cross slip of <c +a> near grain boundaries
[42,43]. As noted above, these <c + a> dislocations are expected to have significantly higher
CRSS than <a> slip systems. Unfortunately, because the slip plane of these dislocations is not
well defined, it is impossible to evaluate the m’ associated with the activation of these
dislocations in response to the shear associated with the [12̅10] (101̅0) prism <a> slip system in
grain 3. Nevertheless, as these dislocations do not propagate a large distance from the
boundary, it is likely that the activation of <c + a> dislocations near the boundary is a
consequence of shear accommodation with the dislocations activated in grain 3. It is also worth
noting that the dominant [112̅0] (11̅00) prism <a> system in grain 2 has a relatively low m’ (0.42)
with the active [12̅10] (101̅0) prism <a> slip system in grain 3, indicating that strain
accommodation between those two systems is unlikely.
Overall, the complex deformation of the polycrystalline patch can be assessed. Because grains 2
and 3 have prism <a> slip systems with similarly low CRSS and oriented with high Schmid factors,
it is reasonable to expect that these two grains would deform early, with perhaps grain 3 slipping
before grain 2 because the Schmid factor for the observed prism slip is somewhat higher in grain
3 (this will also depend on the stresses necessary to nucleate the dislocations at the grain
boundaries). In contrast, grain 1 does not display sharp, planar slip bands as no low CRSS slip
systems are well oriented for high Schmid factors. The activation of the prism <a> systems in
grains 2 and 3 will lead to stress accumulation at their boundaries because the shear in the two
grains is not well accommodated, as indicated by the low m’ between these two systems.
Consequently, nucleation of additional slip systems that can accommodate this grain boundary
stress build-up is necessary. These accommodating systems must have high m’ with these
dominant systems, even if they have relatively low ratios of Schmid factor to CRSS under the
global stress state. Because of this latter condition, these accommodating secondary slip
systems often do not propagate extensively across the grains. This grain boundary strain
accommodation is observed to occur with high m’ slip systems in grain 1 near the boundary with
grain 2, in grain 2 near the boundary with grain 3, and in grain 3 near the boundary with grain 4.
These accommodating slip systems often show less well-defined slip bands than the slip bands
associated with generalized shear of the grains.

The difference in the activation of slip systems as a result of the global stress versus the shear
accommodation suggest that the state of stress involved in these processes evolves differently.
At onset of deformation, the activation of dislocation systems is associated with the most highly
stressed slip systems (resolved global stress) and those with the lowest critical resolved shear
stresses in relation to this resolved stress. As the deformation continues, the local stress state
may evolve differently in different regions of a grain, especially in the vicinity of grain
boundaries, and becomes more heterogeneous than the global state of stress, partially as more
dislocations pile-up at boundaries and as the need to maintain grain boundary integrity grows.
Such complicated local stress state may initiate multiple accommodating slip systems, some of
which may not be favorable under the global stress state and consequently may not propagate
deep into the grain, where the stress state may be different. One would expect this transition in
slip system activity to be gradual, as the state of stress will gradually change between different
regions in a grain. One would also expect this stress state evolution and dislocation system
activity to evolve during the continuing deformation of the polycrystal as a result of the
activation of additional slip systems, local hardening, and grain rotations.

As would be expected, the details of the deformation of a polycrystalline patch are quite
complicated. The approach outlined here has allowed characterization of the specifics of the
dislocation morphology associated with generalized deformation and polycrystal strain
accommodation to be revealed. Specifically, the details of dislocation morphologies within slip
bands have been characterized, facilitating determination of the slip band nucleation points and
propagation directions. Such information is critical for understanding the development of
heterogeneous deformation, and potentially damage nucleation in the form of local work
hardening and crack nucleation, in polycrystalline arrays. Furthermore, this approach has the
potential for characterizing the distribution of dislocation sources. It is important to note that
the slip band morphology observed in this study suggests that all of the dislocation slip bands
nucleated from grain boundaries i.e. dislocations in the slip bands always spread out with
distance from grain boundaries. This does not exclude, however, that in other materials or in
other regions of this sample dislocations nucleate within the grains, which would be observed as
narrow slip bands within the grains that spread as they extend towards the grain boundaries.

