You are on page 1of 16

Materials Characterization 173 (2021) 110941

Contents lists available at ScienceDirect

Materials Characterization
journal homepage: www.elsevier.com/locate/matchar

Comparison of EBSD, DIC, AFM, and ECCI for active slip system
identification in deformed Ti-7Al
Ryan Sperry a, Songyang Han b, Zhe Chen c, Samantha H. Daly c, Martin A. Crimp b,
David T. Fullwood a, *
a
Brigham Young University, Provo, UT, USA
b
Michigan State University, East Lansing, MI, USA
c
University of California, Santa Barbara, California, USA

A R T I C L E I N F O A B S T R A C T

Keywords: The use of SEM-DIC, AFM, ECCI, and HR-EBSD to characterize slip-system activity was assessed on the same
EBSD material volume of Ti–7Al. This study presents a robust comparison of the various methods for the first time,
AFM including an assessment of their advantages and disadvantages, and how they can be used effectively in a
DIC
complementary fashion. The analysis of the different approaches was carried out in a blind, round-robin manner
ECCI
at three different universities. A Ti–7Al specimen was deformed in uniaxial tension to approximately 3% axial
Dislocations
Slip bands strain, and the active slip systems were independently identified using (i) trace analysis; (ii) in-SEM digital image
correlation, (iii) observations of residual dislocations from ECCI, and (iv) long-range rotation gradients through
HR-EBSD, with consistent trace identification in all cases. Displacement data from AFM was used to augment
SEM-DIC displacement data by providing complementary out-of-plane displacement information. Furthermore,
short-range dislocation gradients (measured by DIC) provided insight into the residual geometrically necessary
dislocation (GND) content, and was consistent with the GND content extracted from EBSD data and ECCI images,
confirming the presence of residual GNDs on the dominant slip systems resulting in visible slip bands. These
approaches can be used in tandem to provide multi-modal information on slip band identification, strain and
orientation gradients, out-of-plane displacements, and the presence of GNDs and SSDs, all of which can be used
to inform and validate the development of dislocation-based crystal plasticity and strain gradient models.

1. Introduction Slip-trace analysis has been used since the advent of automated
electron backscatter diffraction (EBSD). An observed line of slip on the
Understanding the deformation of structural alloys relies heavily on sample surface can be compared with potential slip-plane traces asso­
a knowledge of dislocation slip mechanisms. The contributions of ciated with the measured crystal orientation, resulting in the identifi­
dislocation activity to deformation can be studied by direct observations cation of the most-likely active slip plane(s). In-SEM digital image
of dislocation evolution, and/or by monitoring plastic strain that results correlation (SEM-DIC) performed with EBSD can substantially enhance
from this evolution. Many techniques exist for the characterization of slip-trace analysis. For example, the relative displacement ratio (RDR) of
dislocations and related strain. Techniques that can simultaneously Chen and Daly provides high-accuracy analysis of the active slip systems
provide complementary information concerning the deformation pro­ and the resolved slip on those systems [1]. Similarly, surface topography
cesses are particularly insightful, potentially delivering multiple view­ measurements using an atomic force microscope (AFM) can provide out-
points of the same sequence of events, and thereby uncovering new of-plane displacements to complement in-plane measurements and
levels of understanding. This paper analyzes deformation behavior in a determine the resolved shear on specific slip systems [2].
Ti–7Al alloy by simultaneously applying several traditional and Characterization of dislocation content that evolves during the
recently developed surface-based meso-scale characterization tech­ deformation provides another critical view of slip-system activity.
niques. The information regarding specific slip-system activity that is Dislocation structures have historically been characterized by tech­
extracted by each technique is contrasted and compared. niques including etch pit observations, x-ray tomography, and

* Corresponding Author
E-mail address: dfullwood@byu.edu (D.T. Fullwood).

https://doi.org/10.1016/j.matchar.2021.110941
Received 24 July 2020; Received in revised form 28 January 2021; Accepted 29 January 2021
Available online 2 February 2021
1044-5803/© 2021 Elsevier Inc. All rights reserved.
R. Sperry et al. Materials Characterization 173 (2021) 110941

transmission electron microscopy (TEM) (for a review see Hull and study is to assess the relative abilities of SEM-DIC, AFM, ECCI, and HR-
Bacon [3]). At the relevant meso-scale, electron channeling contrast EBSD to characterize slip-system activity from the same material volume
imaging (ECCI) has emerged as a powerful tool for the direct observation of a Ti–7Al sample. The analysis of the different approaches was carried
of individual dislocations in the near surface (~100 nm) regions of bulk out in a blind, round-robin approach at three different universities. This
polycrystals. Using the equivalent of traditional TEM diffraction contrast study presents a robust comparison of the various techniques, including
analysis approaches, ECCI can be used to identify the Burgers vectors of an assessment of the advantages and disadvantages of each of the
dislocations [4–6], and when combined with plane trace analysis, can be techniques and how they can be used effectively in a complementary
used to identify the active slip planes of the dislocations [7–9]. However, manner.
ECCI is generally limited to relatively low dislocation densities produced
at low plastic strains [10,11]. 2. Experimental procedures
At larger deformations, high-resolution EBSD (HR-EBSD) provides a
tool to determine geometrically necessary dislocation (GND) density Ti–7Al was used in this study as a model material because the
based on the local crystal elastic strains (predominantly the rotation hexagonal structure facilitates identification of particular dislocation
component) that result from the accumulated dislocations [12–15]. Burgers vectors based simply on slip band trace analysis. The Ti–7Al
These GND measurements are based upon short-range gradients; i.e. alloy was forged and annealed at 945 ◦ C for 1 h, followed by water
deformation gradients between neighboring scan positions - usually of quenching. Dogbone-shaped samples with gauge dimensions 2 wide ×
the order of 1 m. In deformed structures longer-range (usually meaning 0.8 thick × 10 mm long were wire electro-discharge machined with the
of the order of tens of ms in this paper) rotation gradients often exist. tensile direction along the forging direction. The sample was polished
When long range strain gradients exist in the material, for example along following a sequence of 600, 800, and 1200 grit SiC paper, 3 μm and 1
slip bands, it has been shown that the active slip systems can be iden­ μm diamond paste, and a mixture of colloidal silica and 30% hydrogen
tified from the GND-induced rotation gradients [16]. While HR-EBSD is peroxide with a 4:1 volume ratio. The mechanically finished sample was
generally more suitable for studying higher dislocation densities than then electropolished under 38 V applied voltage, − 35 ◦ C, and with
ECCI [17], recently the dislocation structures developed around nano­ stirring speed of 200 rpm in an electrolyte that contained 30 ml
indentations have been compared using both techniques, and the results perchloric acid, 200 ml butanol, and 300 ml methanol. The electro­
shown to be comparable in terms of Burgers vector and dislocation polishing time was around 2.5 min with the electropolishing rate of
density [17]. approximately 2 μm/min [7]. Before deformation, a rough EBSD scan
Each of these techniques provides invaluable and complementary was taken on the center region of the electropolished sample (Fig. 1)
information regarding the deformation processes, and each can be used using 30 kV accelerating voltage, a 148 μA probe emission current, a
to infer the locally active slip systems. Furthermore, the techniques can 20.0 nm spot size, a 20 mm working distance, and a step size of 5 μm
be performed ‘simultaneously,’ in the sense that a sequence can be with the sample tilted to 70o, using a Tescan Mira III FEG-SEM equipped
designed to gather each information type at a given strain step of a with an EDAX-TSL orientation imaging system. EBSD angular accuracy
deformation process. However, to date there have been no studies car­ is better than 0.5 degrees; for HREBSD disorientation is assumed to be
rying out a comprehensive evaluation of the information gained from accurate to around 1e-4. Note that Neighbor CI Correlation was used to
each method, in relation to the other approaches. The objective of this clean up data points of CI < 0.1. The sample was slightly textured close

