HETEROGENEOUS PLASTIC DEFORMATION
By
Songyang Han
A DISSERTATION
Submitted to
Michigan State University
in partial fulfillment of the
requirements
for the degree of
Materials Science and Engineering
Doctor of Philosophy
2020
ABSTRACT
HETEROGENEOUS PLASTIC DEF
ORMATION
By
Songyang Han
-
titanium,
because the hexagonal crystal structure makes it more prone to polycrystalline compatibility
issues. At the dislocation scale, this compatibility involves the process of ho
w dislocations
being initiated, propagating through grains, and the ability of grain boundaries accommodating
dislocation shear in one grain with shear in its neighboring grain. To characterize the
development of plasticity in a polycrystalline array due
to dislocation nucleation, and slip across
grains and grain boundaries, electron channeling contrast imaging (ECCI) based analysis is used,
since this special scanning electron microscopy (SEM) technique possesses variety observation
scopes, providing a li
nkage between the macro
-
and the micro
-
world.
The first study presented a robust comparison between several techniques for the very
first time: the digital image correlation (DIC), atomic force microscope (AFM), ECCI, and EBSD
(cross
-
correlation). In
this study, a Ti7Al alloy was deformed to 3% plastic tensile strain. The
plasticity evolution of the sample was assessed through BESD
-
slip trace analysis, digital image
correlation, and ECCI contrast analysis. The comparison between different methods rev
ealed
ECCI as a powerful technique in slip system identifications.
The second project was more focused on the interactions between slip systems around
the grain boundary area. The geometry of slip planes and grain boundaries was assessed as a
funct
ion of depth, allowing the analysis of slip transfer parameters, including the geometric
slip planes. Locally accommodation behavior at the grain boundaries w
ere revealed by ECCI.
The third project was the identification of the propagation direction of a slip system
across a polycrystalline grain patch by ECCI. Analysis indicated that slip bands would likely to
become broader as they propagated further into th
e grain from the nucleation points, possibly
due to cross
-
slipping. Together with the trace analysis, a better understanding of the
development of plasticity within polycrystals during heterogeneous deformation was achieved.
The highlight of this work not
only focusing on the infinitesimal change in the local
lattice structure in terms of dislocation nucleation and propagation in a grain, but also involves
the plasticity development of deformation in a macroscopic view, such as the mechanism of
dislocation
across grain boundaries, and the estimation of overall deformation behavior within
a region of grains. More importantly, together with other powerful characterization methods,
ECCI in this study shows a strong potential that successfully links the macros
copic deformation
with dislocation movement during the deformation.
iv
To my wife Dongqing Tao and daughter Freya Han. Thank you for your love and supports.
v
ACKNOWLEDGEMENTS
First and foremost, I would like to gratefully acknowledge my advisor, Dr. Martin Crimp.
Over the numerous conversations of research ideas,
paper discussions, and even non
-
academic
topics, I continuously sharpen my writing skills, being more diligent as a researcher, and
becoming more confident and expressive as an overall person. These improvements are mostly
relied on his wisdom, patience,
and the confidence in me. It will be an honorable memory as a
graduate student of his.
I also give my thanks to the committee members, Professor Philip Eisenlohr, Professor
Carl Boehlert, and Professor Thomas Pence, for their enlightening discussions.
T
(Jason), Harsha Phukan, Mingming Wang for helping me in the research. I also want to thank
Dr. Per Askeland on training the scanning electron microscope and teaching me th
e
maintenance of microscope.
Many thanks go to Dr. Christopher Cowen, formerly at National Energy Technology
Laboratory, Albany OR, for supplying the titanium studied in this work. Acknowledgements also
go to Professor Samantha Daly, and Dr. Zhe Chen
, at University of California Santa Barbara, for
the DIC support, and Professor David Fullwood, at Brigham Young University, for the CC
-
EBSD
support.
Finally, the research is supported by the United States Department of Energy, Office of
Basic Energy Scien
ce Division, through grant number DE
-
SC0001525 and DE
-
FG02
-
09ER46637.
vi
TABLE OF CONTENTS
x
1.
1.1.
1.1.1.
1.1.2.
1.1.3.
1.1.4.
1.2.
Int
1.2.1.
1.2.2.
1.2.3.
Electron
1.2.4.
1.2.5.
1.2.6.
Free surfacing biasing and
limitation of surface
-
1.2.7.
1.3.
2.
37
2.1.
2.2.
2.2.1.
Deformation of Ti
-
2.2.2.
Deformation
40
2.3.
2.3.1.
2.3.2.
2.3.3.
HR
-
3.
3.1.
The overall status of as deformed samples 1
-
3.2.
Slip system
7
3.2.1.
3.2.2.
3.3.
The comparisons of
surface
-
3.3.1.
3.3.2.
vii
3.3.3.
Comparison between the
3.3.4.
3.4.
3
3.4.1.
Categories of slip sy
3.4.2.
3.4.3.
3.4.4.
Comparison of slip
3.4.5.
3.4.6.
3.5.
The direction of sl
3.5.1.
7
3.5.2.
Comprehensive analysis of deformation within grain 1
-
3.5.3.
Conclusions
4.
5.
APPENDICES
01
. 102
APPENDIX B: Removal of polymer film and gold nanoparticles (AuNPs) for DIC patterning
.
106
APPENDIX C: Strain meas
urement after four
-
APPENDIX D: The calibration of MIAR III FEG
-
SEM with SACPs module for ECCI and CC
-
EBSD
APPENDIX E: Procedures of EC
APPENDIX H: Calculation of the
149
BIBL
I
OGRAPHY
viii
LIST OF TABLES
Table 1
.
Slip systems
identified by EBSD slip trace analysis, ECCI, and AFM/DIC
Table A1. Electropolishing parameters
ix
LIST OF FIGURES
Figure 1 Top two crystals present the major slip systems activated during deformation, namel
y:
(Top left) prism slip system on {
} prismatic plane, basal slip system on {0001} basal
plane, and pyramidal slip system on {
} pyramidal plane with <
> Burgers vector;
(Top right) type I slip system on {
} pyramidal pla
ne, and type II on {
}
pyramidal plane. Bottom two crystals show the geometry of two tensile twinning (labeled in
blue and yellow) and two compression twinning (labeled in purple and dark green), which is not
readily observed in room temperatur
e deformation and not the interest of this study
Figure 2a) Sketch of a pure tilt boundary, indicating two crystals are rotated by angle
tilt
about
a rotation axis lying on the grain boundary plane. b) A pure twist boundary is formed, which
looks like a single crystal is twisted into half along a rotation axis lying perpendicular to the grain
boundary with misorientation angle
twist
. c) Sket
ch of
boundary, formed by a 36.9
o
rotation
between two same lattices about a common [001] axis. In this picture, the atom A and B in grain
1 are represented by circles with no fill, while the same atom in grain 2 are pattern filled circles,
and the gr
ain boundary lattice area is limited by the dotted square. After carefully counting, it
can be found that every 2 out of 10 atoms are sharing the same lattice position. It is noticeable
that
value may different about different rotation axis.
(
A
mended
from [57, 65])
Figure 3a) Direct transfer of dislocation across grain boundary. b) Direct slip transfer with
residual dislocations at the grain boundary.
c) Absorption of dislocation slip and dissipated
along grain boundary. d) indirect slip
transfer by absorption and re
-
emission, leaving grain
boundary dislocations. e) Absorption and reflection of dislocations slip with residual grain
boundary dislocations. f) Complicated mechanism, involving both slip transfer and reflection
with the form
ation of grain boundary dislocations
Figure 4 As crystal A with known lattice orientation is continuously deformed, a known type
dislocation slip (blue) is piling up at the grain boundary, where dislocation slip (red) will
be
activated in crystal B due to stress build
-
up as it deformed with crystal A. During the slip transfer,
tangential continuity constrain is required that requires the strain component induced by
dislocations in crystal A be fully balanced in crystal B to
maintain grain boundary integrity.
(Amended from [80])
Figure 5 A sketch of the geometry of slip planes intersecting at a grain boundary plane.
b
,
t
,
n
,
are the Burgers vector, the intersection line direc
tion of the slip plane in the grain boundary
plane, and the slip plane normal that are used in the various slip transmission criterion [80, 81,
grain, and slip
Figure 6a) The simplified mechanism of the formation of electron backscatter patterns (EBSPs).
x
incident beam. b) A Kikuchi map is formed by collecting a
ll the backscatter signals coming out of
different planes on the phosphor screen. c) With each zone axis identified and labeled, one is
able to know the crystal orientation of the scanned grain. This is extremely useful in the
predication of slip systems
during plastic deformation.
(Amended from [107])
Figure 7 Sketched mechanism of
g
b
= 0 and
g
b
x
u
= 0 invisibility criteria in the determination
of dislocation Burgers vector. The dislocations will go completely out of contrast or show low
contrast when the dislocation plane lies parallel to the channeling direction, because the it is
where almost al
l electrons diffracting between planes in the same way, leaving no intensity
differences between distorted region and perfect lattice. On the contrary, larger value of
g
b
suggests more intensity variation around the distorted region, revealing dislocat
ions in better
contrast.
(Amended from [124])
Figure 8a) An example of ECPs collected from a p
-
type boron doped synthetic diamond single
crystal with close to [110] crystal orientation in low mag BSE mode (~ 20x).
The band in the
upper left corner with bright and dark contrast is one of the Kikuchi band formed during incident
attached on the surface. b) An example of EBSPs
collected among one of the grains from a
commercially pure titanium sample in this research (sample 2). The image was taken at a working
distance of 24 mm, a 30 kV accelerating voltage and a 184 µA probe current, with sample tilted
at 70
o
. The edges of
Kikuchi bands are significantly sharper than that of ECPs. c) An example of
SACPs collected from the same target with b) using the same voltage and current, but the
working distance is around 9mm within 10
o
tilt. SACPs provide much accurate information w
here
the closest zone axis the crystal is orientated. With more sharp edges on the channeling bands
and higher special resolution, SACPs fits ECCI analysis more than the other two options
Figure 9 Schematic mechanism on the formation of ECPs (A
mended from [146]), with the incident
beam sweeping angle large enough, intensity of BSE signal changes significantly around the
Figure 10a
-
c) Mechanism of the formatio
n of SACPs. As beam trajectory changes or beam rocking
around a certain point, the lattice channel become open and close with respect to the directions
of the incoming electrons, providing different yield of backscatter electron
. The signal profile
is c
ollected and create a SCAPs on the detector. For a channeling condition that allows the most
electrons channeling into the perfect crystal and leave an overall dark background, lattice
distortion around a dislocation will make more backscatter electrons c
ollected by the BSE
detector. Dislocations will be resolved. d) An example of dislocations (bright) from the dark
background.
Figure 11 An example of the change of channeling contrast with respect to the
deviation
parameter
s
, which s = 0 indicates the optical axis is exactly at the edge of the channeling band.
With optical axis move into or away from s = 0, signal intensity will change dramatically. As the
lattice is no longer aligned symmetrically tow
ards the incident beam due to dislocation distortion,
xi
contrast will occur with bright/dark contrast around a dislocation compared to the overall grey
background. (Amended from [152])
Figure 12a) Overall mechanism of D
IC. The surface is coated with evenly deposition of
nanoparticles, with fiducial marks. The reference image is the upper right square area labeled
with four fiducial marks, with a reference point P (x,y). During deformation, arbitrary shape
change and r
to surface topography, the absolute height difference (Z) is recognize
d by the laser reflection on
the position sensitive detector and recorded upon each position (X) from the starting point.
After scanning the whole area line by line, a 3
-
D topographic map can be created by correlating
height profile (Z) with the plane pro
file (X,Y).
Figure 13 The overall mechanism of the slip trace analysis. The hexagonal cell presents the
crystal orientation, and the red line is the intersection line between the slip plane (grey) and the
sample surface (b
lue). With the profile of each trace (1~12) at the surface, possible slip system
can be identified.
Figure 14a) The dimension of the Ti
-
7Al dog
-
bone tensile sample 1. b) The dimension of the CP
Ti bending
sample 2 & 3.
Figure 15 Sketch of the electropolishing stage. Based on what type of electrolyte is used, the
voltage, temperature while electropolishing, the distance between cathode (stainless steel) and
sampl
e (anode), and the agitating speed of the stir bar will be different and recorded in
Appendix
A
.
Figure 16a1) The SE image from the center of the 3% tensile strained Ti
-
7Al sample 1, with the
tensil
e direction along A2 axis. Almost all grains were deformed, with some grains having more
than one type of slip traces. Surface and grain boundary elevation could be indicated by the
brighter contrast against the dark grains. a2) The corresponding EBSD d
ata in the red region of
a1, which was collected under the same coordinate system as a1. Despite the noises due to
higher deformation strain that disrupts the diffraction (with confident index only 0.65), color
gradient within grains can be clearly seen,
especially where slip bands were densely packed. Slip
traces appear mostly straight, while some curved traces were found near grain boundaries or in
the grains which were heavily deformed. b1) An example of one of the grain boundaries between
two neighbo
ring grains in the center area of sample 2 after 1.5% deformation on four
-
point
-
bending stage. Again, most slip traces appear straight, while the bright and dark contrast on the
slip traces may indicate they may not share the same Burgers vector. Despite
some primary slip
traces that were fully propagated across the grains, some of the slip traces disappear as they
propagated out of the grain boundary. b2) The corresponding EBSD map of the red area of b1,
indicating the formation of slip bands was not si
gnificant enough to affect the quality of the
EBSD, with a high confident index of 0.81, and the topography developed during deformation
was not big enough to be shown in the EBSD compared to a2. c1) SE image of a patch of 4
consecutive grains in the cent
er of sample 3, which was deformed in the same manner with
xii
sample 2 to 1% plastic strain. Slip traces in grains 2
-
4 appear straight while traces in grain 1 are
wavy. Some traces disappear near the grain boundary in grain 2, which is indicated by the whit
e
arrow. Traces in grains 1 and 2 meet at the same point on the grain boundary, so are the traces
in grains 3 and 4, while traces in grains 2 and 3 do not meet at the same point on the grain
boundary. c2) EBSD map of the grain patch in c1. The deformati
on is not large enough for EBSD
to recognize the formation of slip bands at 1% strain level. No crystal rotation is detected. It
should be noted that all the prism cells in a2, b2, and c2 indicate the crystal orientations of the
grains they were located,
with the black dots indicating the optical axis.
4
4
Figure 17a) BSE image of grain 2 in sample 1. Two straight and planar slip traces can be clearly
observed, with their trace outlined as black lines. The hexagonal cells on the right indi
cate the
orientation of this grain. The shaded plane in each cell indicates the slip plane, the blue line
indicates the slip direction, and the red dashed line is the plane trace on the sample surface. For
a potential slip system active during deformatio
n, the red dashed line should match the trace on
the surface. b) BSE image of a patch of grains 1~3. Some slip systems with different Burgers
vectors or slip planes may exhibit similar slip traces at the surface, resulting in uncertainties. The
colored
dashed lines in each grain represent the possible slip systems that leave similar slip
traces. The Burgers vectors and slip planes of all possible slip systems are listed on the right.
The color of the slip system is using the same color with the dashed
line.