5. Summary
The heterogeneous deformation of a polycrystalline microstructural patch has been
characterized using a new ECCI based approach for assessing grain boundary shear transfer and
slip band propagation. The direction of shear band propagation and, consequently, the shear
band nucleation locations have been characterized by assessing the width of shear bands across
grains. Grain boundary shear accommodation has been assessed for both boundaries that have
reacted to dislocation pile-ups and boundaries that dislocation slip bands have nucleated from.
Both types of shear accommodation have been found to follow previously developed slip
transmission criteria. The initiation of either type of grain boundary accommodation is ruled by
a combination of the ease of slip nucleation, the resolved shear stress on the slip system
(indicated by the global Schmid factor), and the geometric compatibility of slip systems in the
neighboring grains. The ECCI based approach outlined here allows linking a dislocation level of
understanding of the deformation mechanisms to the heterogeneous deformation over a
broader length scale than offered by traditional electron microscopy approaches.

6. Acknowledgements
The research is founded by Department of Energy, Office of Science, Basic Energy Sciences
through grant number DE-FG02-09ER46637. The authors would like to thank Dr. Christopher
Cowen, formerly at National Energy Technology Laboratory, Albany OR, for supplying the
titanium studied in this work.
7. References
[1] von Mises, R., 1913, “Mechanik der festen Körper im plastisch deformablen Zustand”, Göttin.
Nachr. Math. Phys. 1. 582-592.
http://www.digizeitschriften.de/dms/resolveppn/?PID=GDZPPN002503697

[2] J. D. Livingston, B. Chalmers, 1957, “Multiple slip in bicrystal deformation”, Acta Metallurgica
5. 6. 322-327. https://doi.org/10.1016/0001-6160(57)90044-5

[3] Z. Shen, R. H. Wagoner, W. A. T. Clark, 1986, "Dislocation pile-up at grain boundary


interactions in 304 stainless steel", Scripta Metallurgica 20. 921 – 926.
https://doi.org/10.1016/0036-9748(86)90467-9.

[4] T. C. Lee, I. M. Robertson, H. K. Birnbaum, 1989, “Prediction of slip transfer mechanisms


across grain boundaries”, Scripta Metallurgica 23. 5. 799-803. https://doi.org/10.1016/0036-
9748(89)90534-6

[5] T. C. Lee, I. M. Robertson, H. K. Birnbaum, 1990, “TEM in situ deformation study of the
interaction of lattice dislocations with grain boundaries in metals”, Philosophical Magazine 62. 1.
131–153. https://doi.org/10.1080/01418619008244340

[6] J. Luster, J. M. Morris, 1995, “Compatibility of deformation in two-phase Ti–Al alloys:


dependence on microstructure and orientation relationships”, Metallurgical and Materials
Transactions A 26. 7. 1745–1756. https://doi.org/10.1007/BF02670762

[7] J. Gong, A. J. Wilkinson, 2011, “A microcantilever investigation of size effect, solid-solution


strengthening and second-phase strengthening for <a> prism slip in alpha-Ti”, Acta Materialia 59.
5970-5981. https://doi.org/10.1016/j.actamat.2011.06.005

[8] T. R. Bieler, P. Eisenlohr, C. Zhang, H. J. Phukan, M. A. Crimp, 2014, “Grain boundaries and
interfaces in slip transfer”, Current Opinion in Solid State and Materials Science 18. 212-226.
https://doi.org/10.1016/j.cossms.2014.05.003

[9] Th. Kehagias, Ph. Komninou, G. P. Dimitrakopulos, J. G. Antonopoulos, Th. Karakostas, 1995,
“Slip transfer across low-angle grain boundaries of deformed titanium”, Scripta Metallurgica 33.
12. 1883-1888. https://doi.org/10.1016/0956-716X(95)00351-U.

[10] T. C. Lee, I. M. Robertson, H. K. Birnbaum, 1992, “Interaction of dislocations with grain


boundaries in Ni3Al”, Acta Materialia 40. 10. 2569-2579. https://doi.org/10.1016/0956-
7151(92)90326-A.