Fig. 1. (b) EBSD crystal orientation map (IPF) from the region marked in red on the Ti–7Al sample (a), before deformation; the confidence index (CI) ranged from
0.75–0.88. (c) The sample is slightly [0001] textured, as indicated by the pole figure and inverse pole figure. The dashed region of the IPF map (b) was patterned for
DIC. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

2
R. Sperry et al. Materials Characterization 173 (2021) 110941

to [0001], but the texture did not play a significant role in the results of visible in SE mode after the removal of the Au NP coating. The crystal
this study. orientation information of the patterned area was again collected by
To facilitate plastic deformation, in-situ tensile testing was per­ EBSD using a step size of 1 μm, with all other parameters the same as in
formed using a Kammrath & Weiss tensile/compression module inside a the previous scans, as shown in Fig. 5. High resolution EBSD scans of
FEI Teneo SEM. Displacement-controlled loading was applied at a rate of areas of interest were also taken using the same parameters, with a step
1 μm/s, to a maximum uniaxial strain of 0.034, and then unloaded to size of 200 nm; this is above the threshold observed in [17] for
zero load. The loading was paused at six strain values, indicated on the convergence of the GND density measurements to a stable plateau, while
stress-strain curve and in the table of Fig. 2, to capture SEM images of the also providing several EBSD scan points between typical slip bands. Each
sample. An automated system [18] was used to capture a total of 8 × 8 high-resolution pattern was saved for subsequent cross-correlation and
= 64 tiles of images to interrogate a sample area of 400 μm × 400 μm. GND analysis; patterns were collected using an exposure time of 0.1 s
Each image had a field of view of 60 μm × 60 μm and a 4096 × 4096 with 480 × 480-pixel resolution.
pixel resolution resulting in a resolution of 14.6 nm. ECCI was carried out using the same Tescan Mira III SEM. Based on
In preparation for DIC data collection, the sample was patterned the EBSD determined orientations of the grains of interest, appropriate
prior to testing with gold nanoparticles (AuNPs) following the method stage tilts and rotations to orient the crystal near specific electron
outlined in [19] and briefly described below: AuNPs with diameters of channeling conditions were determined using the Tools for Orientation
~60 nm were fabricated using the method introduced by Frens [20]. The Determination and Crystallographic Analysis TOCA Software [22].
sample was first soaked in a solution of 2.5 vol% (3-Aminopropyl) tri­ Specific imaging conditions (g) were set up using selected area chan­
methoxysilane (APTMS), 47.5 vol% ethanol, and 50 vol% of DI water for neling pattern (SACP) mode available on this microscope.
20 min, followed by a quick rinse with DI water. This functionalized the Finally, the surface topography developed during plastic deforma­
sample surface with the aminosilane molecules by creating Si-O-Metal tion was measured by atomic force microscopy (AFM) using a VEECO
covalent bonds. The sample was then soaked in the AuNPs suspension Dimension 3100 AFM operating in tapping mode at a speed of 10 μm/
for ~18 h, followed by a quick rinse with DI water, after which the min for every 40 × 40 μm2 area with a vertical resolution of 0.1 nm. The
AuNPs were attached to the sample surface to create random speckle data was processed by the Gwyddion software package [23]. The slip
patterns used for DIC data collection. Digital image correlation (DIC) systems that developed during plastic deformation were then identified
was performed on each image tile using the commercial software Vic2D and analyzed through various techniques using DIC, EBSD, and ECCI.
6 (Correlated Solutions, Inc., Irmo, SC) to calculate the displacement and Data collected can be found as they were published publicly [24].
Lagrangian strain fields, using a subset size of 25 × 25 pixels and a step It is important to note that the sample was marked to ensure each
size of 2 pixels. Fig. 3 presents the axial strain map of the sample at a scan was taken with similar orientations and then programmatically
globally-applied axial strain of 0.028. oriented to match based on grain boundaries and physical attributes of
Following mechanical testing and imaging for SEM-DIC, the coating the sample, thus correcting for any orientation differences as the sample
was removed in order to extract HR-EBSD, AFM, and ECCI data from the was transported between and scanned in different machines.
deformed sample for the other deformation assessment techniques. To
achieve this, a chemical method was developed in the current study to 3. Active slip system identification methods
selectively break the Si–O bonds and remove the AuNPs from the
sample surface. The Si-O-Ti chemical bonds between the surface and the 3.1. Traditional slip trace analysis
polymer chain were fully broken after four hours at 30 ◦ C in a chemical
solution containing tetra-n-butylammonium fluoride (TBAF) [21], EBSD-based slip system analysis was applied to slip traces identified
chloroform, and ethylene glycol with 10: 1: 1 ratio by weight. During the in SEM SE images, DIC strain maps, and AFM maps. Potential slip traces
4-h process, the sample was removed from the solution every hour and associated with slip planes for a particular grain orientation were
cleaned with soap water under sonication (20–40 kHz) for 5–10 min to compared with the observed trace lines using an in-house code. Angles
remove the particles that were physically attached to the metal surface. between potential slip plane traces and observed slip lines (for the most
Final cleaning was accomplished by soap-water washing, ethanol-water- prominent slip trace direction in the grain) are tabulated in Table 1 for
ethanol flashing, and air drying. The success of the uncoating progress grain #1 from Fig. 5a. A common weakness with this approach is that
was verified using SEM in secondary electron mode (SE), backscattered several slip planes may approximately align with the actual observed
electron mode (BSE), and electron channeling mode using the same trace. In these cases, the active slip system can often be inferred based
MIRA III, as shown in Fig. 4. The microscope parameters were similar to upon the likelihood of activation from the magnitude of the Schmid
those used for EBSD, but the working distance was around 8–9 mm and factor with respect to the global tensile direction (included in Table 1 for
the sample was not tilted. the example grain), combined with an analysis of critical resolved shear
The slip traces that developed during deformation were clearly stresses (CRSS) for the various competing slip systems [25–30]. Due to
the hexagonal crystal structure of titanium, slip in the <a > directions of
the basal and prismatic planes has significantly lower CRSS than more
complex <a > and < c + a > slip on higher order pyramidal planes [30],
providing confidence in the slip system identification by trace analysis
for this particular grain, as will be discussed below. It is important to
note that even if a given slip system appears to be weakly active in SEM
images, this can be the result of the Burgers vector lying close to the
parallel with the plane of the sample, while other slip lines might be
more evident solely due to their Burgers vectors being more perpen­
dicular to the sample surface. This limitation of slip trace analysis can be
supplemented through the use of other methods, which will be discussed
further below.
Based upon this analysis, slip system #4 in Table 1 is the most likely
Fig. 2. Engineering stress-strain plot for the Ti–7Al sample. The red dots active slip system associated with the dominant slip trace, as it exhibits a
indicate the stress and strain values at which loading was paused for SEM im­ 1o error between the observed and theoretical slip trace direction, a
aging. (For interpretation of the references to colour in this figure legend, the Schmid factor of 0.49, and a low prismatic CRSS. Fig. 6(a) shows the
reader is referred to the web version of this article.) axial strain (εxx) map of the upper right region of grain 1 at a global

3
R. Sperry et al. Materials Characterization 173 (2021) 110941

Fig. 3. Axial strain map of the sample at a global strain of 0.028.