Figure 18a) BSE image of one of the areas of interest after deformation of sample 1. b
-
f) ECC
images taken at different channeling conditions from the red boxed area. The upper left circles
are the Kikuchi patterns. Each black ar
row across a Kikuchi band indicates the specific channeling
band, and the arrowhead is where the optic axis is focused. Each channeling condition is
identified from the T. O. C. A software. The Burgers vectors of dislocations are identified by
g
.
b
= 0
and
g
.
b
x
u
= 0 contrast analysis, and the slip plane can be revealed through different tilting
and rotating along a certain channeling band (
Appendix E
). A total of four different slip systems
are identified and labeled by colored arrows, namely: (01
0)[2
0] pri
sm slip system (green),
(10
0)[1
10] prism slip system (blue), (10
1)[1
10] pyramidal slip system (purple), and
(0001)[11
0] basal slip system (red). Amended from [185]
Figure 19a) SE image of the upper right corner o
f grain 2 in Ti
-
7Al sample 1. b) AFM color
-
scale
map from the black boxed area in figure a). The black line is the AFM line profile showing the
topography change across the line. The black arrows indicate the edges of slip bands, and the
dashed line ind
icates the undistorted surface plane and is the basis of the height measurement.
c) 3
-
D Greyscale topography map around the line sectioned area, the surface normal is calculated
based on undistorted surface, H is the step of the slip band. d
)
A sketch of
the mechanism to
calculate the local shear distribution across the slip bands. The Burgers vector
b
and slip plane
normal
n
can be directly achieved from the EBSD data once the slip system is confirmed by ECCI
contrast analysis. Height difference across
a slip band H can be directly measured from the AFM
line profile. The distance across a slip band X is 0.3 µm in this study.
(Amended from [185])
.. 53
Figure 20a) color
-
scale AFM map from an area in grain 2. b) The heat map of local shear
distribution across individual slip bands. The local shear contributed by the (01
0)[2
0] prism
xiii
slip system ranges from 0.3 to 0.7, and the shear caused by the (10
0)[1
10] prism slip
system ranges from 0.08 to 0.15. c) The relative shear distr
ibution map of the same area
calculated by the DIC method [185].
Figure 21a) ECC image of the same area with Figure 20. Besides the (01
0)[2
0] (green) and
(10
0)[1
10] (blue) two prism slip systems identified
in the slip trace analysis, (0001) [11
0]
(red) basal dislocations are revealed by the contrast analysis. The prism type dislocations
appear to align perfectly along the slip bands, while the basal dislocations are less uniformly
distributed a
cross the observed area. b) The GND logarithm map of the red boxed area of a).
The GND map is consistent with ECCI observation, although individual dislocations cannot be
resolved as good as the ECCI observation. c) The ECC image at the grain boundary o
f grain 2,
with (01
0)[2
0] (green) and (0001) [11
0] (red) dislocations. Prism dislocations align
close to the slip bands and basal dislocations are more randomly distributed across the
surface. d) The GND map from the red box area of c). The
GND on the other side of the grain
boundary is not available due to misorientation angle exceeding the threshold from the reference
point in grain 2. The GND map is consistent with ECCI observation.
Figure 22a) AFM color
-
scale topograp
hy map at the upper right corner of the grain 1 in sample
1. Two different deformation shear system can be observed by the AFM. b) ECC image of the
same area, two slip systems were identified through the EBSD
-
based slip trace analysis, which
are the (0
1
0)[2
0] prism slip system (blue) and the (
100)[11
0] prism slip system
(red). The slip trace marked in light brown cannot be identified by the slip trace analysis. c) High
magnification ECC image of the red box area in b). Contrast analysis sh
ows several additional
dislocations, including (10
1)[1
10] and (
011)[1
10] two pyramidal slip systems and a
mixture of basal dislocations with majority belong to (0001)[1
10] slip system. The unknown
branched slip trace is caused by the (
011)[1
10] pyramidal slip system.
Figure 23a
-
-
same points on the grain boundary at surface in sample 2. It appears that slip traces that area
far from para
-
-
-
in sample 2. The slip traces do not intersect at the same point on the grain boundary at surface,
despite the slip system is highly activated. h
-
The strong activation of slip systems that propagate up to the grain boundary only occurs in one
of the grains, with the other grain undeformed. At certain circumstances, the unresolved shear
activate new slip systems that are travelling back into the
originated grain.
Figure 24a) 3
-
D geometry of slip system interactions at a grain boundary. The two slip systems
are considered as well
-
correlated
slip systems since they intersect at the same point at the grain
boundary on the free surface. Although slip systems intersect at the same black point at the
surface, it can be clearly seen that they do not meet below the surface since the geometric
orien
tations of the two slip systems are different, with
o
. b) Another 3
-
D geometry of slip
bands interactions in the vicinity of a grain boundary. The two slip systems are defined as non
-
correlated since they do not meet at the surface. However, they m
ay meet at the grain boundary
xiv
plane somewhere below the surface. Although the slip transfer mechanism may be different
between a) and b), current studies of slip transfer ignore this potential difference and directly
use the slip transfer criteria to eval
uate the strain accommodating events at the grain
boundary.
Figure 25a) SE image showing two slip traces that are well
-
correlated at the surface. The despite
a large angle between the slip traces, they
intersect at the same points on a grain boundary.
ECCI facilitated slip trace analysis indicates the slip system in the upper grain is the (0
10)[
110]
prismatic slip system with the Schmid factor 0.48. The slip system in the lower grain is
(10
0)[
2
0] prismatic slip system, with a lower Schmid factor of 0.36. The white arrows
mark one of the well
-
correlated slip traces for comparison. b) ECC image of the same region, but
5 µm below the surface. The observed slip bands are now due to dislocati
on contrast, rather
than topography. It can be clearly seen that the slip bands are misaligned in the electropolished
area, which is indicated by the relative position change of the white arrows. Nevertheless, the
relative spacing and distributions of th
e slip bands remains unchanged. c) high magnification
ECC image of the area in the red boxed area in figure (b). It shows the activation of (
011)[
2
0]
pyramidal secondary slip systems (M =0.45) in the lower grain from an intersection point of
the
slip system form the upper grain at the grain boundary. The secondary slip system propagates
a short distance and appears to merge into the primary slip system in the lower grain. There
might be multiple activation of such secondary slip system, possibly
from different sources at
different depth in the grain boundary plane, which is indicated by the small arrows.
slip bands in the lower grain at the grain boundary. The
red dashed tangent line represents the trace of the grain boundary at the intersection point with
through el
ectropolishing. The grain boundary orientation is significantly changed, which is
indicated by the change of the red dashed tangent line around the same strongest activated slip
band. The slip band spacings and relative distributions remain the same befo
re and after surface
removal. c) Higher magnification ECC image of red tangent line area. A ~50 nm misalignment
between the two slip systems below the surface has been observed. d) ECC image of the red
boxed area in figure (c). A small number of second
ary slip system is observed in the upper grain
in this case.
Figure 27a) SE image of non
-
correlated slip systems on the surface. The small white arrows
indicate the relative positions of the selected sli
p bands on the surface. It can be clearly seen
that these slip bands do not intersect at the same point on the grain boundary. b) ECC image of
the same region below the surface. The relative spacing and distributions of the slip lines remain
unchanged w
ithin either grain, but the relative intersection points on the grain boundary changes,
as indicated by the small white arrows. c) ECC image of the grain boundary area in the red box.
Pyramidal secondary slip system is activated from the intersection
points of the incoming
prism slip system in the upper grain at the grain boundary. d) ECC image of the grain
boundary area in the blue box. Secondary pyramidal slip system is activated to accommodate
the strain induced by the prism slip syst
em in the lower grain.
xv
Figure 28a) SE image of the third interaction type. The slip system appears to be blocked by the
grain boundary and no active slip system is observed in the lower grain. b) ECC image in the red
boxed area. A nu
mber of dislocations within a limited zone is observed, with majority belong to
the basal plane, and some on the pyramidal plane.
7
Figure 29 Comprehensive analysis of slip transfer parameters between well
-
correlated slip (top
1
0 lines), non
-
of the angles
between incoming primary and outgoing primary (black)/ secondary (red) slip
system intersection lines at the grain boundary plane, and the g
between the incoming primary and outgoing primary (black)/ secondary (red) slip systems. b)
Comparison of the global Schmid factors M of the outgoing primary (black)/ secondary (red) slip
rs vectors of incoming and outgoing slip systems, and angle
same
Burgers vectors. Schmid factors for the outgoing primary slip systems are shown in italics
in the cases where there is no correlated slip observed between the incoming primary and
ases because only
limited slip was observed in the outgoing grains in the vicinity of the primary incoming slip
systems. The fine dotted line between the non
-
indicates the similarity in the accommodating beh
avior.
slip systems, b) the angles
between the various outgoing slip systems and the incoming primary
slip system, and c) glob
al Schmid factors M of the outgoing slip systems.
Figure 31a) SE image of grains 1
-
3 in the as
-
deformed sample 3. The slip systems are identified
by the ECCI facilitated slip trace analysis, which are marked by dash lines in different col
ors and
labelled in different colored fonts. It appears the slip lines in grain 2 and 3 are not well
-
correlated.
The slip systems in grain 1 appear to meet the slip systems in grain 2 at the boundary, while the
slip propagation in grain 1 is difficult. b
) General BSE image of the same grains 1
-
3 in the
electropolished sample 3. All the slip traces are removed by the electropolishing and no clear
contrast is seen because the sample is not under a channeling condition. c) One example of ECCI
analysis on t
he electropolished surface. At g = [1
], the slip line contrasts are distinct. These
contrasts come from dislocation contrasts rather than topography. d
-
i) High Mag ECC images
taken from the red boxed area in grain 3 for the Burgers vector identificat
ion. The Burgers vector
is [1
10]. With the information from the slip trace analysis, the slip system that is active in grain
3 is (10
0)[1
10] prism slip system. ECCI identification of slip system in grain 1 and 2 can be
found in the
Appendix F
, fo
llowing the same approach.
6
Figure 32a) electropolished surface of grain 3. b
-
f) ECC images taken at the same channeling
condition at different positions along a slip line. Dislocations are limited in a sharp and narrow
line
from the nucleation point at the grain boundary with grain 2, and spreading out as going
deeper into the grain and will finally distributed around the grain boundary with grain 4.
xvi
Figure 33a) Low mag ECC image of grain 3, the four arrows indicate th
e slip transfer directions of
slip bands with strong ECC contrast. The dashed arrow indicates the opposite transfer direction
-
f) ECC images following
the slip band from the boundar
y with grain 4 into grain interior. Overall dislocation density is
low compared with the case in Figure 31. As the slip band broadening effect further dilutes the
dislocation concentration, the dislocation contrast of the slip band is hardly detectable.
Figure 34a) A wide spread of dislocations in grain 1 from where a dislocation slip in grain 2 is
nucleated at the grain boundary. b) Wide spread of dislocations in grain 1 interior. c) a sharp
slip band is nucleated from the boundary in
grain 2. d) The slip band is wider in grain interior
than it is at the grain boundary. e) Intersection of dislocation slip at the grain boundary with
grain 3, slip systems in grain 2 and 3 are not correlated. Dislocation density is high in grain 2. F)
Accommodating dislocations nucleate from the intersection point of a slip system in grain 3, and
propagate only a short distance with significant broadening effect.
3
Figure 35 Deformation slip in the lower grain interacting with the d
eformation twin in the upper
grain. ECCI analysis in the red boxed area shows the dislocation propagating out of the tip of the
deformation twin.
Figure A1 A scheme shows the general four stages of electropol
ishing with respect to different
voltage and current density ratio.
Figure A2 Left) Below 29 V results in etching of metal with rough surface under optical & electron
microscope. Middle) A good polishing zone result
s in shinning & smooth surface with good
contrast under electron microscope. Right) Above 40 V results in a dimmer surface in optical
microscope. Under electron microscope, pitting occurs, especially at grain boundaries, with
slightly worse SACPs.
Figure A3 Calculation of the surface removal by the Vickers indent. Several Vickers indents are
placed at the surface before electropolishing and final result is the average of each calculated
one.
Figure A4a1) Surface condition of as received sample. a2) SE image of AuNPs taken at high
magnification from the red box, with weak SACPs due to interference. b1) Surface condition
after 1 hour at 30
o
C, showing rem
oval of majority patterning material. b2) SE image showing
one of the unremoved clusters of AuNPs, with sharp SACPs. Particles in these areas are the
focused point during DIC data acquisition. Long time exposure of beam may condense the NPs
into the mate
rial or strengthen the bonding interaction, with detailed mechanism unknown. The
patterns c1) At 4
th
hour, all nanoparticles are consumed, including the clusters. c2) SE image of
the same area with b2, showing clean sample, with sharp slip traces and gra
in boundaries. d1)
24
-
hour reaction time of an undeformed control sample 1, showing the slightly etching of
material. d2) SE image of the etched area in the red box, showing line type etched marks and
small etched cavities. e1) Surface condition of an u
ndeformed control sample 2 after the 3
rd
xvii
hour in solution at 50
o
C, also showing large areas being etched. e2) SE image of the red box
area, showing surface material has been etched away.
Figure A5a) SE image of a random area, sho
wing clear slip traces on surface after the uncoating.
b) 2
-
D AFM map showing topographic information of the same area. c) 3
-
D AFM map showing
clear steps from the slip band and the grain boundary (concaved).
Figure A6 Left) The dist
ance between triple points before deformation. Right) The distance
between same triple points after deformation.
Figure A7 Secondary electron (SE) images taken at the field of view of 3.86 µm that shows: a)
The particle is
out of focus. b) The dirt is in focus, but astigmatic. c) The dirt is stigmatic and in
perfect focus.
Figure A8a) SE image in resolution mode (field of view 3.86 µm) that shows the optic axis (black
cross
) is on a particle. b) SE image in field mode (field of view 65.3 µm) that shows the deviation
of optical axis from the resolution mode. Although it appears blurry, this particle is already in
focus in field mode since the resolution between modes is dif
ferent. c) SE image in field mode
that shows the optical axis is moved back to the same position after aperture alignment after
Figure A9a) An asymmetric SACP aperture with overlapping of p
atterns from surrounding grains.
b) A symmetric, perfect round aperture with interference signals from surrounding grains. c) A
perfect circular aperture within pattern only from the target grain.
Figure A10a) ECC image taken at s <0
, the contrast of dislocations in the lower grain is not perfect.
b) ECC image taken at s = 0, a crisp image of sharp dislocations in perfect contrast. c) ECC image
taken at s > 0, the dislocations in the same area are badly resolved with poor contrast.
Figure A11a) The dislocation tails are in weak contrast, as indicated by the small black arrows
while some other dislocation tails are in strong contrast, as indicated by the small white arrows
when ECCI is taken close to the upper zone axis. b
) The dislocations indicated by the small black
arrow are in strong contrast, on the contrary, the tails marker by the small white arrows are in
weak contrast when the optic axis moves closer to the lower zone axis.
Figure A12 The interface
of T. O. C. A. software shows 3 components in the simulation display,
including the simulated patterns shown on the left pop
-
up window, the corresponding pole
figure shown on the upper
-
right window, and the crystal orientation shown on the bottom
-
right
wi
ndow. The most useful parameters input in the control panel are listed as: 1. The Euler angle;
2. The 90
o
rotation; 3. The acceleration voltage at which ECCI is taken; 4. The magnification at
which ECCI is taken.
122
Figure A13 Left) Simulated channeling patterns with these bands labeled. Right) The real patterns
collected from channeling mode in MIRA III SEM, with the optic axis labeled as black cross.
xviii
Figure A14a)
G
rain 1 as deformed. b)
G
rain 1 after electropolishing. c
-
h) ECC images from the
red box area taken at different
g
vectors, dislocations are (1
)[11
0] and (1
)[11
0].
Figure A15a) Grain 2 as deformed. b) Grain 2 after electropolishing. c
-
h) ECC images from the
red box area taken at different
g
vectors. Dislocations are (1
)[11
0].
Figure A16a) Grain 3 as deformed. b) Grain 3 after electropolishing
. c
-
h) ECC images from the
rex boxed area at different channeling conditions. Dislocations are (10
0)[1
10].
Figure A17a) Grain 4 as deformed. b) Grain 4 after electropolishing. c
-
h) Dislocations at the
boxed area (grain boundary between gr
ains 3 and 4) taken at different channeling conditions.