[11] L. E. Murr, 2016, “Dislocation Ledge Sources: Dispelling the Myth of Frank–Read Source
Importance”, Metallurgical and Materials Transactions A. 47, 5811–5826.
https://doi.org/10.1007/s11661-015-3286-5
[12] V. I. Nikolaichick, I. I. Khodos, 1989, “Review of the determination of dislocation parameters
using strong- and weak-beam electron microscopy”, J. Microsc. 155. 2. 123–167.
http://dx.doi.org/10.1111/j.1365-2818.1989.tb02879.x

[13] J. C. H. Spence, H. R. Kolar, G. Hembree, C. J. Humphreys, J. Barnard, R. Datta, C. Koch, F. M.


Ross, J. F. Justo, 2006, “Imaging dislocation cores – the way forward”, Philosophical Magazine 86.
29-31. 4781-4796. https://doi.org/10.1080/14786430600776322

[14] Y. Guo, T. B. Britton, A. J. Wilkinson, 2014, “Slip band–grain boundary interactions in


commercial-purity titanium”, Acta Materialia 76. 1–12.
https://doi.org/10.1016/j.actamat.2014.05.015

[15] B. Larrouy, P. Villechaise, J. Cormier, O. Berteaux, 2015, “Grain boundary–slip bands


interactions: Impact on the fatigue crack initiation in a polycrystalline forged Ni-based
superalloy”, Acta Materialia 99. 325–336. https://doi.org/10.1016/j.actamat.2015.08.009.

[16] B. E. Dunlap, T. Ruggles, D. T. Fullwood, B. Jackson, M. A. Crimp, 2018, “Comparison of


dislocation characterization by electron channeling contrast imaging and cross-correlation
electron backscattered diffraction”, Ultramicroscopy 184. 125-133.
https://doi.org/10.1016/j.ultramic.2017.08.017.

[17] P. Morin, M. Pitaval, D. Besnard, G. Fontaine, 1979, “Electron–channelling imaging in


scanning electron microscopy”, Philosophical Magazine A 40. 4. 511–524.
https://doi.org/10.1080/01418617908234856

[18] B. A. Simkin, M. A. Crimp, 1999, “An experimentally convenient configuration for electron
channeling contrast imaging”, Ultramicroscopy 77. 1–2. 65–75. https://doi.org/10.1016/S0304-
3991(99)00009-1

[19] M. A. Crimp, 2006, “Scanning electron microscopy imaging of dislocations in bulk materials,
using electron channeling contrast”, Microsc. Res. Tech. 69. 374–381.
http://dx.doi.org/10.1002/jemt.20293

[20] H. Mansour, J. Guyon, M. A. Crimp, N. Gey, B. Beausir, N. Maloufi, 2014, “Accurate electron
channeling contrast analysis of dislocations in fine grained bulk materials”, Scripta Materialia 84–
85. 11–14. https://doi.org/10.1016/j.scriptamat.2014.03.001

[21] S. Zaefferer, N.-N. Elhami, 2014, “Theory and application of electron channeling contrast
imaging under controlled diffraction conditions”, Acta Materialia 75. 20–50.
https://doi.org/10.1016/j.actamat.2014.04.018

[22] H. Kriaa, A. Guitton, N. Maloufi, 2017, “Fundamental and experimental aspects of diffraction
for characterizing dislocations by electron channeling contrast imaging in scanning electron
microscope”, Scientific Reports volume 7, Article number: 9742.
https://doi.org/10.1038/s41598-017-09756-3
[23] W. J. Tunstall, P. B. Hirsch, J. Steeds, 1964, “Effects of surface stress relaxation on the
electron microscope images of dislocations normal to thin metal foils”, Philosophical Magazine 9.
97. 99-119. https://doi.org/10.1080/14786436408217476

[24] J. P. Spencer, C. J. Humphreys, P. B. Hirsch, 1972, “A dynamical theory for the contrast of
perfect and imperfect crystals in the scanning electron microscope using backscattered
electrons”, Philosophical Magazine 26. 1. 193-213.
https://doi.org/10.1080/14786437208221029

[25] G. N. Kumar, B. Hourahine, P. R. Edwards, A. P. Day, A. Winkelmann, A. J. Wilkinson, P. J.