strain of εG = 0.028, with a second family of slip traces observable. Slip of the correct slip system identification [1]. To reduce error caused by
systems #4 ((0110) [2110]) and #5 ((1010) [1210]) were identified to experimental noise, multiple pairs of points along a segment of a slip
be associated with the two families of slip traces in the selected grain. trace line can be used for analysis, where the experimental RDR from all
This is illustrated in the figure by the blue solid lines, which are the pairs of points are averaged. In this work, the experimental RDR values
theoretical trace lines for these systems and match the observed slip were measured for four randomly chosen line segments as indicated by
traces. The intense traces that are oriented from the upper left to the the red solid lines in Fig. 6(a). Fig. 6(c) shows the evolution of the
bottom right correspond to the (0110) [2110] slip system #4. The weak measured RDR for these four lines at different global strains. The theo­
traces lying from the upper right to the bottom left correspond to the retical RDR values for the two slip systems are indicated by the black
(1010) [1210] slip system (#5, with a smaller Schmid factor of 0.23). dashed lines in the inset. At low global strains, the measured RDR values
had a large variance, because the local deformation was very small, and
the measured RDRs were mainly from experimental noise. All of the
3.2. Strain-based methods: DIC, RDR, and AFM labeled slip traces had activated by a globally applied strain of 0.018,
and their experimental and theoretical RDR values were similar. For
Based on SEM-DIC data, the relative displacement ratio (RDR) example, the experimental RDRs of line segments 1 and 2 between
method uses observations of full-field deformation to infer the active slip global strains of 0.018–0.034 were averaged, and the averaged experi­
system that would produce the observed relative displacements across a mental RDR of 1.3 was then compared with the theoretical RDRs of all
slip band [1]. Slip trace analysis can be used to determine the most likely possible slip systems. As summarized in Table 1, the smallest difference
subset of active systems, and RDR analysis can then be used as an between the averaged experimental and theoretical RDRs was found
additional source of information to identify the active system with when comparing with slip system #4, with a difference of 0.25.
higher accuracy, although the RDR analysis alone does not provide Although other slip systems with the same slip direction as slip system
unambiguous identification of active slip systems in the present mate­ #4, including slip systems #1, #7, and #10, also exhibited a difference
rial. Fig. 6(b) briefly summarizes the concept of the RDR. P1 and P2 are a of 0.25, they were not likely active because their theoretical slip trace
pair of points located on different sides of a slip trace, and the relative directions did not match the experimental observation.
displacement between them is: To further investigate the precision of the DIC slip trace analysis, an
AFM topography map was used to cross-validate the DIC method. AFM
S = (∆u, ∆v,∆w) = (u(P2 ) − u(P1 ) , v(P2 ) − v(P1 ) , w(P2 ) − w(P1 ) ) (1)
can capture fine-scale measurements of the out-of-plane displacements
along and across slip bands. The local shear distribution, as illustrated in
where u, v, and w represent the displacements in the x, y, and z di­
Figs. 7(c) and 8, was calculated from AFM and DIC data, respectively,
rections. Assuming the relative displacement results from the deforma­
and compared.
tion of slip system α with a Burgers vector bα = (bαx, bαy , bαz ), and a slip
The AFM map from an area in grain 1 is shown in Fig. 7(a), where the
plane normal vector nα = (nαx, nαy , nαz ), the following relationship can be
background was subtracted and the point with the lowest height was set
obtained:
to 0. The height difference H (nm) across a slip band within a certain
∆u ∆v ∆w distance X (μm) can then be precisely extracted from the AFM line
S= = α = α (2)
bαx by bz profile data (i.e. the blue line in Fig. 7(a), with blue arrows indicating
the slip bands and the corresponding heights). Based on the height
where S represents the magnitude of the relative displacement S and b change across the individual slip bands and the known Burgers vector
represents the length of the Burgers vector. Note that only in-plane direction, the shear across slip bands was calculated, as outlined in [2]
displacements u and v were measured by 2D DIC in the current study; and illustrated in Fig. 7(c). For a given slip system, the number of dis­
w can be measured with AFM. The second and third terms of Eq. (2) can locations that are responsible for the slip band height, N, can be calcu­
be rewritten as: lated from the height difference, H (which is related to the Δw component
/ / in eq. 2 in the DIC/RDR analysis), and the Burgers vector b projected onto
bx by = ∆u ∆v (3)
the surface normal (e):
The ratio bx/by is defined as the theoretical RDR, and the ratio ∆u/∆v H
is defined as the experimental RDR. A smaller difference between the N= (4)
b∙e
experimental and theoretical RDR values indicates a higher confidence

4
R. Sperry et al. Materials Characterization 173 (2021) 110941

Fig. 4. (a and b) Overall surface condition of the


sample coated with Au NPs. Note the electron chan­
neling patterns (inset of b) are blurry due to the
nanoparticle coating. (c) Surface condition after 1 h
of the uncoating procedure. (d) Higher magnification
of the red area in image (c). Note the selected area
channeling pattern is now much sharper, with slip
traces evident, but some clusters of AuNPs still
remain on the surface. (e) The AuNPs are completely
removed after 4 h, as shown in detail in (f), which
shows the red area in (e) at higher magnification.

The average shear γ contributed by each slip system can be estimated calculated Schmid factor (0.49 vs. 0.23) for the same CRSS.
as: Alternatively to the AFM approach, the local shear on active slip
/ ( )/ systems can be calculated using the full-field displacement data from the
H
γ = (bN) (Xn) = b (X∙n) (5) DIC analysis. The overall approach of this method is similar to the AFM
b∙e
characterization, but more precise since it involves multiple in-plane
Using this approach, the local shear can be mapped by determining displacements. For a given material point located at a pixel position P
the gamma on a tile by tile basis. For example, by manually dividing the (x, y, 0), the local deformation was measured by a virtual extensometer
AFM topography map in Fig. 7(b) into 29 × 29 squares tiles (X = 0.3 with a length of Lx = 20 pixels (~0.29 μm), placed along the x-direction
μm), the height difference H across each slip band in each tile can be of the sample, and centered at point P. The virtual extensometer can be
calculated from the AFM line profile, and the local shear can be mapped represented by a vector L = (Lx, 0, 0), and it can measure the relative
as shown in Fig. 7(c). In this case, the maximum shear is 0.70 and the displacement S between points P1(x − 10, y, 0) and P2(x + 10, y, 0). Note
minimum shear is 0.01. As can be seen from Fig. 7(c), the (0110) [2110] that the length of the virtual extensometer (~0.29 μm) was selected to
slip system (#4), which runs from the upper left to the lower right, ex­ approximately match the size of the AFM tile (0.3 × 0.3 μm), so that the
hibits more activity (γ ranging from 0.28–0.62) than the (1010) [1210] local shear was estimated on a similar length scale. Assuming the rela­
slip system (#5, with γ ranging from 0.09–0.16), which runs from the tive displacement results from the deformation of slip system α, the local
upper right to the lower left. This is consistent with the differences in the shear on slip system α, represented by γα, where γrepresents shear and
the superscript α represents the slip system number, is calculated using

5
R. Sperry et al. Materials Characterization 173 (2021) 110941

Fig. 5. (a) Slip traces developed during plastic deformation were clearly observed in SE mode. (b) EBSD inverse pole figure maps showing the orientations of the
various grains. In some regions, local dislocation accumulations result in local orientation variations. Grains 1 and 2 were the focus of the majority of this manuscript.

the following equation: resulting local shear distribution map shown in Fig. 8(f) consists of 33 by
33 windows, each of size 0.29 × 0.29 μm, providing approximately the
S
γα = (6) same resolution as Fig. 7(b). Compared to Fig. 7(b), the relative shear
L∙nα
along and across each slip band is consistent with the calculation result
which is essentially the same as Eq. (5). According to Eq. (2), the from AFM. Several comparisons across several random slip bands in
magnitude of the relative displacement, S, can be calculated with the different areas have been made between the DIC and AFM approaches,
Burgers vector and either ∆u or ∆v: and the results show that DIC is reliable for the shear measurement. This
cross-validates the accuracy of the slip system identification by the DIC
∆u approach. Importantly, this approach can provide more precise local
S=b (7a)
bαx shear information from individual slip systems as well as the axial strain
∆v
S=b α (7b) in this area, which is not accessible from the AFM method.
by