Dislocations are majority (10
0)[
2
0].
Figure A18 One example of dislocations identification in sample 2 after electropolishing. a) ECC
image of slip traces aft
er electropolishing. The contrast is not due to topography but from the
contrast of dislocations. b
-
f) Dislocations taken at different
g
vector. Dislocations are (10
0)
[
2
0].
Figure A19 One exampl
e of dislocations identification in sample 2 after electropolishing on the
other side of the grain. Dislocations are (01
0) [
110].
Figure A20 Neighboring grains in Sample 2. First six ECC images) Identified dislocations are
(1
00)[11
0]. Second six ECC images) Identified dislocations are (1
00)[11
0] and
(1
01)[11
0].
Figure A21 Neighboring grains in Sample 2. First six) Identified dislocations are (1
00)[
20] and
(
101)[11
0
]. Second six) Identified dislocations are (1
00)[
20] and (1
01)[11
0].
Figure A22 Neighboring grains in Sample 2. First six) Identified dislocations are (10
0)[
0]
and (
011)[1
10]. Second six) Identified dislocations are (0
10)[2
0]
and (01
1)[
110].
Figure A23 Neighboring grains in Sample 2. First six) Identified dislocations are (10
0)[
0]
and (
011)[1
10]. Second six) Identified dislocations are (0
10)[2
0].
Figure A24 Neighboring grains in Sample 2
. First six) Identified dislocations are (10
0)[
0]
and (
011)[1
10]. Second six) Identified dislocations are (0001)[
20].
Figure A25 Neighboring grains in Sample 2. First six) Identified dislocations are (1
00)[
20] and
(1
01)[11
0]. Second six) Identified dislocations are (
00)[
20].
Figure A26 Neighboring grains in Sample 2. First six) Identified dislocations are (0
10)[
110].
Second six) Identified dislocations are (
0)[
0].
xix
Figure A27 Neighboring grains in Sample 2. First six) Identified dislocations are (1
00)[11
0].
Second six) Identified dislocations are (
0)[
0].
Figure A28 Neighboring grains in Sample 2. Fi
rst six) Identified dislocations are (1
00)[11
0]
and(1
01)[11
0]. Second six) Identified dislocations are (
0)[
0].
Figure A29 Neighboring grains in Sample 2. First six) Identified dislocations are (0001)[1
0]
and(
011)[1
0].
Second six) Identified dislocations are (
0)[
0].
Figure A30a) AFM color
-
scale topography map of grain 3 in sample 1. b) Slip systems identified
by the trace analysis are the (10
0)[
2
0] and (0
0)[2
0] prism slip systems. c)
ECC
image of the red box area in b). Contrast analysis also reveals two additional slip systems, which
are (1
00)[
20] prism slip system and the (0
11)[2
0] pyramidal slip system. The
pyramidal dislocations are responsible for the curvy sli
p traces. d) AFM color
-
scale
topography map of grain 5 in sample 1. e) (1
00)[11
0] prism slip system is identified by
the slip trace analysis. f) ECC images of the red box area in e). Contrast analysis also reveals a
significant number of (0
10)[2
0] prism dislocations within the observed area.
Figure A31 3
-
D geometry of the well
-
correlated slip systems at the grain boundary by the
correlation of surface and subsurface images. (Amended from [171])
Figure A32 All ca
ses slip interactions at grain boundaries with calculated results.
Figure A33 Slip band broadening effect observed along two different slip bands.
xx
KEY TO ABBREVIATIONS AND SYMBOLS
AFM
A
tomic
F
orce
M
icroscopy
Au
Gold
AuNP
G
old
N
anoparticles
BSE
B
ack
s
cattered
E
lectron
CC
-
EBSD
C
ross
-
C
orrelation
E
lectron
B
ack
s
cattered
D
iffraction
CRSS
C
ritical
R
esolved
S
hear
S
tress
CSL
C
oincident
S
ite
L
attice
DIC
Digital Image Correlation
EBSD
E
lectron
B
ackscattered
D
iffraction
EBSPs
E
lectron
B
ackscattering
P
atterns
ECCI
E
lectron
C
hanneling
C
ontrast
I
maging
ECP
s
E
lectron
C
hanneling
P
attern
s
EDM
E
lectronic
D
ischarge
M
achining
FEG
-
SEM
F
ield
E
mission
G
un
S
canning
E
lectron
M
icroscope
FIB
F
ocused
I
on
B
eam
FSE
F
orward
-
S
catter
E
lectron
GNDs
G
eometrically
N
ecessary
D
islocations
hcp
hexagonal close
-
packed crystal structure
HR
-
EBSD
High Resolution Electron
Backscattered Diffraction
LRB
Lee
Robertson
Birnbaum
xxi
MATLAB
M
atrix
L
aboratory
Ni
Nickel
OIM
Orientation Imaging Microscopy
OPS
C
olloidal
S
ilica
Suspension
SACP
s
S
elected
A
rea
C
hanneling
P
attern
s
SE
S
econdary
E
lectron
SiC
S
ilicon
C
arbide
TBAF
T
etra
-
n
-
B
utylammonium
Fl
uoride
TEM
T
ransmission
E
lectron
M
icroscopy
Ti
Titanium
Ti
-
7
-
Al
Titanium Aluminum Alloy
b
Burgers vector
g
vector that describes the electron imaging/channeling condition
angle between
Burgers vectors
l
rotation axis between intersecting lattices
M
Schmid factor
m'
predictive parameter of slip transfer using
n
slip plane
normal
S
stress
q[
a
ngle between
s
lip
p
lane normals
s
deviation factor
t
intersection line of a slip system on the grain boundary plane
xxii
qy
resolve shear
stress
qy
c
critical resolved shear stress
u
dislocation line direction
the reciprocal of total atoms at Coincident Site Lattice to the
tota
l atoms
rotation angle between two intersecting lattices
strain
P
stress on the slip
system
angle between intersection lines of slip systems at a grain boundary plane
Å
Angstrom, 10
-
10
meters
1
1.
Introduction
1.1
Heterogeneous deformation of commercially pure titanium
With the increasing demand for
-
titanium and titanium alloys across a wide range of
industries, due to their good corrosion resistance, high strength to density ratio, and biomedical
compatibilities [1
-
5], the desire to manipulate and precisely predict th
e performance of such
hexagonal materials during service also grows stronger. However, wide replacement of cubic
materials (i. e. steel) with such promising metals is still impossible, because the detailed
deformation evolution mechanisms of hexagonal tit
anium are not yet well understood. Cubic
makes it easier to precisely model the deformation textures [6
-
8]. On the contrary, such
exist in commercially pure
-
titanium (and other hcp
metals) [9
-
12] since the orientation change and slip
-
twin distribution are not consistent to give
predictable textures under the same deformation mode [13, 14]. Nevertheless, such
inconsistency challen
ges the establishment of a reliable model, and the key to elucidate this
unpredictable deformation behavior is the full understanding of the mechanisms of
heterogeneous deformation of hcp titanium.
1.1.1
Plastic heterogeneity of commercially pure titanium
Het
erogeneous deformation is common in all polycrystalline metals. Plastic
heterogeneity happens due to strain variations from grain to grain since all the grains
experience different deformation processes due to varying crystal orientations. In plastic
def
ormation models, Talyor and Houttee [15, 16] suggested that strains could be equally
distributed among all the grains under a macroscopically uniform deformation. This worked
2
well for materials with cubic symmetry because each grain is able to activate mu
ltiple slip
systems with comparable critical resolved shear stress (CRSS,
c
) to accommodate the
partitioned strain. Thus, heterogeneous deformation is not a serious problem for cubic
materials. However, this assumption failed on materials with lower sym
metry, such as hcp
titanium. As shown in
Figure 1
, there are several deformation slip systems in commercia
lly
pure titanium, including {
} <
> prismatic slip, {
} <
> basal slip,
Figure 1 Top two crystals present the major slip systems activated during deformation, namely: (Top left)
prism slip system on {
} prismatic plane, basal slip system on {0001} basal plane, and pyramidal
slip system o
n {
} pyramidal plane with <
> Burgers vector; (Top right) type I slip system
on {
} pyramidal plane, and type II on {
} pyramidal plane. Bottom two crystals show the
geometry of two tensile twinning (labeled in blue and yello
w) and two compression twinning (labeled in purple
and dark green), which is not readily observed in room temperature deformation and not the interest of this
study.
3
{
} <
> pyramidal slip, and {
} <
> pyramidal slip ({
}
pyramidal is rare). Tensile and compressive twinning does
exist as additional
deformation systems, but twinning activation is dependent on elemental composition [17] and
thus not easily predictable. Due to the anisotropic nature of the low symmetry of hexagonal
crystal lattice, the activities of these deformatio
n systems are dramatically different, with the
prismatic slip systems being the most observed among all deformation systems [18
-
25].
Such different activation of deformation systems is due to a large deviation in the CRSS value
among these deformation
systems [26]. Although CRSS values are sensitive to elemental
compositions [27
-
30] and dependent on testing methods [31
-
38], the CRSS of the prismatic slip
system is more likely found to be the lowest at room temperature in commercially pure
titanium
1
.
The CRSS value of the basal slip system is the second lowest, around 1.2 ~ 2.6 times
larger than the prism slip [28, 36, 37]. Pyramidal deformation systems, including pyramidal
and pyramidal slip systems, usually have CRSS values around 1
.3 ~ 8 times larger
than the prism slip system [25
-
28, 34
-
37]. As a result, it is much harder for other slip
systems, especially pyramidal slip systems, to be activated. As indicated by Von Mises [39], an
individual grain needs at least five independ
ent slip systems to accommodate change shape.
Thus, every titanium crystal needs to activate multiple different slip system types to maintain
polycrystalline integrity. In reality, this constraint is hard to achieve since slip activities in
titanium is n
ot uniformly distributed due to the variability of CRSS of the different slip system
types. Thus, plasticity models of titanium often fail since heterogeneity is not able to be
1
Basal may become lower than prism slip system at levitated temperature and in some alloys [38], that is why CRSS
value is material dependent.
4
correctly modelled to reflect the real
-
life deformation.
In addition to
the CRSS issue, due to the low symmetry of titanium, the deformation of
a single crystal through one type of slip system at a certain orientation usually means the
activation of the other slip system is not favorable. For example, if [11
] is favorably a
ctivated
with a high Schmid factor, then [1
10] and [2
] slip are typically not equally activated at the
same time since their Schmid factors are usually low. Thus, the deformation of a crystal has to
follow a certain direction, with other slip activati
on somewhat suppressed. However, in
polycrystal deformation, it is not quite possible to relief all the strain by the activation of the
primary slip system, and the suppression of other slip systems cause incomplete strain
relaxation, thus cause strain ac
cumulation within the grain. In a polycrystal, some grains are
orientations on
ly allow high
-
CRSS
-
value deformation systems [40
-
42]. Based on the uneven
distribution of plastic strain among the grains and the incomplete strain relaxation due to lack
of slip activation, strains that are not fully relaxed may accumulate at the grain b
oundaries.
1.1.2
Accommodation at grain boundaries
Micro
-
cracks will form at the grain boundaries if they fail to sufficiently accommodate
the strain localization on both sides [43
-
46]. At the grain
-
scale level, in order to maintain grain
boundary integri
ty, the resulting strain must be released by either transferring across the grain
boundary into the neighboring grain, or sometimes reflecting back to the original grain. At the
nanoscale or atomic level, dislocations that are carrying the strain get rest
rained at a grain
boundary, because the extensive atomic disordered grain boundary interface disrupts the
5
propagation of dislocations within a confined plane. The continuous piling
-
up of dislocations
around the grain boundary can be sources for the disloc
ation activations in either grains [47
-
48], and cause strain hardening of the grain boundary if the strain is not completely carried
dislocation slip encourages t
he further disruption of the grain boundary atomic configuration,
creating more defects and voids at the interface between the joining grains, and eventually
create micro
-
cracks on the grain boundary. This is believed as the precursor of material failure
[45, 50, 51]. In the heterogeneous deformation of titanium, the capability of a grain boundary
to accommodate the strain is difficult. As discussed previously, individual grains with varying
crystal orientations deform differently, so the grain boundarie
s have to deal with different
amounts of strain from different directions at the same time. As a result, it is critical to
understand the nature of grain boundaries and how grain boundaries accommodate the
heterogeneous strain from the grains.
1.1.3
The geome
try of grain boundary and its effect on dislocations
It is important to understand how a grain boundary reacts to dislocation shear based on
its geometric characters. Compared with the unique atomic arrangement within a lattice, a grain
boundary is a disordered interface between the joining crystals. Such
disordered structural
defects are usually high energy sites and can be dislocation sinks and sources [52
-
55] based on
how the lattice is misorientated at the grain boundary.
By and large, there are several different ways to describe the crystallography of
a grain
boundary, namely: tilt/twist boundary, symmetric/asymmetric boundary,
and the
boundary.
Tilt and twist boundaries are formed when two adjoining crystal lattices share a same rotation
6
axis
l
, as illustrated by
Figure 2 a&b
. For a pure tilt boun
dary, the rotation axis
l
lies in the
boundary plane. On the contrary, the axis
l
lies perpendicular to boundary plane in a pure twist
boundary [56]. This approach defines the configuration of a grain boundary by five
-
degrees of
freedom, including the gr
ain boundary normal
n
, two crystallographic orientations of the
neighboring crystals
n
1
and
n
2
, the rotation axes
I
, and the rotation angles
(the
total
misorientation can be broken down into a combination of
tilt
and
twist
about the respective
Figure 2a) Sketch of a pure tilt boundary, indicating two crystals are rotated by angle
tilt
about a rotation axis
lying on the grain boundary plane. b) A pure twist boundary is formed, which looks like a single crystal is
twisted into half along a rotation axis lying perpendicular to the grain boundary with misorientation angle
twist
. c)
Sketc
h of
boundary, formed by a 36.9
o
rotation between two same lattices about a common [001]
axis. In this picture, the atom A and B in grain 1 are represented by circles with no fill, while the same atom in
grain 2 are pattern filled circles, and the gra
in boundary lattice area is limited by the dotted square. After
carefully counting, it can be found that every 2 out of 10 atoms are sharing the same lattice position. It is
noticeable that
value may different about different rotation axis.
(
A
mended f
rom [57, 65])
7
rotation axes
l
)[57]. In this approach, the
geometry of a grain boundary cannot be simply
defined by a fixed rotation and twist angle, since different rotation and twist combinations can
achieve the same misorientation structure of the boundary lattice. Specifically, to further
describe a pure tilt
/twist boundary, the concept of symmetric and asymmetric boundary is then
applied based on the relationship of the tilt axis direction to the grain boundary plane. For
example, the symmetric boundary is defined when the grain boundary plane is mirrored ab
out
the shared axis direction. By specifying the family of the rotation axis and the information
discussed above (i. e. {210}<001> symmetric tilt boundary, with misorientation angle 53.1
o
), a
more precise definition of the grain boundary is thus presented
to outline the grain boundary
atomic structure [58
-
64] for molecular dynamic modeling. Another common approach that
describes the geometric configuration of the grain boundary is the
boundary, usually used
together with the
coincident site lattice (CSL
) model [65]. The grain boundary is simply viewed
as a region of interpenetrating lattice points between the neighboring misorientated grains.
There will be some lattice points in that region where the atoms from the adjacent grains
overlap. Those points
are called coincident site lattice points. The
value is the value of the
total number of atoms over the number of atoms that are in coincident sites. An example of
5 boundary is presented in
Figure 2c,
since 2 out of 10 atoms on the grain boundary p
lane are
in coincident sites. This boundary is formed by a 36.9
o
rotation between two perfect cubic
lattices about a common [001] axis. This pure geometric model categorizes some specific grain
boundaries out of the common boundaries and provides another way to quantify the
misorientation. Grain boundaries with low
q7
values (more atoms share the coincident sites)
suggest there are little mismatch and little lattice disorder between the adjacent grains.