Parbrook, G. England, C. T. Cowan, 2012, “Rapid nondestructive analysis of threading
dislocations in wurtzite materials using the scanning electron microscope”, Phys. Rev. Lett. 108,
135503. https://link.aps.org/doi/10.1103/PhysRevLett.108.135503

[26] Y. N. Picard, M. E. Twigg, J. D. Caldwell, C. R. Eddy Jr., M. A. Mastro, R. T. Holm, 2009,


“Resolving the Burgers vector for individual GaN dislocations by electron channeling contrast
imaging”, Scripta Materialia 61. 8. 773-776. https://doi.org/10.1016/j.scriptamat.2009.06.021

[27] M. E. Twigg, Y. N. Picard, 2009, “Simulation and analysis of electron channeling contrast
images of threading screw dislocations in 4H-SiC”, Journal of Applied Physics, 105. 093520.
https://doi.org/10.1063/1.3110086; 2008, “Diffraction contrast and Bragg reflection
determination in forescattered electron channeling contrast images of threading screw
dislocations in 4H-SiC”, Journal of Applied Physics 104, 124906.
https://doi.org/10.1063/1.3042224

[28] S. Zaefferer, 2003, “A study of active deformation systems in titanium alloys: dependence
on alloy composition and correlation with deformation texture”, Mater. Sci. Eng. A 344. 20-30.
https://doi.org/10.1016/S0921-5093(02)00421-5
[29] J. P. Hirth, 1972, “The influence of grain boundaries on mechanical properties”,
Metallurgical Transactions 3. 3047-3067. https://doi.org/10.1007/BF02661312

[30] C. Blochwitz, J. Brechbühl, W. Tirschler, 1996, “Analysis of activated slip systems in fatigue
nickel polycrystals using the EBSD-technique in the scanning electron microscope”, Mater. Sci.
and Eng. A 210. 1-2. 42-47. https://doi.org/10.1016/0921-5093(95)10076-8

[31] F. Bridier, P. Villechaise, J. Mendez, 2005, “Analysis of the different slip systems activated by
tension in a α/β titanium alloy in relation with local crystallographic orientation”, Acta Materialia
53. 555–567. https://doi.org/10.1016/j.actamat.2004.09.040

[32] J. R. Seal, M. A. Crimp, T. R. Bieler, C. J. Boehlert, 2012, “Analysis of slip transfer and
deformation behavior across the / interface in Ti–5Al–2.5Sn (wt.%) with an equiaxed
microstructure”, Mater. Sci. and Eng. A 552. 61–68. https://doi.org/10.1016/j.msea.2012.04.114
[33] J. J. Fundenberger, M. J. Philippe, F. Wagner, C. Esling, 1997, “Modelling and prediction of
mechanical properties for materials with hexagonal symmetry (zinc, titanium and zirconium
alloys)”, Acta Materialia 45. 10. 4041-4055. https://doi.org/10.1016/S1359-6454(97)00099-2

[34] P. Acar, A. Ramazani, V. Sundararaghavan, 2017, “Crystal plasticity modeling and


experimental validation with an orientation distribution function for Ti-7Al Alloy”, Metals 7. 459-
471. http://doi.org/10.3390/met7110459

[35] J. C. Williams, R. G. Baggerly, N. E. Paton, 2002, “Deformation behavior of HCP Ti-Al alloy
single crystals”, Metallurgical and Materials Transactions A 33. 3. 837-850.
https://doi.org/10.1007/s11661-002-0153-y

[36] J. Gong, A. J. Wilkinson, 2009, “Anisotropy in the plastic flow properties of single-crystal α
titanium determined from micro-cantilever beams”, Acta Materialia 57. 19. 5693-5705.
https://doi.org/10.1016/j.actamat.2009.07.064.