The small area indicated by the black box in Fig. 6(a), which is 3.3. Dislocation observation methods: ECCI
approximately the same area analyzed by AFM in Fig. 7(b), was selected
for analysis. Figs. 8(a) – 8(c) shows the resulting axial strain map, the u- ECCI was used to image individual dislocations in the near surface
displacement map, and v-displacement map, respectively. As two slip region of the deformed sample and to characterize the slip systems.
systems (#4 and #5) were identified to be active, the relative Fig. 9(a) shows a relatively low magnification secondary electron image
displacement was considered as a sum of the relative displacement of grain 1, in which dominant slip bands, running from the upper left to
associated with each of the active slip systems: the lower right, are evident (traced in green). Trace analysis indicates
S = S4 + S5 (8) that these slip planes are (0110) prism planes. Additional slip bands that
run from the upper right to the lower center are seen in the right side of
The components of the relative displacement associated with the two the grain (traced in blue) and are identified as (1010) prism planes. This
active slip systems (∆u4, ∆v4 for slip system #4 and ∆u5, ∆v5 for slip slip band contrast is primarily a result of the topography from the slip
system #5) were calculated by a system of equations: bands but may also be reinforced from the dislocations lying in the slip
⎧ bands [7]. Fig. 9(b) shows a higher magnification image of the defor­


⎪ ∆u4 + ∆u5 = ∆u (9a)

⎪ mation structure, outlined in the red box in 9 (a), that reveals the in­

⎨ ∆v4 + ∆v5 = ∆v (9b) dividual dislocations. It is evident in these higher magnification images

⎪ b4y ∆u4 − b4x ∆v4 = 0 (9c) that the dominant slip bands are made up of densely packed dislocations



⎪ b5y ∆u5 − b5x ∆v5 = 0 (9d) that bow out considerably, but have ends that align along the slip band

direction (in ECCI, dislocations will go out of contrast as they extend
further below the surface). The dislocations lying in the slip bands that
where the two equations, (9a) and (9b), were obtained from Eq. (8) by run from the upper right to the middle of the image appear to bow out on
considering individual components along each axis. Applying Eq. (3) to planes that are more parallel with the surface. In addition to the dislo­
each active slip system can provide two additional equations: (9c) for cations associated with the slip bands evident in Fig. 9(a), significant
slip system #4 and (9d) for slip system #5. This provides four inde­ numbers of dislocations are also observed on slip planes that have not
pendent equations for four unknowns. For more than two active slip developed strong surface slip traces. These dislocations have bowed out
systems, applying Eq. (3) can provide only provide only one additional on (0001) basal planes and exist in a significant number of dislocation
equation but two more unknowns (∆uα and ∆vα for slip system α), and pile-ups, identified by the red brace in Fig. 9(b).
this would result in an indeterminate problem, and an optimal solution The Burgers vectors of individual dislocations were identified using
would be used instead. ECCI g・ ・b = 0 and g・ ・b x u = 0 invisibility criteria [4], where g is the
Using the method described by Eqs. (6) and (7), the local shear on channeling vector, b is the Burgers vector, and u is the dislocation line
slip systems #4 and #5 were calculated for each pixel in the selected direction. Compared to other slip trace methods that have the potential
area, as shown in Figs. 8(d) and (e). The sum of the local shear on the for error based on misidentification of the slip system, which results in a
two slip systems was calculated, and then coarsened by an average filter “fit” factor (RDR, rotation angle) between the theoretical trace and the
with a window size of 20 × 20 pixels and a step size of 20 pixels. The observed trace, ECCI requires a judgement of whether a dislocation is

6
R. Sperry et al. Materials Characterization 173 (2021) 110941

Table 1
Designations of the slip systems considered in this study: the Schmid factor, the theoretical RDRs, and the angle between the axis of lattice rotation and theoretical line
direction for each system in grain #1 indicated by Fig. 5(a). (Note: the experimental RDR in this table was 1.30, which was calculated by averaging the experimental
RDR of two slip-trace line segments for the dominant family of traces, at global strains between 0.018 and 0.034.)
Slip system Slip system Slip system Schmid Angle between theoretical and Difference between experimental Angle between measured axis of
type number factor measured slip trace (deg) and theoretical RDR rotation and line direction (deg)

Basal <a> 1 (0001) [2 1 0.07 29 0.25 87


1 0]
2 (0 0 0 1) [1 2 0.02 29 1.56 88
1 0]
3 (0 0 0 1) [1 1 0.09 29 5.33 90
2 0]
Prismatic <a> 4 (0 1 1 0) [2 1 0.49 1 0.25 3
1 0]
5 (1 0 1 0) [1 2 0.23 59 1.56 3
1 0]
6 (1 1 0 0) [1 1 0.26 62 5.33 3
2 0]
Pyramidal 7 (0 1 1 1) [2 1 0.40 7 0.25 26
<a> 1 0]
8 (1 0 1 1) [1 2 0.19 69 1.56 27
1 0]
9 (1 1 0 1) [1 1 0.28 69 5.33 28
2 0]
10 (0 1 1 1) [2 1 0.47 4 0.25 31
1 0]
11 (1 0 1 1) [1 2 0.21 50 1.56 31
1 0]
12 (1 1 0 1) [1 1 0.19 57 5.33 29
2 0]
Pyramidal <c 13 (0 1 1 1) [1 1 0.33 7 2.91 77
+ a> 2 3]
14 (0 1 1 1) [1 2 0.12 7 1.59 75
1 3]
15 (1 0 1 1) [2 1 0.37 69 2.48 74
1 3]
16 (1 0 1 1) [1 1 0.47 69 2.91 78
2 3]
17 (1 1 0 1) [1 2 0.06 69 1.59 79
1 3]
18 (1 1 0 1) [2 1 0.09 69 0.81 74
1 3]
19 (0 1 1 1) [1 1 0.28 4 2.39 76
2 3]
20 (0 1 1 1) [1 2 0.03 4 1.48 78
1 3]
21 (1 0 1 1) [2 1 0.27 50 0.81 78
1 3]
22 (1 0 1 1) [1 1 0.38 50 2.39 75
2 3]
23 (1 1 0 1) [1 2 0.01 57 1.48 74
1 3]
24 (1 1 0 1) [2 1 0.09 57 2.48 79
1 3]

either out of contrast or in contrast for a given g. Errors are minimized planes contained [1210] dislocations (blue). These two slip systems
by making these judgements for both in and out of contrast for multiple were consistent with those identified by RDR/DIC. Additionally, the
ECCI images taken with different g. In the present study, all dislocations dislocations observed to be bowed out on basal planes (red) had [1120]
were assessed for four different g. Fig. 10, with accompanying table, Burgers vectors.
shows such an analysis for the region of grain 1 shown in Fig. 9. In this In contrast to other surface analysis techniques, ECCI depends on
study, the slip planes were identified using the trace analysis approach contrast variations due to lattice distortions rather than surface dis­
described above. In all cases, the identified Burgers vectors were placements. Thus, it is capable of identifying dislocations that are
consistent with these slip planes. In the area studied in Fig. 9, four responsible for wavy slip traces and those that do not contribute to the
different Burgers vectors were identified. The dominant (0110) prism formation of an apparent slip trace. One typical example is the identi­
slip bands (green) contained [2110] dislocations and the (1010) prism fication of slip systems in grain 2, as shown in Fig. 11. Only (1100)

7
R. Sperry et al. Materials Characterization 173 (2021) 110941

Fig. 6. (a) Axial strain map of a


selected area of interest including part
of Grain #1, at a global strain of εG =
0.028. The theoretical slip trace di­
rections for two slip systems were
indicated by the blue solid lines. The
experimental RDRs were measured for
the four slip trace lines segments indi­
cated by the red solid lines. (b) Illus­
tration of the relative displacement
across a slip trace. P1 and P2 are a pair
of points located on different sides of a
slip trace. The relative displacement
between P2 and P1, represented as (Δu,
Δv, Δw), is measured at P2 using P1 as
reference. (c) Evolution of the experi­
mental RDR measurements at different
global strains for the four slip trace line
segments. The theoretical RDRs for the
two slip systems are indicated by the
black dashed lines in the inset. (For
interpretation of the references to
colour in this figure legend, the reader
is referred to the web version of this
article.)