8
Further studies have found that low
boundaries usually show unique behaviors than non
-
ones during deformation.
For instance, some low
boundaries are found to prevent creep
formation in a Ni alloy [66], while coherent twin
3 boundaries are particularly good for
inhabiting cracks [67], etc. Nonetheless, the energy barrier of the grain boundary for
dislocation tra
nsmission and nucleation is also found to be related to the
value. As indicated
by
[62], the energy barrier for dislocation transmission and nucleation at a
q7
3
boundary was particularly high, suggesting this boundary was an effective block
to dislocations.
In that study, kinetic factors including the geometry of the loading orientations of the bicrystals,
and the Schmid factors, were also found to have strong impacts on the dislocation/grain
boundary interactions.
It should be realized tha
t the grain boundary models discussed above
are mainly developed from cubic or other higher symmetry systems [52
-
69], until recently, K.
Glowinski et al [70] applied this concept to hexagonal systems. In the study, they categorized
the grain boundary geom
etry with rotation axes/planes and
q7
boundaries, and specified the
similarity and difference with the cubic systems in the atomic configuration. That study helped
establish a system to correctly represent the grain boundary configuration for the hcp syste
m.
All in all, these lattice models indeed provide clues on plasticity transfer across grain
boundaries in plasticity modeling, and specifically, reveal how dislocations dissociate and cross
-
slip at/within special grain boundary interfaces [58, 62, 68, 6
9]. However, for the convenience
of modeling
-
based studies and to reduce complexities, there are some assumptions or
simplifications in these models. Such compromises make the modeling less effective at
representing real
-
life heterogeneous deformation of
polycrystals, with the reasons listed as
follows:
9
Although there have been a considerable amount of studies of special grain
boundaries (i. e.
q7
3, 5, 11, etc.) that show consistent behaviors of the dislocations, the
behaviors of dislocations at non
-
q7
bou
ndaries with uncategorized disordered lattice
configurations cannot be easily predicted. The understanding of grain boundary
accommodations with dislocation shear in heterogeneous deformation should also
include general boundaries, and these special model
s do not work well.
In the models, the dislocation shear is usually started from an intentionally
induced defect within a grain and then propagated to the grain boundary. For
convenience, the dislocation type was also given, and the shear was considered t
o be
homogeneous among the same type slip bands within the grain. However, the direction
of real
-
life deformation shear as well as the activation of dislocations are not always
predetermined since the heterogeneous deformation is not limited to only one g
rain,
but also its surrounding grains. The amount of shear carried by each slip band was also
not equal during the deformation.
Modeling studies until recent usually limit the interactions between one
incoming dislocation with one grain boundary, includin
g new dislocation initiations or
the absorption/reflection of the incoming dislocation. However, except in special cases,
there will typically be activation of multiple slip systems, which make the grain boundary
accommodation events more complicated than
the models.
Collectively, a more reliable approach is thus needed to study the accommodation
behavior by efficiently revealing the dislocation/grain boundary interactions in hexagonal
titanium. The active slip systems should be correctly identified,
the direction of shear transfer
10
should be grasped, and the models should be also applied on common grain boundaries.
As is discovered by Sangid et al [62], the Schmid factors, the orientation of crystals, and
the geometry of grain boundary interfaces are
all important elements affecting the
interaction between dislocations and boundaries. Moreover, since dislocations are
found to dissociate into dislocation partials or cross
-
slip during propagation within grain
boundary lattices, these studies also sugge
st that residual dislocations left in the grain
boundary will be an important factor in the dislocation/grain boundary interaction
studies.
1.1.4
Interactions between dislocations at general grain boundaries
Different from studies that focused on the effect of g
rain boundary atomic
configurations on the dislocations, numerous studies have successfully illustrated the
interactions between dislocation slip and unspecified grain boundaries. Bayerschen et al. and
other researcher [71
-
75] have summarized several poss
ible accommodating mechanisms when
an incoming dislocation meets a grain boundary. These mechanisms are illustrated in
Figure 3
,
namely: a) The direct slip transfer of an incoming dislocation into the neighboring grain without
leaving any residual dislocations in the grain boundary. b) The direct transfer of an incoming
dislocation across the grain boundary by initiating a diffe
rent type of outgoing dislocation in the
joining grain and thus leaving residual dislocations in the grain boundary. c) The full absorption
of an incoming dislocation into the grain boundary. This process creates grain boundary
dislocations that can be m
oved elsewhere under applied stress. d) An indirect slip transfer
process, including the absorption of an incoming dislocation, and the re
-
emission of an
outgoing dislocation at the boundary. Since the incoming and outgoing dislocations are not
11
directly connected, this process usually involves the participation of grain boundary
dislocations. e) The absorption of an incoming dislocation and initiation of outgoing
dislocations back to the same grain. f) The direct slip transfer of d
islocation, creating new
dislocations both in the adjacent grain and back into the original grain.
Situations e) and f) are not as common as the other accommodating models. Li et al
[76] found that it was more energetically favorable for a dislocation t
o cross a grain boundary
than being reflected back. It should be noted that, although not common, such mechanisms
have been both observed by the in
-
situ transmission electron microscopy (TEM) studies [35, 77,
Figure 3a) Direct transfer of dislocation across grain boundary. b) Direct slip transfer with residual dislocations
at the grain boundary. c) Absorption of dislocation slip and dissipated along grain boundary. d) indirect slip
transfer by absorption and
re
-
emission, leaving grain boundary dislocations. e) Absorption and reflection of
dislocations slip with residual grain boundary dislocations. f) Complicated mechanism, involving both slip
transfer and reflection with the formation of grain boundary dis
locations.
12
78] and included in the plastic modeling [76,
79]. The mechanism f) emphasizes the conditions
for multiple activation of dislocation slip at a grain boundary during accommodation events,
which is important in the heterogeneous deformation.
Livingston and Chalmers [80] were among the first of several res
earchers studying the
activation of multiple slip systems in plastic deformation. In their studies, they induced a
deformation shear that was carried to the grain boundary through a known slip system by one
Fi
gure 4
) and studied how the shear was
-
crystal systems
with different orientation combinations were tested. The strain components (
,
, and
shown in
Figure 4
) cre
compatibility of the bi
-
crystal system. This theory was referred as tangential continuity, and it
Figure 4 As crystal A with known lattice orientation
is continuously deformed, a known type dislocation slip
(blue) is piling up at the grain boundary, where dislocation slip (red) will be activated in crystal B due to stress
build
-
up as it deformed with crystal A. During the slip transfer, tangential conti
nuity constrain is required that
requires the strain component induced by dislocations in crystal A be fully balanced in crystal B to maintain
grain boundary integrity.
(Amended from [80])
13
req
uired a total of four degrees of freedom within the two adjoining grains during the
accommodation events. This included the situations of one incoming slip system in the parent
grain A being accommodated by three outgoing slip systems in the receiving gra
in B, two slip
systems (including the incoming one that was already known) in A being accommodated by two
slip systems in B, and three by one, respectively. This concept considered the potential for self
-
accommodation, where the strain accommodation was n
ot limited to the neighboring grain,
but also the grain where the strain was originated. Nonetheless, with the combination of the
pile
-
up stress and the geometric tangential continuity, Livingston and Chalmers outlined a
criterion that was used to predict
P
i
= P N
1i
= P [(
n
1
n
i
) (
b
1
b
i
) + (
n
1
b
i
) (
n
i
b
1
where P was the stress of the slip system,
n
1
&
b
1
was the slip plane normal and Burgers
n
i
&
b
i
was the corresponding parameters for any active slip
This criterion is useful for predicting
the primary (and sometimes secondary) slip
prediction of minor sli
p systems, due to more complicated mechanisms at the boundary. By
removing the stress component that needed to be calculated/measured from case to case, this
criterion was later simplified to a pure geometric constraint. This criterion is now well known
as the N factor, and is widely applied in many later studies as a slip transfer criterion [81
-
87]:
N
in
-
out
= (
n
in
n
out
) (
b
in
b
out
) + (
n
in
b
out
) (
n
out
b
in
14
Influenced by Livingston and Chalmers [80] and their collaborators [88], sequent
ial slip
transfer criteria have been developed in order to predict the slip systems activated as a result
of shear accommodation at grain boundaries, with the geometry of slip systems and grain
boundary illustrated by
Figure 5
. Among these criteria, the M
2
factor was established by Shen
et al. [81], evaluating slip transfer events from a different perspective than the N factor:
M = (
t
in
t
out
) (
b
in
b
out
) = cos
cos
This criterion considere
d the angle
between the intersection of the line directions (
t
)
of slip planes at the grain boundary plane, and the angle
q/
between the Burgers vectors (
b
) of
the slip systems on both sides of the grain boundary. At the same time, Lee et al. [83
-
86] lai
d
2
M is usually a symbol of the Schmid factor (SF) in many research. To avoid the misuse of M, this factor is us
ually
used as the LRB factor after Lee et al.
Figure 5 A sketch of the geometry of slip planes intersecting at a grain boundary plane.
b
,
t
,
n
, are the Burgers
vector, the intersection line direction of the slip plane in the grain boundary plane, and the slip plane normal
that are used in the
various slip transmission criterion [80, 81, 95]. Slip plane I (blue) is usually considered from
15
out criteria based on the M factor. It agreed that slip transmissions can happen when M was
maximized, but two additional stress components should also be included: First, the resolved
Second, the magnitude of
residual Burgers vector left in the grain boundary should be minimized. This combination
criteria, known as the LRB criteria, provided significant insights regarding the importance of
residual Burgers vector in slip transmissions
as well as its influence on grain boundary
deformation [89
-
94].
Another more convenient criterion was outlined by Luster and Morris [95], referred as
the geometric compatibility factor:
n
in
n
out
) (
b
in
b
out
) = cos
cos
where
q}
vectors. This simplified version of the N factor has been extensively used [96
-
100] since the
angle
q}
between plane normals is easily acquired from electron backsca
tter diffraction (EBSD),
whereas the measurement of angle
requires grain boundary orientation assessment, which is
not available only through surface analysis.
Meanwhile, a
qp
function was created by Werner and Prantl [101], dealing with slip
transfer b
etween different phases:
qp
= cos (
q}
q}
) cos (
5
Slip transmission was expected only when the angle
q}
a limited value (
q}
c
= 15
o
c
= 45
o
). The application of this
qp
16
since it was mainly for intra
-
phase slip transmissions.
For the most part, the application of different criteria has fulfilled different
requirements in the study of slip transmissions [74,
75, 102], and in particular, combined
criteria that coupled some of the geometric parameters with accumulated shear stress
[74] or
the Schmid factor M [103] have made more statistically reliable predictions of slip activity.
However, one may realize tha
t despite the factor PN
1i
outlined by Livingston and Chalmers et al.
[80, 88] that have indicated the need for multiple activations of slip systems during slip
accommodation, as well as Shen et al. [81] that have discussed the observation of slip
multiplic
ity within the vicinity of grain boundaries, many follow
-
up criteria have become more
and more simplified, assuming:
one deformation system during one slip transfer activity, altho
ugh comparisons of
parameters between different slip systems are quite common.
Similar to the limitation of many modeling studies, the direction of
grain. Many studies have
focused on the cases where a known incoming slip piled
-
up at
Undoubtedly, the simplified criteria are extremely useful, especially when the target of
interest is limited to bicrystals. However, this is far from accurate in the study of
heterogeneous deformation of polycrystals. For the first assumption, based on b
oth the
independent slip system to maintain integrity) [39], it is necessary to have more than one
17
accommodating slip system to be activated to fully accommodate th
e strain at the grain
boundary
3
. Accommodation by multiple slip systems was recently reported by Su et al. [104].
e grain boundary. The activation of
double accommodation was a complicated competition between many factors including the
Burgers vectors. Despite this res
earch, until now, how other deformation systems affect slip
transmission is still not clear. For the second assumption, it only worked perfectly in the bi
-
crystal system in an ideal condition but not precise in the heterogeneous deformation, where
shear t
ransfer is not necessarily limited to one given direction. As reviewed by Bayerschen et
train is primarily
in real
-
independently due to the applied stress, and the
needs to be transferred out at the grain boundary. Thus, it is not reasonable to say the
accordingly. Nevertheless, these ideal models neglect the situation
that dislocations can be nucleated at the grain boundary and carry the accumulated shear out
of the grain boundary by propagating into both grains. This is also an important mechanism to
3
Although grain boundary dislocations can also carry away accumulated strain at the boundary,
the
migration of
grain boundary dislocations may cause
severe
grain boundary
movement or
cracking.
18
protect the integrity of the grain boundary, since the grain boundary can be a source for
dislocation activation. So far, whether slip transfer criteria can be effectively applied in this
situation is still unanswered.
Thus, in order to further the und
erstanding of heterogeneous deformation of
commercially pure titanium, it is necessary to figure out if the classic slip transfer criteria or
initiation at a g
rain boundary and propagation into the adjoining grains). Additionally, it may
also be necessary to identify the direction of strain transfer within patches of grains. If possible,
it is insightful to identify which grain is actively deforming with respe
ct to the applied stress and
which is deforming passively to accommodate the deformation of its neighbor. By extension, if
one is able to locate the grain boundaries where the flow of strain is concentrated and is not
able to be well accommodated, such gr
ain boundaries may be vulnerable to damage nucleation
during the deformation.
1.2
Introduction of experimental techniques
There are generally several analytical methods used to identify the deformation slip
systems, the nature of the dislocations (in terms o
f Burgers vectors, slip directions and slip
planes), and the relative strain distributions across the deformed material. Rather than simply
laying out numerical expressions that are boring and non
-
intuitive, the following sections
provide a brief introduc
tion and a comparison of different analytical methods that are used for
dislocation
-
level characterization. The purpose of this section is to elucidate the advantage of
using electron channeling contrast imaging (ECCI) in this study, since it is capable
of both
grasping the macroscopic deformation of the material and providing microscopic detailed
19
information including the nature and relative distribution of the dislocations. When carefully
planned, ECCI is able to avoid the biasing of the free surface a
nd provide information on how
deformation shear is accommodated within and between grains associated with other
techniques.
1.2.1
Electron backscatter diffraction (EBSD)
EBSD is used to acquire the crystal orientation distributions and the changes in
orientatio
n during the loading process, both of which are important information in the study of
heterogeneous deformation. The overall set
-
up for the EBSD technique is shown in
Figure 6a
.
The crystal orientation of a grain is achieved based on the electron backsc
attering patterns
(EBSP), which appears as a map of intersecting pairs of parallel Kikuchi
-
lines on a phosphor
screen (
Figure 6b
) [106, 107]. In order to maximize the backscattered signal collected by the
detector, a high surface normal tilt angle of 70
o
is generally used. As the incident beam
electrons inelastically scattering in all directions within a crystal, some electrons hit the crystal
4
and will be elastically scattered and form reinforced beams
of electrons ex
iting the sample surface. As the inelastically scattered electrons vary in
form a surface of a cone, referred as the Kossel cone. Since the scattering even
ts occur in a
very small volume and therefore can be considered to occur at single planes, a lattice plane
thus will be represented as a pair of Kossel cones, which manifest as two parallel Kikuchi lines
4
n
= 2 d
hkl
si
n
B
, where n is a positive integer,
is the wavelength of the electron beam, d
hkl
is the distance of the
Miller indexed (h k l) lattice plane, and
B
constructive signals. Based on diffe
rent lattice structure, some combination of h, k, l will result in constructive
reflections, enhancing the signal, while in other cases results in destructive/forbidden reflections and thus give
weak signal.