[37] H. Li, D. E. Mason, T. R. Bieler, C. J. Boehlert, M. A. Crimp, 2013, “Methodology for


estimating the critical resolved shear stress ratios of α-phase Ti using EBSD-based trace analysis”,
Acta Materialia 61. 7555–7567. https://doi.org/10.1016/j.actamat.2013.08.042

[38] N. Benmhenni, S. Bouvier, R. Brenner, T. Chauveau, B. Bacroix, 2013, “Micromechanical


modelling of monotonic loading of CP a-Ti: Correlation between macroscopic and microscopic
behavior”, Mater. Sci. Eng. A 573. 222–233. https://doi.org/10.1016/j.msea.2013.02.022

[39] S. Naka, A. Lasalmonie, P. Costa, L. P. Kubin, 1988, “The low-temperature plastic


deformation of α-titanium and the core structure of a-type screw dislocations”, Philosophical
Magazine A 57. 717-740. https://doi.org/10.1080/01418618808209916

[40] J. Kacher, I. M. Robertson, 2016, “In situ TEM characterization of dislocation interactions in
α-titanium”, Philosophical Magazine 96. 14. 1437-1447.
https://doi.org/10.1080/14786435.2016.1170222

[41] E. Clouet, D. Caillard, N. Chaari, F. Onimus, D. Rodney, 2015, “Dislocation locking versus easy
glide in titanium and zirconium”, Nature Materials 14. 931-936.
http://dx.doi.org/10.1038/nmat4340

[42] H. Numakura, Y. Minonishi, M. Koiwa, 1986, “<1̅1̅23> {101̅1} slip in titanium polycrystals at
room temperature”, Scripta Metallurgica 20. 1581-1586. https://doi.org/10.1016/0036-
9748(86)90399-6
[43] M. H. Yoo, J. R. Morris, K. M. Ho, S. R. Agnew, 2002, “Nonbasal deformation modes of HCP
metals and alloys: Role of dislocation source and mobility”, Metallurgical and Materials
Transactions A 33A. 13. 813-822. https://doi.org/10.1007/s11661-002-1013-5
Figure 1. The geometry of slip transfer across grain boundaries and
associated slip transmission criteria [2-6].
Figure 2. a) Backscattered electron image of grains 1, 2, 3 and 4 before electropolishing. Slip bands in each grain are clearly visible, however, the topography is
too large for ECCI analysis. The white arrow indicates a less distinct of slip band near the grain boundary; the red arrow indicates the only slip band in grain 2
that intersects the boundary with grain 3 at a common point on the boundary. b) Backscattered electron image of the same area after electropolishing. Only
faint contrast is remains from the slip bands.
Figure 3. a) EBSD map of grains 1-3 after four-point bending straining. Hexagonal cell in each grain indicate the orientations. b) EBSD map of the same area
after electropolishing. The step size in (b) was courser than in (a). The slight differences in the orientations between the two maps can be attributed to slight
changes in the mounting during the two EBSD sessions
Figure 4. Backscattered electron images collected prior to
electropolishing showing the details of the slip bands in grains 1-3.
The traces for potential slip planes are indicated on each image with
the dashed lines, the colors of which correspond to the slip systems
indicated to the right. Because some slip planes have similar traces,
it is not possible to fully attribute the slip lines to specific systems
solely on the slip trace analysis.
Figure 5. a) Electron channeling contrast image of a distinct slip band. b-f) Higher magnification ECC images collected
from the red square region in (a) showing the details of the dislocation structure in the slip band collected using different
channeling conditions.
Figure 8. a) ECC image of a dominant slip band in grain 3. b-d) Higher
magnification ECC images of the slip band, the positions indicated by the
respective letters b, c, and d in (a), showing how the band broadens across the
grain, indicating the band propagated from the upper left of the grain towards
the lower right.
Figure 9. The evolution of the slip band morphology across grain 1 is shown in
the low magnification ECC image in (a) and the corresponding higher
magnification images in (b-d). The slip band nucleated from the high strain
region indicated by the strong, bright contrast in (b), and spread out
dramatically as it propagated in to grain 1, as shown in (c and d).
Figure 10. a) ECC image of grain 3 shows that four predominant slip bands
from the upper left to lower right. A less distinct slip band shown in (b-d)
spreads from the lower right to upper left, suggesting this band propagated in
the opposite direction of the dominant slip bands in this grain.
Figure 11. a) ECCimages of the dominant slip bands in grain 2 propagating from
the upper left to the lower right. b-d) show the nucleation of an additional
accommodating slip band in grain 2 from the boundary with grain 3, which
rapidly spreads as it propagates 6-7µm towards the upper left in grain 2.

You might also like