[1120] (red trace) and (0110) [2110] (blue trace) slip systems were 3.4. Dislocation-based observations: GND content from EBSD
identified by slip traces in Fig. 11(a). The yellow trace was not identified
as it did not match any theoretical traces. This unknown slip system was Relating to the observation of individual dislocations using ECCI,
identified as the (1011) [1210] slip system in the more detailed ECCI arrangements of geometrically necessary dislocations (GNDs) can also
analysis. Additionally, basal <a > slip systems that did not form an indicate the active slip system where strain gradients are present. This is
apparent slip trace were also identified, with the majority of them being particularly true where slip bands approach grain boundaries, resulting
(0001) [1210] slip. in dislocation pileups [11,26]. The GNDs cause orientation gradients
that can be identified using EBSD, from which the corresponding slip

8
R. Sperry et al. Materials Characterization 173 (2021) 110941

Fig. 7. (a) AFM-measured surface topography map in grain 1. The blue line traces the AFM profile showing the measured topography across the line, shown below
the AFM map. The small blue arrows indicate two prominent slip bands and their corresponding heights. The black boxed region denotes where the local strain was
determined by both AFM and DIC shown in Fig. 8. (b) AFM topography map of the black boxed area in (a). (c) Map of the local shear distribution of black boxed area,
consisting of 29 × 29 square tiles, with each tile having an area of 0.09 μm2. (For interpretation of the references to colour in this figure legend, the reader is referred
to the web version of this article.)

system can be inferred. The dislocation content in the slip bands typi­ systems on the identified slip plane(s), and deformation systems with
cally consists of a distribution of dislocations with the same Burgers line directions close to the axis of rotation were identified as possibly
vector that glide until encountering an obstacle, and create a pileup active.
[32,33]. Thus, the extra planes of atoms from each dislocation all on one Fig. 12 demonstrates two selected points that were used for this
side of the slip plane create long-range lattice strain and cause a rotation method. These points were found in grain 1. Using these two points, the
in the lattice about the line direction of the dislocations (under the lattice rotation was used to identify possible active slip systems
simplifying assumption of pure edge dislocations for this analysis). By following the method described above. Table 1 includes these results,
identifying the axis of rotation and comparing with possible line di­ showing the angle between the line direction and axis of rotation for
rections, the active slip system can be identified [16]. each slip system. A lower angle suggests a higher likelihood of the slip
For the slip bands in question (e.g. grain #1), a neighboring refer­ system being active. The three closest slip systems were the prismatic
ence point towards the center of the grain was selected, and the rotation slip systems, as they all have parallel line directions with an angle of
from this point to each evaluation point along the band was determined. only three degrees between measured and theoretical line directions.
A transformation matrix was determined from the EBSD data, indicating Since there are three possible systems identified, the slip trace analysis
the orientation of the lattice at each point. A third transformation matrix was then used to identify the active slip system responsible for the slip
from the reference point to the evaluation point was then created: traces (along with the Schmid factor, if necessary). It can be seen from
Table 1 that the (0 1 1 0) [2 1 1 0] system lies on the slip plane selected
Gr,e = G−r 1 Ge (10)
through slip trace analysis, and thus is identified as the active slip sys­
where Grand Geare the transformation matrices representing the orien­ tem; it also has the highest Schmid factor.
tation of the reference point and the evaluation point, respectively. Gr, e GND fields often produce only very local strain gradients, with the
is the rotation matrix from the reference point to the evaluation point. dislocation slip system changing periodically across a given grain; in
This transformation matrix Gr, e was then converted into an axis-angle which case, the two points selected in Fig. 12 might have simply been
rotation, giving the axis about which the lattice is rotating. This axis fortuitously connected to the correct active slip system. In order to
was then compared to the possible line directions of the deformation determine the robustness of this approach along a given slip band, we

9
R. Sperry et al. Materials Characterization 173 (2021) 110941

Fig. 8. (a) Axial strain, (b) u displacement, and (c) v displacement of a selected area indicated by the black box in Fig. 6(a). (d) The local shear on slip system #4. (e)
The local shear on slip system #5. (f) The sum of local shear coarsened by a window size of 20 × 20 pixels. Standard deviation for Ɛxx is 0.0069.

examine the consistency of slip system associated with GNDs along the map, and then summing the number in each point for all points of the
entire slip band; if GNDs are associated with a single system, then the band. The burgers vector magnitude of the slip system is denoted by b,
long-range rotation gradient along the slip band will match the local and h is the spacing of the slip bands. By multiplying both sides by Δx,
gradients used to calculate GND density. If the local GND density along the resulting equation can be used to estimate the rotation based on the
the slip band is associated with a single slip system, the dislocation number of dislocations along the band. This can then be used to estimate
content and the long-range lattice rotation gradient along the slip bands a rotation from the total number of GNDs, which can be taken from the
can be approximately related according to [31]: GND density maps, and compare it to the measured rotation.
This potential relationship between local GND content and long-
Δθ nb
= ρb = (11) range lattice rotation was compared for over 4000 different line seg­
Δx Δxh
ments, along parallel slip bands. The theoretical lattice rotations
where Δx is the length along the slip band and Δθ is the relative rotation calculated from the short-range strain gradient GND content (Eq. 11)
between two points. The number of dislocations along the band is rep­ was compared with measured lattice rotations using Eq. 10, and plotted
resented by n, which can be found by calculating the number of dislo­ in Fig. 14.
cations in each volumetric point of the slip band using the GND density The strong correlation indicates that the measured GND content is

10
R. Sperry et al. Materials Characterization 173 (2021) 110941

Fig. 9. (a) Secondary electron image showing two prism slip band orientations, (0110) (green) and (1010) (blue) in grain 1, determined by the trace analysis. (b)
Higher magnification ECCI image from the red region in (a) showing the local dislocation structure. Dislocations are identified in both the slip bands noted by the
green and blue braces, but additional dislocations are observed on the basal planes (red).