20
when the two cones are projected onto the detector
screen. With Bragg diffraction occurring
from all structure factor allowed planes, cone pairs will develop from all allowed set of planes,
and diffraction patterns are formed as intersections of numerous Kikuchi bands (
Figure 6b
).
Based on this mechanism diffraction patterns provide angular information of the crystal. For a
known material, the identification of Kikuchi bands (
Figure 6c
) through the Hough transform
[108] will reve
al its crystal orientation. After the application of EBSP in the 1970s [106],
continuous development of automatic patterning and phase identification methods [107
-
110]
have made the EBSD a widely used scanning electron microscopy technique for near surfac
e
characterization. EBSD is able to provide accurate information on crystalline orientation
Figure 6a) The simplified mechanism of the formation of electron backscatter patterns (EBSPs). Each pair of
Kikuchi map is formed by collecting a
ll the backscatter signals coming out of different planes on the phosphor
screen. c) With each zone axis identified and labeled, one is able to know the crystal orientation of the
scanned grain. This is extremely useful in the predication of slip systems
during plastic deformation.
(Amended from [107])
21
distributions, grain
-
scale misorientations, size and phase variations, and elastic strain
distribution across a bulk sample.
A well
-
prepared sample is generally
needed, since the highly topographical surface will
leave residual deformation at the surface that leads to local strains and blurs the Kikuchi lines
for a precise orientation detection [111]. It should also be realized that the inelastic electron
intera
ction volume is strongly influenced by the sample tilt. This means, at a high tilt of 70
°
for
EBSD, the spatial resolution along the tilt axis is usually better than perpendicular to this axis.
In a large area EBSD scans, both the top and bottom part of
a sample will be out of focus if the
center is in well focus. Nevertheless, modern high
-
speed EBSD provides spatial resolution from
30 to 100 nanometers [107, 108, 111] with good angular resolutions between 0.5
o
and 2.0
o
, and
orientation precision of 0.5
o
[107, 112]. Several factors simultaneously affect the performance
of the EBSD. The atomic number of material, the geometry of mounting, the probe current, the
accelerating voltage, and clarity of the pattern can strongly affect the spatial resolution; w
hile
the scanning speed and the calibration of the pattern center will both affect the angular
resolution [113
-
117].
With the development of the high
-
resolution EBSD technique (HR
-
EBSD) [118
-
123], the
angular precision has increased to 0.01
o
. This
technique can resolve as low as 10
-
4
elastic strain
across a deformed area by comparing the relative distortion of the pattern collected from an
area to the reference pattern from a presumably strain
-
free area. However, this technique still
needs refineme
nt to improve the resolution and the speed for data treatment [122].
1.2.2
Transmission electron microscopy (TEM)
To understand how a crystal is deformed and to evaluate slip transfer, a method is
22
usually needed to identify the deformation slip (and twinning)
within and between polycrystal
patches. As is mentioned in section 1.1.4, transmission electron microscopy (TEM) has been
widely used for the characterization of slip activity. This technique uses high voltage electrons
(generally 100 ~ 400 keV) that
can penetrate through the sample. The sample is oriented so
diffraction patterns and Kikuchi bands that represent the lattice parameters on the back focal
p
lane below the sample. Thus, the lattice distortion around the defects will be resolved due to
contrast variations from the defect
-
free background. Based on this contrast mechanism,
dislocations can be visualized, with their Burgers vectors identified through the
g
b
= 0 and
g
Figure 7 Sketched mechanism of
g
b
= 0 and
g
b
x
u
= 0 invisibility criteria in the determination of
dislocation Burgers vector. The dislocations will go completely out of contrast or show low contrast when the
dislocation plane lies parallel to the channeling direction, because the it is where almost al
l electrons
diffracting between planes in the same way, leaving no intensity differences between distorted region and
perfect lattice. On the contrary, larger value of
g
b
suggests more intensity variation around the distorted
region, revealing dislocat
ions in better contrast.
(Amended from [124])
23
b
x
u
= 0 invisibility criteria as shown in
Figure 7
(where
g
is the channeling/diffraction vector,
b
is t
he Burgers vector, and
u
is the line direction) [124
-
126]. The dislocation line directions,
the dislocation types (edge or screw), and the slip planes can also be identified by tilt
-
and
-
rotate operations. In addition to the contrast analysis for defect i
maging, with continuous
advancement of electron sources and the special resolution, latest TEM allows the study of
grain boundary configuration and the distortions of atomic arrangement due to dislocation
inductions within the vicinity of the boundaries at
approximately atomic level [127
-
129].
Despite the high resolution and the capability of doing in
-
situ slip transfer experiments, TEM
also suffers a series of limitations [130
-
134], one of which is the requirement of thin foils. Thin
foil sample preparat
ion can be difficult and time consuming, but may also result in artifacts
during improper preparation. Another limitation of TEM is the observation volume, making it
difficult to collect appropriate levels of information for statistical analysis.
1.2.3
Elect
ron channeling contrast imaging (ECCI)
With the advancement of scanning electron microscope, other surface characterization
techniques [135
-
138] have been introduced, such as the electron channeling contrast imaging
(ECCI) [137
-
141]. Among those technique
s, ECCI is particularly strong at the identifications of
near surface dislocation Burgers vectors and line directions [142
-
145], and thus serves as
competitive approach to the TEM. This technique can resolve dislocation image peak widths as
small as 15 n
m (comparable to bright field TEM) and is able to capture the dislocations
distributed within 100 nm of the surface. ECCI is a non
-
destructive technique and a similar
contrast analysis as TEM, but ECCI is collecting signals from backscattered electrons ra
ther than
the electrons penetrating a TEM thin foil.
24
Figure 8a) An example of ECPs collected from a p
-
type boron doped synthetic diamond single crystal with close
to [110] crystal orientation in low mag BSE mode (~ 20x).
The band in the upper left co
rner with bright and
dark contrast is one of the Kikuchi band formed during incident beam sweeping the sample. It disappeared in
larger mag.
. b) An example of EBSPs collected
among one of the g
rains from a commercially pure titanium sample
in
this research (sample 2).
The i
mage
wa
s
taken at a working distance of 24
mm, a 30 kV accelerating voltage and a 184 µA probe current, with sample
tilted at 70
o
. The edges of Kikuchi bands are
significantly sharper than that of ECPs. c) An example of SACPs
collected from the same target with b) using the same voltage and current, but the working distance is around
9mm within 10
o
tilt. SACPs provide much accurate information where the closest z
one axis the crystal is
orientated. With more sharp edges on the channeling bands and higher special resolution, SACPs fits ECCI
analysis more than the other two options.
25
-
analysis,ECCI
requires the sample to be oriented to specific channeling conditions in order to maximize
contrast and facilitate defect analysis. The orientations with respect to the incoming electron
beam can be established using crystallographic orienta
tion information from either low
-
mag
electron channeling patterns (ECPs) or higher magnification selected area channeling patterns
(SACPs), with example patterns shown in
Figure 8
[146]. EBSD can also be used to facilitate
ECCI by inferring the necessary
tilts and rotations to achieve proper two
-
beam channeling
conditions[111, 142]. ECPs (
Figure 8a
) are typically formed at low magnification in a single
crystal or a grain with a large size. Such patterns were often used when the ECCI technique was
first e
stablished. The mechanism for the ECPs formation is sketched in
Figure 9
[146]. While
the electron beam will strike the sample parallel to the optic axis in the center of a scan, as the
Figure 9 Schematic mechanism on the formation of ECPs (Amended from
[146]), with the incident beam
the diffraction contrast at the surface.
26
beam is scanning across a sample, the electron beam trajectory will
vary. At low magnifications
this variation in trajectory angle will be maximized. As these trajectories vary, the electron
beam strikes the lattice planes at different angles, and the Bragg diffraction (channeling)
behavior changes. Subsequently, the b
ackscattered electron yield varies as the beam is
moving, forming a pattern of lines, known as an electron channeling pattern, indicative of the
Figure 8a
in
thus a strong BSE signal is achieved. The edge of the band indicates the lattice planes are
detected at the screen. Comparing the qualities of the patterns in
Figure 8a
-
c
, it appears that
the ECPs are blurry and show worse contras
t, which is not good for a precise establishment of
channeling condition [147] . In theory, ECPs can only provide crystal orientations up to 1
o
, thus
are not very suitable for accurate dislocation related studies since it is hard to precisely tilt the
samp
le to an exact channeling condition. Additionally, ECPs are limited to large
-
grain samples
or single crystals, which are not readily applicable for heterogeneous deformation studies due
to the need for large numbers of grains.
The EBSPs technique (
Figur
e 8b
), with a precision accuracy around 0.5~2.0
o
, has
replaced ECPs in most applications for the determination of crystal orientation. By calculating
the rotation and tilt angles needed to reach the edge of a specific channeling band, it is possible
to es
tablish channeling condition for ECCI analysis based on the EBSP
-
determined crystal
orientation [149]. However, this approach is not intuitive, and the precise establishment of a
27
channeling condition is challenged by the uncertainty induced during the sta
ge movement from
a high
-
tilt EBSD orientation to a low
-
tilt ECCI condition
5
. This makes the fine adjustment of the
impossible, since this fine adjustment typically ne
eds angular accuracy within 0.1
o
, which is
beyond the capability of EBSD. Although this problem can be partially resolved by doing ECCI
using a forward
-
scatter electron (FSE) detector with a similar high
-
tilt setup [149, 150], this
technique also suffers
issues similar to EBSD, with image/diffraction contrast shadowed by the
severe topography and variation of focus across the tilted area [140]. High tilt ECCI also suffers
from image foreshortening.
The SACPs (
Figure 8c
), acquired by electron beam rocking
about a point close to the
surface rather than sweeping across the sample, overcome the limitations of ECPs and are thus
able to be used on small grains (20
µ
m). The advantages of SACPs are: 1. the SACPs technique
have a smaller angular range with a bett
er spatial resolution. 2. The SACPs have angular
accuracy with respect to the beam trajectory within 0.1
o
, which allows the precise
establishment of the channeling condition
g
and the deviation parameter
s
for the
enhancement of dislocation contrast. With a precise calibration of the beam shift on a
crossbeam field emission gun (FEG) SEM equipped with a Gemini column, a high
-
resolution
SACP can be established with a spatial resolution of 500 nm, allowing t
he capability to perform
a quantitative ECCI analysis [147, 148].
As sketched in
Figure 10a
-
c
5
The pattern center of EBSP is a chronic proble
m in HR
-
EBSD that still needs improvement since many factors such
as accelerating voltage, working distance, etc. can affect the position of the center. Without knowing the exact
pattern center, rotation & tilt angles calculated based on EBSD is unreliabl
e.
28
trajectory changes when rocking around the focused point. The diffraction pattern around this
point is thus created since the BSE yield changes with the rocking angle. Once a ce
rtain SACP is
achieved, it is possible to set up a channeling condition based on the lattice orientation of the
crystal. Any near
-
surface stacking fault and line defects can thus be resolved since the lattice
distortion changes the diffraction interaction
of electrons in defect
-
free lattice, providing a
Figure 10a
-
c) Mechanism of the formation of SACPs. As beam trajectory changes or beam rocki
ng around a
certain point, the lattice channel become open and close with respect to the directions of the incoming
electrons, providing different yield of backscatter electron
. The signal profile is collected and create a SCAPs
on the detector. For a
channeling condition that allows the most electrons channeling into the perfect crystal
and leave an overall dark background, lattice distortion around a dislocation will make more backscatter
electrons collected by the BSE detector. Dislocations will be
resolved. d) An example of dislocations (bright)
from the dark background.
29
different BSE contrast. A typical example is shown in
Figure 10d
, where, at a specific
channeling condition, the perfect crystal lattice allows most of the electrons to channel into the
material, resulting
in an overall dark background. Because the near
-
surface dislocations distort
the perfect lattice, the scattering behavior between the incident electrons and the distorted
lattice is different than that in the perfect crystal. With more backscatter elect
rons collected by
the BSE detector around the dislocations, dislocations appear as brighter dots or lines
depending on their orientation with respect to the surface. The mechanism responsible for the
bright
-
dark dislocations contrast is shown in
Figure 11
[141, 152, 153].
Figure 11
(left) shows
Figure 11
left) is exactly at one of the channeling bands on a perfect
lattice, with the deviation parameter
s
= 0. Due to the lattice distortion from the dislocation,
the cha
nneling planes are deviated from the exact Bragg condition, resulting in a different
backscatter signal yield from the background yield level (
Figure 11
right).
Once a specific channeling conditions with a proper channeling contrast have been
Figure 11 An example of the change of channeling contrast with respect to the deviation parameter
s
, which s =
0 indicates the optical axis is exactly at the edge of the channeling band. With optical axis move into or away
from s = 0, signal intensity will change dramatically. As the lattice is no longer aligned symmetrically towards
the incident beam due to dis
location distortion, contrast will occur with bright/dark contrast around a
dislocation compared to the overall grey background. (Amended from [152])
30
established
, dislocation identification can be achieved through ECCI
g
b
= 0 and
g
b
x
u
= 0
contrast analysis [140, 150, 154]. It should be noted that since there is always elastic relaxation
of dislocation core at the free surface, there are situations that di
slocations do not fully
disappear after adjusting the deviation parameter
s
the surface. One should also realize that because the working distance is around 10 mm for
ECCI analysis at 30 kV, the tilt angle
for a larger samples is often limited to about 20
o
, which
can limit the ability to carry out contrast analysis
6
. Thus, it is not always possible to obtain all
the channeling conditions necessary to achieve
g
b
= 0 and
g
b
x
u
= 0, nor is it always
p
ossible to identify line directions by traveling between major zone axes following a channeling
band (i. e. the sample need to tilt 35.16
o
to travel from [110] to [111] zone axis, procedures can
be found in
Appendix V
). However, it is easier to identify t
he inclination direction of dislocation
as well as its slip plane in ECCI, since there is only one free surface for the scanned sample.
1.2.4
Other surface plastic evolution analysis techniques
A number of other techniques that are capable of providing informa
tion on the
evolution of heterogeneous deformation has been developed, such as the digital image
correlation (DIC) and the atomic force microscopy (AFM). DIC was first experimentally applied
by Sutton et al. [155] to the full
-
field (2
-
d) measurement of th
e displacements during
mechanical testing. With continuous improvements in computing technique and imaging
qualities [156
-
158], this technique is now capable of resolving 1 nm horizontal displacement at
the surface. The mechanism is schematically descri
bed in
Figure 12a
. In this method, the area
6
the mounting stage may collide with the detector, and the dramatic drop of BSE yield at higher tilt angles
depending on the material
31
of interest is covered by nanoparticles and labeled by several fiducial marks. Following
deformation, a strain map can be created since the displacement evolution history between
nanoparticles within the area
by correlating sequential images captured during deformation
with the initial reference image [159
-
163]. AFM, with a vertical precision of 0.1 Å [164, 165], is
able to provide a relative strain map based on the height difference across the probed area
[44
]. The simplified mechanism of AFM is shown in
Figure 12b
. While the probe scanned
across the surface, height difference across the area will oscillate the cantilever, resulting in a
deviation in laser reflection from the tip onto a photodiode. The heig
ht variation (Z) at
different coordinate points (X, Y) on the surface, is then used to construct a topography map.
Both techniques have their own advantages and limitations. For instance, DIC is good for
measuring the in
-
plane displacements, but cannot o
bserve dislocation scale movements, while
Figure 12a) Overall mechanism of DIC. The surface is coated with evenly deposition of nanoparticles, with
fiducial marks.
T
he reference image is the upper right square area labeled with four fiducial marks, with a
reference point P (x,y). During deforma
tion, arbitrary shape change and rotation of this area is reflected by the
As the probe is deviating
from its original position due to surface topography, the absolute
height difference (Z) is recognized by the
laser reflection on the position sensitive detector and recorded upon each position (X) from the starting point.
After scanning the whole area line by line, a 3
-
D topographic map can be created by correlating hei
ght profile
(Z) with the plane profile (X,Y).
32
AFM offers high accuracy for tracing out
-
of
-
plane displacement, but suffers from slow probing
speed and artifacts [166
-
168]). Thus, they usually complement with SEM
-
EBSD [162, 163, 44]
based crystallographic infor
mation or other techniques that can compensate for the limitations
in the study of polycrystal deformation. With the incorporation of different surface analytical
techniques, one is able to perform the slip trace analysis that identifies the slip/twin sys
tems
that may be activated during the deformation [24, 164, 169, 170] based on the morphologies of
the slip traces developed during the deformation.
1.2.5
Problems of the classic trace analysis
The
re have been a number of recent studies that have taken the advantages of surface
analytical techniques (i.e. slip trace analysis [24]) for statistical analysis of dislocation activity
since the identifications of slip systems can be much easier to achieve
with computer
assistance. The mechanism of slip trace analysis is shown in
Figure 13
. With the Euler angle
detected by the EBSD, the crystal orientation can be visualized, the intersection line of a slip
Figure 13 The overall mechanism of the slip trace analysis. T
he hexagonal cell presents the crystal orientation,
and the red line is the intersection line between the slip plane (grey) and the sample surface (blue). With the
profile of each trace (1~12) at the surface, possible slip system can be identified.
33
plane with the sample surface can thus be drawn,
which is referred as the slip trace. With the
profile of all slip traces, it can be used to identify the active slip system by comparing the
observed trace to the calculated ones. Current EBSD based slip trace analysis still have some
problems. For exa
mple, it cannot identify the slip system on basal plane since they show the
same trace at the surface. Moreover, it cannot precisely differentiate slip systems that show
similar slip traces (5 vs 9, 6 vs 12 in
Figure 13
). Additionally, this method only w
orks on the
grains that exhibit straight slip traces since the identification of slip system is solely based on
the observation. Thus, this method is currently blind to cross
-
slip identification (and wavy
traces, which will be discussed in this study). T
his limitation is seldom discussed because
researchers will always select another grain that have easier identified slip systems, or choose
the slip system with the highest Schmid factor among possible alternatives. Although the slip
trace analysis is mor
e precise in the identification of slip systems with similar slip traces with
the help of AFM [168], and DIC [162, 163], current method is still not perfect, especially for
wavy traces. Thus, to study the strain accommodation simply relying on the slip tr
ace analysis
is dangerous, since different slip interactions may indicate different strain accommodation
mechanisms during the deformation.
1.2.6
Free surfacing biasing and limitation of surface
-
based analysis
Despite the limitation of the current slip trace ana
lysis, the free surface may also bias
the slip activation and slip transfer events on the surface. For example, current surface
-
based
analysis, such as AFM and DIC, is not sensitive to the slip systems that do not contribute to the
topography change at th
e observed surface. This means there may be some dislocations that
are not correctly identified by the slip trace analysis. This may be a severe issue near the grain
34
boundary, because dislocations from other slip systems may actually play more important
role
in the slip transfer, but are not detected by the slip trace analysis [171]. This ignorance will lead
to improper/incomplete understanding of strain accommodation during the heterogeneous
plastic deformation.
One other limitation to the surface
-
bas
ed analysis (AFM, DIC, EBSD) in the study of slip
accommodation at grain boundary is illustrate in Figure 5. It appears that two slip systems
interact differently at the grain boundary at the sample surface (meet at the same point on the
grain boundary or
not). No matter which type of interaction, evaluation of slip transfer events
is solely based on surface observations. Because it is hard to reveal the geometry of the slip
system and the grain boundary [172
-
174] from surface observation
7
, the geometrically
boundary plane orientations [75, 97, 98]. Additionally, due to not knowing the local
accommodation mechanisms at the grain boundary plane
(especially in the area between the
divergent slip planes below the surface), it is risky to directly use the slip transfer criteria in the
real
-
life deformation.
1.2.7
The advantage of ECCI over other techniques
To solve the problems in current surface
-
based a
nalysis, ECCI is thus needed. One major
advantage of ECCI over other surficial techniques (DIC, AFM) is the capability to identify slip
planes, slip directions, and the Burgers vectors, which is critical in the plasticity study [143,
144]. Additionally,
-
7
Unless using destructive FIB milling on the grain boundary area [172
-
174]. However, FIB milling may lose
informatio
n on slip interactions at the milled area.
35
geometry of a grain boundary plane and the slip planes by the correlation of images taken at
different depths [96, 174]. On the other hand, although not comparable with the darkf
ield TEM
that is able to resolve small dislocation width, ECCI is none
-
destructive and thus is suitable for
continuous deformation study of bulk material [140, 141]. As an SEM
-
based technique, a
broader field of view of ECCI offers the deformation informa
tion from the macroscopic level to
the dislocation level. This technique links the macroscopic and the microscopic world, which is
good for both detailed mechanism study and statistical analysis.
1.3
The objective of this research and design plan
This
Ph.D. thesis aims to solve several open questions about how plastic strain is
accommodated within grains and across grain boundaries during the plastic deformation. The
answers to these questions will provide better guidance for the establishment of a rel
iable
plastic model in the future.
ECCI will be used to study the dislocation slip evolution during heterogeneous plastic
deformation, with a particular focus on slip band/grain boundary interactions. This objective
will be carried out in a number of step
s: First, a robust comparison of a number of approaches
for characterizing heterogeneous deformation will be carried out. This will include how these
various techniques give consistent and/or complementary information. Second, how grain
boundary strain
accommodation is achieved between the interacting slip systems in order to
maintain grain boundary integrity. These studies will examine this behavior across 3
-
D volumes
by carrying out ECCI studies at different depth from the surface. It will be shown t
hat this
approach allows a more robust assessment of the parameters that affect the accommodation
behavior than is facilitated by surface studies alone. Finally, it will be shown that ECCI
36
facilitates the assessment of the sequence of slip activity across
patches of multiple grains,
facilitating a better understanding of the development of heterogeneous deformation.
In order to address the first objective, a surface analytical experiment was performed on
a Ti
-
7
-
Al tensile sample to
facilitate the comparison of ECCI with AFM, HR
-
EBSD, and DIC. The
highlight is to emphasize the convenience of using ECCI in the identification of slip systems. An
additional consequence of this study is to introduce a method of removing DIC patterning
without damaging the surface, facilitating further EBSD and ECCI analysis. To fulfill the second
-
1.5% plastic
strain by four
-
point
-
bending. With controlled electropolishing tech
niques, comparison of
images of slip bands at and below surface reveal how the free surface is biases the slip
-
slip planes and grain boundary planes available from
the images at and below the surface,
several slip transfer parameters have been used to evaluate slip transmission at the grain
boundary. After the identification of the propagation direction of dislocation following slip
bands within series of neighbori
ng grains, it is possible to estimate the direction of deformation
flow traveling within the grain patches, facilitating the determination of deformation
sequences, and locate the grain boundary where the plastic strain was not sufficiently resolved.
37
2.
Experimental Procedures
2.1
Samples preparation
Sample 1 was a Ti
-
7Al
-
which had already been cut by electron discharge machining (EDM) into a 42 x 8.2 x 2.2 mm
tensile bar with a 10 mm gauge length. The dimensions of sample 1 are shown in
Figure 14a
.
Samples 2
and 3 were EDM sectioned into two 25 x 3 x 2.5 mm bars, as shown in
Figure 14b
,
from the
-
titanium provided by Dr. Christopher Cowen (formerly at National Energy
Technology Laboratory). The sectioned samples experienced several grinding steps using silic
on
carbide (SiC) grinding paper from 400, 600, 1200, down to 4000 grit using a polishing wheel at a
speed of 200rpm. Final polishing of the samples was accomplished on a Struers MD
-
Chem
polishing cloth at 300rpm with the mixture of 5:1 volumetric ratio of
0.05 µm colloidal silica
suspension (Struers OP
-
S) and 30% hydrogen peroxide solution for 30 minutes.
All three
samples
were electropolished with a polishing cell, as sketched by
Figure 15
, using different
electrolytes and different parameters [175
-
179].
Samples 1 and 2 were electropolished in a
Figure 14a) The dimension of the Ti
-
7Al dog
-
bone tensile sample 1. b) The dimension of the CP Ti bending
sample 2 & 3.
38
solution that
contained 30 ml perchloric acid, 200 ml butanol, and 300ml methanol, using an
applied voltage of 38 V at
-
35
o
C. Sample 3 was electropolished using 24 V at
-
30
o
C using a
solution containing 10 wt% magnesium perchlo
rate and 90 wt% methanol. Detail of the
electropolishing mechanisms, parameters, and a comparison between the two methods are
recorded in
Appendix A
.
Before deformation, grain orientations of the samples were characterized using EBSD,
with grain boundar
y and surface conditions (after electropolishing) checked by general
secondary electron (SE) and backscattered electron (BSE) imaging mode using a Tescan Mira III
FEG
-
SEM equipped with an EDAX
-
TSL orientation imaging system. EBSD was performed using a
30
kV accelerating voltage with a 148 µA probe emission current, a 20.0 nm spot size, and an 18
mm working distance with the samples tilted to 70
o
. The instrument parameters for SE/BSE
(and later ECCI) observations were the same as used for EBSD, but the wor
king distance was
Figure 15 Sketch of the electropolishing stage. Based on what type of electrolyte is used, the voltage,
temperature while electropolishing, the distance between cathode (stainless steel) and sample (anode), and
the agitating speed of the stir bar will be
different and recorded in
A
ppendix
A
.
39
around 8
-
10 mm (with a maximum stage tilt of 20
o
for ECCI analysis). If not specifically
mentioned, all images and analysis (including AFM) during each experimental stage were taken
at consistent conditions, with the sample placed in the
same orientation on stage.
2.2
Samples deformation
2.2.1
Deformation of Ti
-
7Al sample and uncoating
After mapping the crystal orientation distribution of sample 1 using EBSD, it was sent
ith densely
deposited gold nanoparticles (AuNP) through surface condensation reactions [162, 163]. It was
then plasticly deformed to ~3% tensile strain (the coordinate system for deformation,
observation, and analysis remained consistent and is sketched i
n
Figure 16a1
), with a full
-
field
displacement of particles and strain development monitored through DIC at different strain
levels.
After receiving the sample back from the Daly group, the sample 1 was soaked for a
total of 4 hours at 30
o
C in a solutio
n of tetra
-
n
-
butylammonium fluoride (TBAF) [180
-
184],
chloroform, and ethylene glycol with a weight ratio of 10: 1: 1 respectively. During this
uncoating process, sample 1 was taken out every 1 hour and cleaned with soap water using
sonication for 5
-
10 mi
nutes. Final cleaning was accomplished by dipping the sample into
dishwashing soap, wiping off the soap with cotton, flashing with ethanol
-
water
-
ethanol, and air
drying. The overall uncoating approach was successful, with no AuNPs left on the surface,
re
sulting in a smooth surface and sharp SACPs. The detailed coating removal procedure is
recorded in
Appendix B
.
40
2.2.2
Deformation of samples 2 and 3
Samples 2 and 3 were plasticly deformed in a four
-
point
-
bending stage to around 1.5%
and 1% surface tensile str
ain, respectively, with tensile strain measured as outlined in
Appendix
C
(the coordinate systems for samples 2 and 3 are shown in
Figures 16b1&c1
).
2.3
Samples analysis
All deformed samples (1
-
3) were placed on the SEM stage for the observation of slip
tr
aces developed during the deformation, with post
-
deformation EBSD data collected to update
the crystal orientations from deformation. Combined with the information from the slip traces
and the corresponding EBSD orientation profiles, surficial slip trace
analysis was performed
using an in
-
house developed MATLAB code [24], where the crystal orientation, the slip plane
that may leave the slip trace on the surface, and the potential Burgers vector were input, with
the global Schmid factor M calculated based o
n the Euler angle of the crystal and the geometry
of potential slip systems relative to the surface tensile direction. Based on the geometry of slip
planes and the crystal orientation of a grain and its neighboring grains, the alignment of slip
systems ac
[95].
2.3.1
ECCI analysis on sample 1 and electropolished samples 2&3
ECCI analysis was carried out directly on the plasticly deformed Ti
-
7Al sample 1, both
within grains and n
ear grain boundaries, facilitating the identification of dislocations (the
alignment of SEM for ECCI is in
Appendix D
). The Burgers vectors were identified through ECCI
g
b
= 0 and
g
b
x
u
= 0 contrast analysis [140, 150, 154], and the slip planes and line
directions were roughly estimated by tilting along one of the Kikuchi bands, with detailed
41
procedure presented in
Appendix E
. Subsequent to this characterization, samples 2 and 3 were
fu
rther electropolished
using the same electropolishing conditions outlined in
Appendix A
,
which eliminated all surface topography. The depth of material removed was determined by
applied current and electropolishing time and was directly measured by the Vi
ckers indent. The
resulting materials removal was approximately 5
µ
m from sample 2 and
approximately 20 µ
m
being from sample 3.
EBSD was again carried out on these samples following this surface removal. ECCI was
then performed on samples 2 and 3 in order to identify dislocations, dislocation propagation
investigation behavior, and dislocation interactions at grain boundaries.
2.3.2
AF
M analysis on sample 1
The topography developed due to slip band development during the deformation was
measured using a
VEECO Dimension 3100 AFM operating in tapping mode
at a speed of 10
µ
m/min for every 40 x 40
µ
m
2
area. The data from AFM was processed
using the Gwyddion
software package
8
, with the background surface normalized (polynomial 3) and the regions
having the lowest height were automatically assigned as zero during the analysis.
2.3.3
HR
-
EBSD analysis on sample 1
HR
-
EBSD was performed on the area
s where ECCI was performed on sample 1, using a
sample tilt of 70
o
, a working distance of 20 mm, and a 20.0 nm spot size. Each high
-
resolution
pattern for the cross
-
correlation was taken at an exposure time of 0.1 s with a 480 x 480
-
pixel
resolution and t
he EBSD patterns were saved. As indicated by Dunlap et al. [137] and Ruggles
et al. [138], the step size will affect the GND density distribution determinations; the GND
8
Available free at
http://gwyddion.net/
42
analysis in this research was based on 200 nm effective step size, a parameter that
can resolve
dislocations as best as possible [136, 137]. This facilitated a semi
-
quantitative comparison
between the GND measurements and the ECCI and AFM data.
43
3.
Results and Discussions
The results presented here are primarily in the form of a large number of ECC images
that allow the determination of crystallographic details, primarily Burgers vectors, dislocation
line directions, and slip band morphologies.
These are related to underly
ing crystallographically
dependent parameters based on EBSD analysis, including global Schmid factors, the angle
between slip plane intersections in grain boundaries,
, and the resulting compatibility factors
These interdependent factors vary
from case to case.
Thus, rather than
presenting the results of various experiments in isolation, the author believes the best way to
descriptively convey the research is to combine the results and discussion, presenting
combined results for various case
s in order to tell complete stories without leaving unanswered
questions, rather than laying out results fragmentally.
Nevertheless,
sections 3.1&3.2
will
outline the generalized approaches and observations used for carrying out the specific studies.
Sec
tions 3.3
-
3.5
will outline the advantages of ECCI technique over other technique in the
heterogeneous plastic deformation study, especially in terms of slip accommodation activities
within grain interior and at grain boundary area.
3.1
The overall status of
as deformed samples 1
-
3
The deformation of samples 1
-
3 at their respective strain levels (sample 1 at 3%, sample
2 at 1.5%, and sample 3 at 1%) was dominated
by heterogeneous slip systems.