lying on a consistent slip system, presumably relating to dislocation The EBSD data can be used to generate a GND density map that as­
pileups at the ends of the slip bands. Hence, almost any pair of points sociates with the local strain gradient. This was done for the current
along a slip band can be used to extract the relevant rotation (Eq. 10) to sample using the open source OPENXY software [34] which is accurate
be used for slip system analysis. The observed agreement also confirms to 1011 m− 1. Following the guidance on parameter selection specified in
that the GND step size selection is appropriate for cancelling out mea­ [35], an effective step size of 990 nm was selected for the map, which
surement noise and SSDs. As shown in Fig. 13, the trendline between the demonstrated a stable level of average GND density in the scan relative
two methods has a slope of 1.06, with a p value of zero (or lower than the to step size. This step size is also sufficiently large to eliminate noise
computing roundoff error resulting in zero, which is characteristic of a from measuring techniques and nearby statistically significant disloca­
large dataset such as this one). The R-squared value of the data is 0.735, tions (SSDs), and thereby reveals the content of the GNDs on the
suggesting that while the correlation is strong, it is not perfect. Likewise, dominating slip systems responsible for the visible slip bands. Note that
the trendline does not pass through the origin. This could possibly be most slip bands are more than 1 m apart; however, some are closer than
due to multiple active slip systems contributing partially to the lattice this, resulting in GNDs from both bands being included in the GND
rotation in different directions. This would artificially increase the calculation for points near these bands (such as near the bottom left
rotation measured (y-axis) as there are two activated slip systems in this corner of the GND map in Fig. 13).
grain. Other causes can be attributed to noise in the EBSD, affecting the Since dislocations (including GNDs) are highly visible in the ECCI
accuracy of GND data generated from the short-range strain gradients. scans, they can be compared to the GND density maps generated from
Additionally, dislocation numbers calculated from the GND density the EBSD data. Fig. 14 shows a comparison between an ECCI scan in
maps assumed a constant line length, based upon the step size used; in grain 1 and the GND density map generated from the EBSD scan in grain
reality, dislocation line lengths within a given volume would vary (in­ 1. The locations of high GND density correlate reasonably well with the
crease or decrease) depending upon the actual position of the disloca­ locations where high GND density is visible in the ECCI images. For
tions relative to the scan points. Another factor could be errors in the example, the high-density locations in the GND density map show a
estimation of the band spacing, h. Both the ECCI data and the long-range maximum of about 1014.7. Assuming an evenly spaced array of dislo­
strain gradients from the EBSD scans indicate residual dislocations that cations, this would amount to a spacing of roughly 50 nm between
are associated with the dominant slip systems identified by the other dislocations, which is reasonably consistent with the ECCI image (simple
methods. counting of dislocations in the ECCI image across a number of high
Slip traces in the secondary direction (top right to bottom left as density regions indicates identifiable spacing of around 70 nm in both
indicated by the ECCI and DIC data) cannot be readily seen in the IPF dimensions; higher density may exist in the noisy regions). This in­
gradients in Fig. 12. Thus, slip trace analysis and point selection for this dicates that both the ECCI and GND data are identifying similar sets of
method may rely on SEM SE images, optical microscopy, or some other residual dislocations, and therefore the same localized short-range strain
means. Using this approach, the slip system was identified as the (1 0 1 gradients. This also indicates that the vast majority of dislocations that
0) [1 2 1 0] system, matching the results of the other methods used. appear in the ECCI image are in fact GNDs, since a large number of SSDs
This method was then implemented on various slip traces to probe its would result in a lower observed GND content.
more general viability. These results and their comparison with the The strain gradients extracted from the SEM-DIC data should also be
other two methods will be discussed later in this paper. consistent with those identified via the GND data. Long-range strain
gradients along a slip band, as measured by DIC, can be used to calculate
4. Quantitative comparison of dislocation activity from EBSD, the number of dislocations that remain in a pileup along that band [16].
ECCI and DIC The DIC displacement values directly relate to the number of disloca­
tions that have glided through a region of interest. The difference be­
The ability of the different methods to identify the active slip system tween displacement values measured for neighboring regions directly
(s) relies upon evidence of current or past dislocations extracted by the correlates with the number of dislocations that glided through one re­
individual methods. Hence, in this section we briefly compare the gion but not the other; i.e. the number that are left as residual GND along
observation of residual dislocations, and their related strain gradients, the slip band. By dividing the resulting difference in displacements by
by the three methods: EBSD, ECCI and DIC. the Burgers vector, the number of residual GND can be calculated.
Fig. 15 shows an example of a band and the corresponding strain curve

11
R. Sperry et al. Materials Characterization 173 (2021) 110941

Fig. 10. (a-d) ECCI images captured at different channeling conditions from the same region as in Fig. 9(b). The electron channeling conditions are indicated by the
black arrows on the upper right corners. The colored ovals and fonts indicate the Burgers vector of the various dislocations evident in the figures. Different dislocation
Burgers vectors are in strong or weak contrast depending on g, as outlined in the accompanying table. The majority of the dislocations are (0110) [2110] (green
traces in Fig. 9) and (1010) [1210] (blue), with some (0001) [1120] (red) dislocations.

12
R. Sperry et al. Materials Characterization 173 (2021) 110941

Fig. 11. (a) Slip traces developed during deforma­


tion in grain 2, colored lines represent the traces on
the surface. (b) ECCI image of the red area in (a),
where more dislocations types were identified, with
each type marked by the colored arrows. Note: black
dots along slip traces are likely caused by incomplete
removal of polymer or Au nanoparticles from the DIC
process. (For interpretation of the references to
colour in this figure legend, the reader is referred to
the web version of this article.)

Fig. 13. Plot of theoretical lattice rotation, calculated from short-range strain
maintained by GND content along bands (Eq. 11), and measured lattice rotation
as calculated from Eq. 10. There is a correlation between the two with an R-
squared value of 0.735. Dotted line shows the trendline between the two values.
Fig. 12. IPF map of grain boundary of grain 1 in the sample, with the green dot
and blue square representing the reference and evaluation points, respectively. summed in Table 2.
Interchanging these selected points does not affect the result, as it only causes As seen from the table, all methods identified the same active pris­
the rotation to be in the opposite direction about the same axis. (For inter­ matic slip systems in both grains, suggesting that all are viable methods
pretation of the references to colour in this figure legend, the reader is referred
that can identify the activated slip system. However, each method has
to the web version of this article.)
limitations and advantages.
The EBSD method is advantageous in that it can be relatively low
extracted from the band. The strain values at the beginning and end of cost, requiring only EBSD scans to be performed and not requiring the
the curve can be used to calculate the number of dislocations along that preparation or data collection of the DIC and ECCI techniques. However,
pileup in the band. this depends upon microscope setup and resources, and may not be
The number of GNDs along ten different bands were calculated using available to all. A limitation of the EBSD long-range rotation method is
this process and compared to the number of GNDs calculated from the that it can only be used to identify the activated slip system if slip band
GND density map based from short-range strain gradients. Fig. 16 shows traces can be visually identified, since points along the trace need to be
the comparison of calculated GNDs from the strain map versus the EBSD- selected. This can be difficult using strictly the EBSD data and IPF im­
obtained GND density map. There is a strong correlation between the ages, as the traces are only apparent from subtle colour gradients, which
two parameters along the red dashed trend line. The slope of this line is often only show evidence of the primary slip system or none at all.
1.062 with a p value of 6 e-6 suggesting that both sets of data are However, this problem can be addressed through SEM or optical mi­
accurately sharing information about the same dislocation pileups. croscopy, methods that typically can be used to visualize the slip traces.
Thus, further supporting the idea that in the presence of strain gradients, Additionally, this method is dependent on long-range rotation gradients
each method is measuring characteristics of the same active slip system in the crystal lattice that are created by GND along a slip band. If there
despite each method measuring distinct features of the deformed were no rotation / strain gradient associated with the bands, identifying
sample. the active slip system would be impossible using this method. We note
that HR-EBSD is much more sensitive to rotation gradients than regular
5. Slip system identification comparison EBSD, and could therefore identify gradients that would not be apparent
in regular EBSD. From a previous study in a Ni super-alloy [16], and the
All slip system identification methods were applied to two grains to current material, strain gradients generally occur along slip bands of
identify multiple active slip systems in each grain. These results are some polycrystalline materials – particularly within 1/3 of a grain

13
R. Sperry et al. Materials Characterization 173 (2021) 110941

Fig. 14. (a) ECCI image of grain 1 with dislocations visible. (b) GND map generated from short-range strain gradients showing localized distribution of GND. Maps
match up quite well showing similar dislocation content between areas in the red boxed areas in (a) and (b). Red line in (b) indicates cross-section of map used to
generate GND density profile seen below map. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of
this article.)