No deformation
twins were observed within the targeted areas, whi
ch was confirmed by the SE images and
EBSD orientation map of these areas, as shown in
Figure 16
. Although comparisons before and
after deformation are not shown, the crystal orientations were not distinguishably changed
with deformation.
As shown in
Fig
ure 16a1
, the collections of lines lying at the surface of
44
Figure 16a1) The SE image from the center of the 3% tensile strained Ti
-
7Al sample 1, with the tensile
direction along A2 axis. Almost all grains were deformed, with some grains having more than one
type of slip traces. Surface and grain boundary elevati
on could be indicated by the brighter contrast
against the dark grains. a2) The corresponding EBSD data in the red region of a1, which was
collected under the same coordinate system as a1. Despite the noises due to higher deformation
strain that disrupts
the diffraction (with confident index only 0.65), color gradient within grains can
be clearly seen, especially where slip bands were densely packed. Slip traces appear mostly straight,
while some curved traces were found near grain boundaries or in the g
rains which were heavily
deformed.
b1) An example of one of the grain boundaries between two neighboring grains in the
center area of sample 2 after 1.5% deformation on four
-
point
-
bending stage. Again, most slip traces
appear straight, while the bright a
nd dark contrast on the slip traces may indicate they may not share
the same Burgers vector. Despite some primary slip traces that were fully propagated across the
grains, some of the slip traces disappear as they propagated out of the grain boundary. b2
) The
corresponding EBSD map of the red area of b1, indicating the formation of slip bands was not
significant enough to affect the quality of the EBSD, with a high confident index of 0.81, and the
topography developed during deformation was not big enough
to be shown in the EBSD compared to
a2.
45
Figure 16
(cont
d)
c1) SE image of a patch of 4 consecutive grains in the center of sample 3, which
was deformed in the same manner with sample 2 to 1% plastic strain. Slip traces in grains 2
-
4 appear
s
traight while traces in grain 1 are wavy. Some traces disappear near the grain boundary in grain 2,
which is indicated by the white arrow. Traces in grains 1 and 2 meet at the same point on the grain
boundary, so are the traces in grains 3 and 4, while t
races in grains 2 and 3 do not meet at the same
point on the grain boundary. c2) EBSD map of the grain patch in c1. The deformation is not large
enough for EBSD to recognize the formation of slip bands at 1% strain level. No crystal rotation is
detected
. It should be noted that all the prism cells in a2, b2, and c2 indicate the crystal orientations
of the grains they were located, with the black dots indicating the optical axis.
46
sample 1 indicate the traces of the slip systems [185]. It can be s
een that the slip traces have
varying directions, suggesting the grains were deformed in different directions due to shear in
different directions.
With the exception of a few grains, almost all grains in the red region show
apparent slip traces, with hal
f of the grains showing more than one type of slip traces.
The
overall grain size in sample 1 was somewhat smaller than that in samples 2&3, and it appears
that the distance between slip traces in sample 1 were smaller
. It is also noticeable that some
gr
ains and grain boundaries show a significant contrast difference than their surrounding
environment as a result of local lattice rotation due to deformation leading to different electron
channeling contrast. For example, the upper grain boundary and the r
ight region of grain 3
show a much brighter contrast than the rest area in this grain, which indicates there is a large
variation of lattice orientation in that area due to the deformation. The orientation change due
to plastic deformation was also captur
ed by the EBSD orientation map (
Figure 16a2)
. For
example, the densely packed slip traces are seen as fine purple lines in contrast to the overall
pink background in grain 4, suggesting a large local orientation change in those areas. In
contrast, the or
ientation variations of samples 2&3 appeared much straightforward since they
experienced a much lower plastic strain. At 1~1.5% plastic strain, only ~30% of grains in the
observed areas show clear slip traces on the surface. In addition, at this low plas
tic strain, all of
the traces within a deformed grain (
Figure 16b1&c1
) had the same orientation, suggesting that
no other slip systems were activated. It is noticeable that some grains are not fully
recrystallized and have shown interesting diffraction features (grain 5 in
Figure 16c1
), such
grains are rarely observed and
not the focus of this research.
47
3.2
Slip system identifications
3.2.1
EBSD based slip trace analysis
In order to understand the heterogeneous deformation of a p
olycrystal, it is critical to
correctly identify the slip systems that are active during the plastic deformation. Based on the
slip traces developed during the plastic deformation and crystal orientation information from
the EBSD, the active slip systems
can be readily identified with the EBSD based slip trace
analysis [24]. An example of slip trace analysis is shown in
Figure 17
.
Figure 17a
shows grain 2
in sample 1 after 3% plastic deformation [185]. Two types of straight slip traces are observed in
g
rain 2, with one lying from the upper left to the bottom right (#1) and the other from the
upper right to the bottom left (#2). The hexagonal cells on the right represent the lattice
Figure 17a) BSE image of grain 2 in sample 1. Two straight and planar slip traces can be clearly observed, with
their trace outlined as black lines. The hexagonal cells on the right indicate the orientation of this grain. The
shaded plane in each cell i
ndicates the slip plane, the blue line indicates the slip direction, and the red dashed
line is the plane trace on the sample surface. For a potential slip system active during deformation, the red
dashed line should match the trace on the surface. b) BS
E image of a patch of grains 1~3. Some slip systems
with different Burgers vectors or slip planes may exhibit similar slip traces at the surface, resulting in
uncertainties. The colored dashed lines in each grain represent the possible slip systems that
leave similar slip
traces. The Burgers vectors and slip planes of all possible slip systems are listed on the right. The color of the
slip system is using the same color with the dashed line.
48
orientation of grain 2. The grey shaded planes in the cell represent th
e slip plane, and the blue
lines are the slip directions. The red dashed lines in the cells, referred as the slip (plane) traces,
show the intersection line of the slip plane with the sample surface. If the red dash line is
parallel to the observed trace
s on the sample, then the slip plane that leaves this slip trace on
the surface is identified as the deformation plane (similarly, the colored dashed lines are the
slip traces from the corresponding slip systems in
Figure 17b
). Because there is generally
only
one Burgers vector on each specific plane due to lower symmetry of
-
titanium (with the
exception of the (0001) basal plane, which can have three Burgers vector [11
0], [1
10], and
[
110]), the slip system can thus be identified.
In some cases, slip systems with different Burgers vectors or slip planes may show
similar slip traces, adding some uncertainty in the slip system identifications. In
Figure 17a
,
particularly, (0
10) [2
0] prism slip system and (10
1) [
23] pyramida
l slip
system both show similar traces with the observed trace #1. Similarly, (01
1) [
3] pyramidal
slip system and (10
0) [1
10] prism slip system show similar traces that match the
observed trace #2. Slip trace analysis is also
not capable if identifying a specific plane for wavy
traces, since it is difficult to match curvy surface traces with the calculated ones. An example of
this can be found in grain 1 of sample 3 (
Figure 17b
[96]), due to the curved traces, four
possible sl
ip systems may be active according to the EBSD
-
based slip trace analysis.
Additionally, grain 2 and grain 3 may deform differently and might have different slip
interactions based on how the slip systems are identified. In general, previous studies [43,
74,75, 97, 98] using this EBSD
-
based slip trace analysis have overcome this difficulty by
selecting the slip system with the highest global Schmid factor (in this study this was calculated
49
using an in
-
house MATLAB code), by choosing the slip system with th
e lower CRSS value
9
, or a
combination of these approaches.
The latter approach is statistically reasonable because such
slip systems are more likely to be active. However, this statistical hypothesis is not safe to use
in the study of slip interaction
s at grain boundaries, and it is not viable in extreme conditions (i.
e. slip systems having similar Schmid factor/CRSS ratio). In
Figure 17b
, there are a total of 6
possible slip systems within grains 1
-
3 (uncertainties cannot be eliminated), resulting i
n a more
complicated situation in the evaluation of slip transfer events. Thus, a more reliable method is
needed to facilitate the identification of slip systems. In this thesis, the difficulty is overcome
by identifying the Burgers vector using ECCI.
3.2.2
I
dentification of slip systems using ECCI
To eliminate the uncertainties from surface trace analysis, ECCI was either directly
carried out on the as
-
deformed sample surfaces or after electropolishing to remove some of
the near
-
surface material. The Burgers
vectors of the slip systems can be identified by ECCI
based
g
b
= 0 and
g
b
x
u
= 0 invisibility criteria [140, 150, 154], where
g
is the channeling
vector,
b
is the Burgers vector, and
u
is the dislocation line direction. Once the Burgers vector
b
is id
entified, the slip system that forms a certain slip trace can be identified
10
based on the
combined knowledge of the Burgers vector and the subset of possible slip planes from the slip
trace analysis.
One example of the ECCI identification of dislocation Burgers vectors is given for grain 1
9
-
titanium, prism slip system is more likely to be activated than the pyramidal
due to a lower CRSS
value
. Thus, it is more likely to consider prism s
lip system is the active one
than other candidates, unless the
Schmid factor of prism slip is extremely low and is not possible to be activated.
10
Dislocation line direction
u
can be used to
reveal the inclination of the slip plane, and
thus
eliminate the
situation when a pyramidal and a prism slip system shows a similar trace with the same Burgers vector
.
50
Figure 18a) BSE image of one of the areas of interest after deformation of sample 1. b
-
f) ECC images taken at
different channeling conditions from the red boxed area. The upper left circles are the Kikuchi patterns. Each
black arrow across a Kikuchi ban
d indicates the specific channeling band, and the arrowhead is where the optic
axis is focused. Each channeling condition is identified from the T. O. C. A software. The Burgers vectors of
dislocations are identified by
g
.
b
= 0 and
g
.
b
x
u
= 0 cont
rast analysis, and the slip plane can be revealed
through different tilting and rotating along a certain channeling band (
Appendix
E
). A total of four different
slip systems are identified and labeled by colored arrows, namely: (01
0)[2
0] prism sli
p system (green),
(10
0)[1
10] prism slip system (blue), (10
1)[1
10] pyramidal slip system (purple), and (0001)[11
0]
basal slip system (red). Amended from [185]
51
of sample 1 in
Figure 18
[185]. A series of ECC images of the same area have b
een taken at
different channeling conditions (at different
g
). If the dislocation shows little or no contrast at a
certain
g
, then the Burgers vector can be identified through the ECCI
g
b
= 0 and
g
b
x
u
= 0
contrast analysis. For example,
dislocations with the Burgers vector of [2
0] should show
good contrast at
g
= [
], [
2], [1
00], [11
0] (
Figure 18b
-
e
), and show weak contrast at
g
= [01
1] (
Figure 18f
). In this particular example, a total of three different Burgers vector
were
identified. After combining the Burgers vector and the slip plane traces, four different slip
systems were finally identified, namely: the (01
0)[2
0] prism slip system (green), the
(10
0)[1
10] prism slip system (blue), the (0001)[11
0] basal
slip system (red), and a
small number of (10
1)[1
10] pyramidal slip system (purple). The majority of dislocations
that contributed to the formation of slip bands in grain 1 were the (01
0)[2
0] prism slip
system (green) and the (10
0)[1
10] prism slip system (blue). Additionally, with the Euler
angle acquired from the EBSD software, the global Schmid factor can be calculated based on
the global stress state. Based on the experimental observation, it is clear that the (01
0)[2
0]
pr
ism slip system (#1 slip system in
Figure 17a
, green in
Figure 18
) is more active since the
global Schmid factor (M) is 0.49, while the other prism slip system, (10
0)[1
10] (#2 slip
system in
Figure 17a
, blue in
Figure 18
) is less active because i
t has a much smaller M, 0.23.
The type dislocations on pyramidal and basal planes do not contribute to the deformation
due to the low Schmid factor (0.19 and 0.09 respectively) and higher CRSS value.
Similar approaches were also applied on the electr
opolished samples 2 & 3 (
Appendix
F
), and perfectly confirmed the slip systems active during the plastic deformation. For example,
in sample 3 (
Figure 17b
), the active slip system in grain 1 was dominated by the (1
)[11
0]
52
prism slip system (M=
0.15), with dislocation cross
-
slip on (1
) pyramidal plane
(M=0.23). The active slip system in grain 2 was also identified to be (1
)[11
0] prism slip
system (M=0.41), and the slip system in grain 3 was the (10
0)[1
10] prism slip system
(M=0.
48). With precise identification of slip systems, the study of shear associated with each
slip band during heterogeneous deformation and the investigation of slip/grain boundary
activities is thus reliable.
3.3
The comparisons of surface
-
based techniques
Th
e main purpose for analysis of the deformation of sample 1 was to compare different
surface
-
based technologies in the evaluation of plastic deformation. Relative techniques
involved in this study were ECCI, atomic force microscopy (AFM), and EBSD cross
-
co
rrelation.
Meanwhile, digital image correlation (DIC) data was provided by Dr. Zhe Chen in Professor
data [185, 186].
3.3.1
Local shear distribution across the surfac
e
AFM was used to monitor the topography change due to deformation slip activation
during the plastic deformation of sample 1. Each AFM grid covered a maximum area of 40 x 40
µ
m
2
to maintain considerable accuracy (large area scan will sacrifice accuracy)
11
. An example of
AFM topography map is shown in
Figure 19
.
Figure 19b
is a color
-
scale topography map from
the black boxed area in grain 2 (SE image) as shown in
Figure 19a.
The topography change due
to two different slip systems can be detected from th
e AFM (no research has used AFM to
11
Higher scanning speed will sacrifice the preciseness, so the tiny surface change will not be recorded by the
system.
53
Figure 19a) SE image of the upper right corner of grain 2 in Ti
-
7Al sample 1. b) AFM color
-
scale map from the
black boxed area in figure a). The black line is the AFM line profile showing the topography change acro
ss the
line. The black arrows indicate the edges of slip bands, and the dashed line indicates the undistorted surface
plane and is the basis of the height measurement. c) 3
-
D Greyscale topography map around the line sectioned
area, the surface normal is
calculated based on undistorted surface, H is the step of the slip band. d) A sketch
of the mechanism to calculate the local shear distribution across the slip bands. The Burgers vector
b
and slip
plane normal
n
can be directly achieved from the EBSD data once the slip system is confirmed by ECCI contrast
analysis. Height difference across a slip band H can be directly measured from the AFM line profile. The
distance across a slip band X is 0.3
µ
m in this study
.
(Amended from [185])
54
identify slip systems), which is consistent with the SEM observation (
Figure 18a
). In the AFM
data analysis, the background is subtracted and the lowest point in the map is set to zero
height, so it is easier to find the undistort
ed surface plane that has not experienced
deformation (surface plane normal
e
). Based on the line section profile across the area (
Figure
19b&c
), the height change H (nm) across a certain distance (X
µ
m) due to the activation of each
slip system can be a
cquired. With the crystal orientation achieved from the EBSD, the Burgers
vector
b
and slip plane normal
n
of each identified slip system (by ECCI) can also be calculated
from the in
-
house MATLAB code [24], as outlined earlier. The method for local shear
distribution was established by Yang et al. [44], with the mechanism shown in
Figure 19d
[185].
For a certain slip system that contributes to the formation of a slip band, the number of relative
slip planes N can be calculated from the height change acro
ss the slip band H and the projection
of the Burgers vector
b
[b
x
b
y
b
z
] onto the surface plane normal
e
[e
x
e
y
e
z
]:
N =
The calculation of shear is tile based. With equation (6), the averaged shear contributed
by each slip band
along a certain distance X in each tile can be estimated as
:
=
=
Where
n
[n
x
n
y
n
z
] is
the slip plane normal of the certain slip system. The map of relative
local shear distribution across one area in grain 2 is shown in
Figure 20b
, which reveals the
topography change of this area according to the AFM map (
Figure 20a
). The deformation she
ar
contributed by the
(01
0)[2
0] prism slip system
is from 0.3 to 0.7, and the shear from
the
(10
0)[1
10] prism slip system ranges from 0.08 to 0.15. This is also consistent with
55
the argument that (01
0)[2
0] prism slip system (M = 0.49) i
s more active than the
(10
0)[1
10] prism slip system (M = 0.23) in slip trace analysis and the SEM observation.