Fig. 15. Example of strain values and selected slip band that can be used to estimate GNDs. Section of axial strain map with selected portion of a slip band identified
with the red circle (a). Corresponding strain curve along the slip band (b). There is a negative gradient in the strain values along the band in the direction of the
dotted arrow, suggesting a dislocation pileup that can be measured. Strain values at beginning and end of the curve would be used to estimate the GND content of the
pileup. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

diameter from a GB. accurately measure local crystal orientations with respect to the electron
The RDR method is a reliable means to accurately identify active slip beam in order to set up specific electron channeling conditions typically
systems, even if strain gradients are not present. However, it is limited to requires specialized SEM modules (i.e. SACPs as mentioned previously),
identifying slip systems along slip bands that are visible in the DIC data and the level of user experience needed for this can be challenging since
and requires a certain level of strain for the RDR values to converge to determining channeling vectors and identifying dislocations need a
their final values. The RDR method also requires SEM-DIC data collec­ certain level of expertise. It can also be time consuming for large area
tion in addition to the EBSD data collection, which may increase costs dislocation analysis compared to the DIC and EBSD.
depending on experimental resources. This requires the experience and AFM is a strong supplemental technique capable of measuring out-of-
equipment necessary for surface polishing, speckle pattern deposition/ plane strain across an as-deformed sample surface. With AFM charac­
removal as well as SEM and EBSD detectors. terization of the Burgers vector, slip plane normal of the slip systems, it
ECCI is typically strong in identifying the nature of dislocations, is possible to calculate the relative strain contribution by each of the slip
including the Burgers vectors and slip planes, but will potentially also bands with relative comparable preciseness. However, AFM used by it­
identify dislocations that do not contribute to the apparent slip bands. self is not able to correctly identify the slip system and crystal orienta­
Hence, there is an assumption that residual dislocations in the material tion and must be used in conjunction with EBSD or other identification
are associated with the active slip system. The distribution of disloca­ methods.
tions can be clearly visualized, and includes the full GND plus SSD
dislocation density, when the dislocation density is reasonably low. The 6. Conclusions
information from ECCI clearly complements the strain and GND density
information from DIC/AFM and HR-EBSD. The need to be able to A sample of titanium alloy was prepared for microscopy and

14
R. Sperry et al. Materials Characterization 173 (2021) 110941

necessarily associated with slip band formation but rather tend to


particularly occur near grain boundaries. The resulting slip systems
identified were consistent with all other methods.
Strain gradients associated with the slip bands (particularly in the
presence of grain boundaries) provided another view of the remnant
dislocation field. Short range strain gradients in EBSD data were used to
generate a dislocation density map that was consistent with ECCI ob­
servations of residual dislocations. GND were then used to calculate
theoretical long-range rotations that were consistent with the measured
rotations. Using DIC displacement data, strain gradients were charac­
terized and indicated a required GND content; this was found to be
consistent with the GND maps as well, and related inferred active slip
systems.
The combined information from the different methods reviewed in
this paper is particularly useful to inform dislocation-based crystal
plasticity and strain gradient models that are being developed and
refined. The concurrent measurement of strain and orientation gradi­
ents, GNDs, and SSDs, is critical to guide and validate the developing
models.
Fig. 16. Comparison of GNDs calculated along a slip band from the strain map
using the GND density map from the EBSD data and the strain map from the DIC Data availability
data. Red dashed line shows the trend of the data with a slope of 1.062 and an
R2 value of 0.868 suggesting the strain map and the GND density map have The raw data required to produce these findings are available to
information about the same dislocations. (For interpretation of the references to download from doi.org/10.13011/m3-5jy8-0c92. The processed data
colour in this figure legend, the reader is referred to the web version of required to reproduce these findings are available to download from doi.
this article.) org/10.13011/m3-5jy8-0c92.

Table 2 Declaration of Competing Interest


Slip systems identified by each method in each grain. ECCI method was able to
identify dislocations on slip systems that were not creating visible slip bands. The authors declare that they have no known competing financial
Grain 1 Grain 2 interests or personal relationships that could have appeared to influence
Slip trace and Schmid factor
the work reported in this paper.
(0 1 1 0) [2 1 1 0] (1 1 0 0) [1 1 2 0]
(1 0 1 0) [1 2 1 0] (0 1 1 0) [2 1 1 0]
SEM-DIC method (0 1 1 0) [2 1 1 0] (1 1 0 0) [1 1 2 0] Acknowledgements
(1 0 1 0) [1 2 1 0] (0 1 1 0) [2 1 1 0]
ECCI method (0 1 1 0) [2 1 1 0] (1 1 0 0) [1 1 2 0] This work was supported at BYU by the U.S. Department of Energy
(1 0 1 0) [1 2 1 0] (0 1 1 0) [2 1 1 0] (DOE), Office of Science, Basic Energy Sciences (BES), under Awards DE-
(0 0 0 1) [1 1 2 0] (0 0 0 1) [1 2 1 0] SC0012587 and DE-SC00012483. Z. Chen and S. Daly were supported by
(0 0 0 1) [1 1 2 0] the U.S. Department of Energy, Office of Basic Energy Sciences, Division
(0 0 0 1) [2 1 1 0] of Materials Sciences and Engineering under Award DE-SC0008637 as
(1 0 1 1) [1 2 1 0] part of the Center for Predictive Integrated Structural Materials Science
(1 0 1 1) [1 2 1 0] (PRISMS) at the University of Michigan. S. Han and M. Crimp were
EBSD method (1 1 0 0) [1 1 2 0] (1 1 0 0) [1 1 2 0] supported by the Department of Energy, Office of Science, through
(0 1 1 0) [2 1 1 0] (0 1 1 0) [2 1 1 0] Award DE-FG02-09ER46637.

displaced under uniaxial tension to approximately 3% global axial strain References


to develop distinct bands of localized slip. SEM, DIC, EBSD, AFM, and [1] Z. Chen, S.H. Daly, Active slip system identification in polycrystalline metals by
ECCI data were collected from the sample. Active slip systems were digital image correlation (DIC), Exp. Mech. 57 (2017) 115–127, https://doi.org/
identified in two grains using slip trace analysis in conjunction with 10.1007/s11340-016-0217-3.
[2] Y. Yang, L. Wang, T.R. Bieler, P. Eisenlohr, M.A. Crimp, Quantitative atomic force
Schmid factors. From the high resolution DIC data, the RDR method was
microscopy characterization and crystal plasticity finite element modeling of
used to confirm the identified active slip systems with higher confi­ heterogeneous deformation in commercial purity titanium, in: Metall, Mater.
dence. This approach is particularly important where the global Schmid Trans. A Phys. Metall. Mater. Sci (2011) 636–644, https://doi.org/10.1007/
factor may not reflect the local stress field. s11661-010-0475-0. Springer.
[3] D. Hull, D.J. Bacon, Introduction to Dislocations, 5th ed., Elsevier Ltd, 2011
The AFM data was then used to augment the DIC data by providing https://doi.org/10.1016/C2009-0-64358-0.
out-of-plane displacements and therefore shears. Shear profiles gener­ [4] B.A. Simkin, M.A. Crimp, An experimentally convenient configuration for electron
ated from the DIC data and the AFM data were found to be comple­ channeling contrast imaging, Ultramicroscopy. 77 (1999) 65–75, https://doi.org/
10.1016/S0304-3991(99)00009-1.
mentary and consistent. [5] M.A. Crimp, B.A. Simkin, B.-C. Ng, Demonstration of the g⋅bxu=0 edge dislocation
Analysis of ECCI data identified residual dislocations along the slip invisibility criteria for electron channelling contrast imaging, Philos. Mag. Lett. 81
bands, confirming the identity of the activated slip systems in the two (12) (2001) 833–837.
[6] M. Pitaval, P. Morin, J. Baudary, E. Viaria, G. Fontaine, Advances in crystalline
grains; other dislocations lying on distinct slip systems, without devel­ contrast from defects, Scan. Electron Microsc. 1 (9177) (2019) 439–444.
oping distinct slip bands, were also observed. [7] S. Han, P. Eisenlohr, M.A. Crimp, ECCI based characterization of dislocation shear
The EBSD data was independently used to infer the active slip sys­ in polycrystalline arrays during heterogeneous deformation of commercially pure
titanium, Mater. Charact. 142 (2018) 504–514, https://doi.org/10.1016/j.
tems in the grains through measurement of long-range orientation gra­
matchar.2018.06.003.
dients along the slip bands. Such orientation gradients are not [8] B.A. Simkin, M.A. Crimp, T.R. Bieler, A factor to predict microcrack nucleation at
grain boundaries in TiAl, Scr. Mater. 49 (2003) 149–154.

15
R. Sperry et al. Materials Characterization 173 (2021) 110941

[9] S. Zaefferer, N.-N. Elhami, Theory and application of electron channelling contrast [22] I. Gutierrez-Urrutia, S. Zaefferer, D. Raabe, Coupling of electron channeling with
imaging under controlled diffraction conditions, Acta Mater. 74 (2014) 20–50. EBSD: toward the quantitative characterization of deformation structures in the
[10] M.A. Crimp, J.T. Hile, T.R. Bieler, M.G. Glavicic, Dislocation density measurements sem, JOM. 65 (2013) 1229–1236, https://doi.org/10.1007/s11837-013-0678-0.
in commercially pure titanium using electron channeling contrast imaging, TMS [23] Gwyddion – Free SPM (AFM, SNOM/NSOM, STM, MFM, …) data analysis software,
Lett. 1 (2004) 15–16. (2004). http://gwyddion.net/ (accessed June 12, 2020).
[11] M.A. Crimp, Scanning Electron microscopy imaging of dislocations in bulk [24] Z. Chen, R. Sperry, S. Han, Comparing EBSD, DIC, AFM and ECCI for active slip
materials, using Electron channeling contrast, Microsc. Res. Tech. 69 (2006) system identification in Ti7Al, 2020, https://doi.org/10.13011/m3-5jy8-0c92.
374–381. [25] X.L. Nan, H.Y. Wang, L. Zhang, J.B. Li, Q.C. Jiang, Calculation of Schmid factors in
[12] T.J. Ruggles, D.T. Fullwood, Estimations of bulk geometrically necessary magnesium: analysis of deformation behaviors, Scr. Mater. 67 (2012) 443–446,
dislocation density using high resolution EBSD, Ultramicroscopy. 133 (2013) 8–15, https://doi.org/10.1016/j.scriptamat.2012.05.042.
https://doi.org/10.1016/j.ultramic.2013.04.011. [26] Y.N. Wang, C.I. Chang, C.J. Lee, H.K. Lin, J.C. Huang, Texture and weak grain size
[13] A.J. Wilkinson, G. Meaden, D.J. Dingley, High-resolution elastic strain dependence in friction stir processed mg-Al-Zn alloy, Scr. Mater. 55 (2006)
measurement from electron backscatter diffraction patterns: new levels of 637–640, https://doi.org/10.1016/j.scriptamat.2006.06.005.
sensitivity, Ultramicroscopy. 106 (2006) 307–313, https://doi.org/10.1016/j. [27] L. Wu, S.R. Agnew, D.W. Brown, G.M. Stoica, B. Clausen, A. Jain, D.E. Fielden, P.
ultramic.2005.10.001. K. Liaw, Internal stress relaxation and load redistribution during the twinning-
[14] A.J. Wilkinson, E.E. Clarke, T.B. Britton, P. Littlewood, P.S. Karamched, High- detwinning-dominated cyclic deformation of a wrought magnesium alloy, ZK60A,
resolution electron backscatter diffraction: an emerging tool for studying local Acta Mater. 56 (2008) 3699–3707, https://doi.org/10.1016/j.
deformation, J. Strain Anal. Eng. Des. 45 (2010) 365–376, https://doi.org/ actamat.2008.04.006.
10.1243/03093247JSA587. [28] R. Hill, The elastic behaviour of a crystalline aggregate, Proc. Phys. Soc. Sect. A. 65
[15] J.R. Seal, T.B. Britton, A.J. Wilkinson, T.R. Bieler, M.A. Crimp, Characterizing slip (1952) 349, https://doi.org/10.1088/0370-1298/65/5/307.
transfer in commercially pure titanium using high resolution electron backscatter [29] B. Clausen, C.N. Tomé, D.W. Brown, S.R. Agnew, Reorientation and stress
diffraction (HR-EBSD) and Electron channeling contrast imaging (ECCI), Microsc. relaxation due to twinning: modeling and experimental characterization for mg,
Microanal. 18 (2012) 702–703, https://doi.org/10.1017/S1431927612005363. Acta Mater. 56 (2008) 2456–2468, https://doi.org/10.1016/j.
[16] R. Sperry, A. Harte, J.Q. da Fonseca, E.R. Homer, R.H. Wagoner, D.T. Fullwood, actamat.2008.01.057.
Slip band characteristics in the presence of grain boundaries in nickel-based [30] H. Li, C.J. Boehlert, T.R. Bieler, M.A. Crimp, Analysis of slip activity and
superalloy, Acta Mater. 193 (2020) 229–238, https://doi.org/10.1016/j. heterogeneous deformation in tension and tension-creep of Ti–5Al–2.5Sn (wt %)
actamat.2020.04.037. using in-situ SEM experiments, Philos. Mag. 92 (2012) 2923–2946, https://doi.org/
[17] B.E. Dunlap, T.J. Ruggles, D.T. Fullwood, B. Jackson, M.A. Crimp, Comparison of 10.1080/14786435.2012.682174.
dislocation characterization by electron channeling contrast imaging and cross- [31] M.F. Ashby, The deformation of plastically non-homogeneous materials, Philos.
correlation electron backscattered diffraction, Ultramicroscopy. 184 (2018) Mag. 21 (1970) 399–424, https://doi.org/10.1080/14786437008238426.
125–133, https://doi.org/10.1016/j.ultramic.2017.08.017. [32] J.C.M. Li, Y.T. Chou, The role of dislocations in the flow stress grain size
[18] Z. Chen, W. Lenthe, J.C. Stinville, M. Echlin, T.M. Pollock, S. Daly, High-resolution relationships, Metall. Mater. Trans. A 1 (1970) 1145–1159, https://doi.org/
deformation mapping across large fields of view using scanning Electron 10.1007/BF02900225.
microscopy and digital image correlation, Exp. Mech. 58 (2018) 1407–1421, [33] Z. Shen, R.H. Wagoner, W.A.T. Clark, Dislocation pile-up and grain boundary
https://doi.org/10.1007/s11340-018-0419-y. interactions in 304 stainless steel, Scr. Metall. 20 (6) (1986) 921–926, https://doi.
[19] A.D. Kammers, S. Daly, Self-assembled nanoparticle surface patterning for org/10.1016/0036-9748(86)90467-9.
improved digital image correlation in a scanning Electron microscope, Exp. Mech. [34] GitHub - BYU-MicrostructureOfMaterials/OpenXY, (n.d.). https://github.
53 (2013) 1333–1341, https://doi.org/10.1007/s11340-013-9734-5. com/BYU-MicrostructureOfMaterials/OpenXY (accessed June 12, 2020).
[20] G. Frens, Controlled nucleation for the regulation of the particle size in [35] T.J. Ruggles, T.M. Rampton, A. Khosravani, D.T. Fullwood, The effect of length
Monodisperse gold suspensions, Nat. Phys. Sci. 241 (1973) 20–22, https://doi.org/ scale on the determination of geometrically necessary dislocations via EBSD
10.1038/physci241020a0. continuum dislocation microscopy, Ultramicroscopy. 164 (2016) 1–10, https://
[21] E.J. Corey, A. Venkateswarlu, Protection of hydroxyl groups as tert- doi.org/10.1016/j.ultramic.2016.03.003.
Butyldimethylsilyl derivatives, J. Am. Chem. Soc. 94 (1972) 6190–6191, https://
doi.org/10.1021/ja00772a043.

16

You might also like