Based on a similar mechanism shown in
Figure 19d
, the local shear distribution of slip
systems can also be calculated by
measuring the difference of in
-
plane displacement [185, 186]:
=
Where S
x
y
z]
is the relative displacement that can be calculated from the Burgers
vector and in
-
plane displacement across a slip band:
=
x/b
x
or
=
y/b
y
The resulting shear distribution map calculated by the DIC is shown in
Figure 20c
[185].
The deformation shear across each slip band calculated by the DIC method is similar with the
AFM result. For example, the shear contributed by s
lip band #1 is ranging from 0.26 to 0.48 and
the shear caused by slip band #2 is ~0.16 calculated by the AFM method. Similarly, the DIC
method shows a similar shear amount from slip band #1 (0.25~0.45) and #2 (~0.14). The
difference might come from diffe
rences in the accuracy of AFM and DIC data collection (DIC is
Figure 20a) color
-
scale AFM map from an area in grain 2. b) The heat map of local shear distribution across
individual slip bands. The local shear contributed by the (01
0)[2
0] prism
slip system ranges from 0.3
to 0.7, and the shear caused by the (10
0)[1
10] prism slip system ranges from 0.08 to 0.15. c) The
relative shear distribution map of the same area calculated by the DIC method [185].
56
more sensitive to in
-
plane displacement, while AFM is more focused on out
-
of
-
plane
measurements) or the position variance used in the calculation (the two points at a distance of
X µm across the
slip band are different in AFM and DIC calculation), etc. Nevertheless, the
overall results from the two methods are consistent with each other. Besides this typical
example, comparisons between both methods have been made across several other slip band
s,
all giving reasonable results. This suggests the two methods are reliable and consistent in
revealing the local shear distribution in the plastic deformation.
3.3.2
Dislocation based characterizations of plastic deformation
Different from the AFM and DIC
approaches that directly reveal the shear that cause
deformation, the EBSD cross
-
correlation method is able to estimate the residual elastic shear
after slip band formations by investigating the distribution of geometrically necessary
dislocations (GNDs).
Since the methodology study of parameter selection of GNDs is not the
focus of this study and has been well illustrated by Fullwood et al.[117, 136
-
138], the specific
parameter selection for this approach will not be discussed. In this study a 990 nm st
ep size
was used for the cross
-
correlation calculation in the open source OPEN
-
XY software
12
[187],
which gives a relatively stable GNDs density (990 nm is where GND does not significantly
change with step size) and eliminates potential noises from data co
llection. Since the
dislocations are visible by ECCI, it is possible to compare the reliability of the GNDs to ECCI [137,
titanium is currently unavailable) and
ECCI is shown in
Figure 21
.
Figure 21a
shows ECCI
12
The latest
OPEN
-
XY has added the GNDs calculation of
titanium, however, the split GNDs for
titanium is
currently unavailable. This is the reason for showing only total GNDs, rather than GNDs for each types of
dislocations in this study.
57
Figure 21a) ECC image of the same area with Figure 20. Besides the (01
0)[2
0] (green) and (10
0)[1
10]
(blue) two prism slip systems identified in the slip trace analysis, (0001) [11
0]
(red) basal dislocations
are revealed by the contrast analysis. The prism type dislocations appear to align perfectly along the slip
bands, while the basal dislocations are less uniformly distributed across the observed area. b) The GND
logar
ithm map of the red boxed area of a). The GND map is consistent with ECCI observation, although
individual dislocations cannot be resolved as good as the ECCI observation. c) The ECC image at the grain
boundary of grain 2, with (01
0)[2
0] (green) and (
0001) [11
0] (red) dislocations. Prism dislocations
align close to the slip bands and basal dislocations are more randomly distributed across the surface. d)
The GND map from the red box area of c). The GND on the other side of the grain boundar
y is not available
due to misorientation angle exceeding the threshold from the reference point in grain 2. The GND map is
consistent with ECCI observation.
58
observation of the same area with
Figure 20
, where the shear is calculated by the AFM and DIC.
E
CCI correctly reveals the two prism slip systems, (01
0)[2
0] (green) and (10
0)[1
10]
(blue), that have been identified in the slip trace analysis. The dislocations of the two types are
uniformly aligned on the slip bands. More importantly, ECCI c
ontrast analysis also reveals a
number of (0001) [11
0] basal dislocations (red) that are less uniformly distributed within
the vicinity of the slip bands.
Figure 21b
shows the logarithm map of total GNDs from the red
boxed area in
Figure 21a
. It ap
pears the dislocation density along the slip band (10
14.5~15
) is
more than one magnitude larger than the overall background (10
13~13.5
), with some dislocations
(10
14~14.5
) distributed randomly near between slip bands. The total GND density is consistent
with the ECCI observation (10
14
corresponds to 50 nm between dislocations) that dislocations
are more localized around the slip bands. Similarly, ECCI and GND map at one of the grain
boundary in grain 2 also consistent dislocation distributions on the lef
t side of the grain
boundary (
Figure 21c&d
), although only (01
0)[2
0] (green) prism dislocations and (0001)
[11
0] basal dislocations (red) are detected. Again, the prism dislocations are found
confined in the slip bands, whereas the basal dislocations are less uniformly distributed
between slip bands.
3.3.3
Comparison between the classic EBSD based slip tr
ace analysis and ECCI
It may be noticeable in the discussion of
Figure 21
that EBSD based slip trace analysis is
not able to identify all the slip systems developed during the plastic deformation. Some slip
systems may not develop well defined slip bands
for a number of reasons: if cross slip is easy,
dislocations may not form well defined slip traces or if the slip plane is aligned close to parallel
to the observed surface. Additionally, if an experimentally observed slip trace does not match
59
the theoretical trace, the slip system for that trace cannot be identified. Slip system
identification in grain 1 of sample 1 is an example showing these limitations of current slip trace
a
nalysis (
Figure 22
). Both AFM topography map and the SEM image show two different slip
traces. EBSD
-
based slip trace analysis indicates the slip traces are caused by (
100)[11
0] (red)
and (0
10)[2
0] (blue) prism slip systems (
Figure 22a&b
).
But there is another type of
slip trace observed close to the upper grain boundary of grain 1 (brown), which cannot be
identified, since the slip trace does not match any of the theoretical traces. ECCI contrast
analysis reveals this unidentified slip sy
stem to be (
011)[1
10] pyramidal slip system (light
brown in
Figure 22c
). The slip trace might be distorted by local lattice rotation or the
topography development during the plastic deformation, and thus deviate from the theoretical
traces. Additio
nally, a large number of basal dislocations are detected by ECCI, with the
majority belonging to (0001)[1
10] slip system. Based on the MATLAB code calculation, at this
Figu
re 22a) AFM color
-
scale topography map at the upper right corner of the grain 1 in sample 1. Two
different deformation shear system can be observed by the AFM. b) ECC image of the same area, two slip
systems were identified through the EBSD
-
based slip tr
ace analysis, which are the (0
10)[2
0] prism slip
system (blue) and the (
100)[11
0] prism slip system (red). The slip trace marked in light brown cannot be
identified by the slip trace analysis. c) High magnification ECC image of the red box a
rea in b). Contrast
analysis shows several additional dislocations, including (10
1)[1
10] and (
011)[1
10] two pyramidal slip
systems and a mixture of basal dislocations with majority belong to (0001)[1
10] slip system. The
unknown branched sli
p trace is caused by the (
011)[1
10] pyramidal slip system.
60
crystal orientation, the basal plane is close to parallel to the sample surface, t
hus no slip band
can be easily developed that can be identified in the slip trace analysis. Furthermore, the basal
and (10
1)[1
10] pyramidal dislocations are also observed (brown), but the density is
small compared to other dislocation types. Th
us, these dislocations have not developed slip
bands on the surface. Similar situations where more dislocation types can be identified by ECCI
than the slip trace analysis were also observed in other grains of sample 1 (
Appendix G
).
3.3.4
Advantages and disad
vantages of EBSD, DIC/AFM, and ECCI
A summary of slip system identification using EBSD based slip
trace analysis, ECCI, and
AFM/DIC [185] is shown in Table 1. Classic EBSD
-
based slip trace analysis is not good at
differentiating slip systems that have similar slip traces
13
. With the help of the Schmid f
actor/CRSS ratio, some pyramidal slip
systems can be eliminated (i.e. (2
2)[
113] in
13
This is partially reso
lved by the EBSD cross
-
correlation that use lattice rotation across a slip band to determine
the possible slip system [189]. This method is not the classic slip trace analysis.
Table 1
.
Slip systems identified by EBSD slip trace analysis, ECCI, and AFM/DIC
Grain 1
Grain 2
Grain 3
Grain 5
EBSD
-
slip trace
analysis
(0
10)[2
0]
(0
11)[
]
(
100)[11
0]
(2
2)[
113]
(0
10)[2
0]
(10
1)[
23]
(01
1)[
3]
(10
0)[1
10]
(0
10)[2
0]
(
01
1)[
23]
(10
0)[
0]
(10
1)[
23]
(1
00)[11
0]
(2
2)[
113]
ECCI
(0
10)[2
0]
(
100)[11
0]
(
011)[1
10]
(0001)[1
10]
(10
1)[1
10]
(01
0)[2
0]
(10
0)[1
10]
(10
1)[1
10]
(0001)[11
0]
(0
10)[2
0]
(10
0)[
0]
(0
11)[2
0]
(
100)[11
0]
(1
00)[11
0]
(01
0)[2
0]
AFM
and
DIC [18
5
]
(0
10)[2
0]
(
100)[11
0]
(01
0)[2
0]
(10
0)[1
10]
N. A.
N. A.
Note: Colored fonts
indicate different slip traces, black fonts are dislocations that are only observed in ECCI.
Shadowed slip systems are not likely to be activated because of the lower Schmid factor/CRSS ratio.
61
grain 1&5), however, this estimation is not always safe (such as grain 2&3). On the other hand,
AFM and DIC [185, 186] are more precise in the slip system identification, since the physical
displacement
14
across a slip band indicates shear direction (i.e. Burgers vector), which
supplements the trace analysis. Unfortunately, these methods are not good for traces that do
not match the theoretical generated traces (grain 1 in
Figure 22
) or the slip systems t
hat do not
show apparent slip traces (i.e. (0001) dislocations in grain 1&2). On the contrary, ECCI is
especially good for the identification for slip systems since this identification is based on
dislocation contrast rather than topography/displacement c
hange. Compared to the classic slip
factor (a factor defines the deviation of an observed trace from a theorical one. Zero deviation
means a perfect match),
of an observed trace to the theoretical one dislocation contrast
different channeling conditions
g
are taken, there is almost no uncertainty in the identification
of
dislocations regardless of the morphology of the slip traces. Thus, ECCI is able to identify
every dislocation slip activation during the deformation that is ignored by the classic slip trace
analysis (ECCI row in
Table 1
). This is
particularly
important
when studying the strain
roles in these events.
Slip system identification techniques, no matter ECCI or DIC/AFM, need the crystal
orientation information achieved from the EBSD. Although not shown in this study, most
14
Currently, this method is only available on DIC, however, since the mechanis
m are similar, it is possible to
develop a program for AFM
.
62
recent EBSD cross
-
correlation is able to more precisely identify slip systems by rela
ting the
GND
-
induced rotation gradient across the slip bands to the slip system identifications [185,
188]. Yet EBSD cross
-
correlation is time consuming and needs extremely large data volume (30
Gb for an area shown in
Figure 21
at a step size of 200 nm,
5 hours per scan).
DIC and AFM are generally useful tools for studying plasticity. DIC is more accurate in
monitoring the in
-
plane displacement, while AFM is good for determining out
-
of
-
plane
topography change. With the crystal orientation information
, both methods give reliable
information of the local shear distribution across slip bands. Particularly, by correlating the
relative displacement across a slip band to a theoretical slip system, DIC is more precise in the
identification of the slip plane
, thus improve the reliability of the slip trace analysis. However,
this improved slip trace analysis is still blind to the curvy traces that will develop from cross
-
slip,
traces that deviate from theoretical traces, and slip systems that do not form a sl
ip trace. This
is a chronic issue that all the slip trace analysis techniques suffer from. Although this issue may
not be significant in the study of overall plasticity development in the bulk sample, this issue
may lead to severe problems in slip transf
er studies, since such irregular or hidden slip traces
usually exist near the grain boundary. Incorrect identification of these slip systems may bias
the understanding of slip interactions at the grain boundary.
Compared to the techniques discussed above,
ECCI is extremely strong in the slip system
identification since it is able to identify the Burgers vector of a slip system by contrast analysis.
This means ECCI is not affected by the morphologies of slip traces at the free surface. By
revealing all ty
pes of dislocations, it is possible to investigate the accommodating events at the
grain boundaries at dislocation level, which is generally ignored in the studies that heavily rely
63
on the slip trace analysis. Thus, ECCI is a perfect solution to the estab
lishment of a precise slip
trace analysis. The only drawback for ECCI is the time
-
consuming issue, since it generally need
at least 6 different
g
vectors (must include at least one
g
b
x
u
= 0 or
g
b
= 0) to identify
Burgers vector and several more ti
lt & rotate operation to reveal slip planes. Additionally, too
much topography change after the high strain deformation will add difficulty for the
establishment of channeling condition and resolving individual dislocations.
3.4
The reveal of the geometry of
slip systems at grain boundaries
As discussed in the previous section, one of the benefits of using ECCI in the study of slip
accommodation is the capability to precisely identify all active slip systems at grain boundaries.
This is particularly importan
t in the understanding of slip transfer events between grains in hcp
titanium. The stress state at the grain boundary may not be the same as it is within the grain,
thus slip activation may be completely different at the grain boundary. Such local activa
tion of
unexpected slip systems may not be correctly revealed by the slip trace analysis due to lack of
slip traces
15
. This leads to a bias in the evaluation of slip transfer events during heterogeneous
deformation due to the ignorance of local slip accommodation mechanisms at the grain
boundary.
3.4.1
Categories of slip system interaction at surface
Different slip transfe
r mechanism may be revealed by varied interactions between an
grain boundary were generally categorized into three types in early studies of slip transfer
ev
ents based on the TEM observations [25, 77
-
87]. These classifications are still widely used in
15
If the slip plane of a slip system is close to parallel to the surface, slip band is hard to be revealed at the surface.
64
Figure 23a
-
-
-
-
g) SE images
-
same point on the grain boundary at surface, despite the slip system is highly activated. h
-
i) SE images of
f slip systems that propagate up to the grain
boundary only occurs in one of the grains, with the other grain undeformed. At certain circumstances, the
into the
originated grain.
65
the studies of heterogeneous deformation to describe how slip systems interact at the free
surface. The first type describes the situation that two slip systems in their respective grains
meet at the same point at the grain
boundary. It appears the dislocation slip coming from a
intersection point with the grain boundary. The interaction between the two slip systems is
ll
-
Figure 23a
-
d
, such well
-
correlated slip interaction
can happen not only between slip traces that are close to parallel to each other (Figure 23a
-
c),
but also between those that are far from parallel (Figure 23d). The second t
ype of slip
interaction is shown in Figure 23e
-
g. Despite the strongly activation of slip bands in both grains,
the slip bands do not meet at the same point at the grain boundary, regardless of whether or
not their slip traces appear to be parallel align
-
directly transferred into
Figure 23h&i
. The slip system is highly activated only in one of the grains. It appears the slip
ry in
between, and no shear is transferred across the grain boundary. This slip interaction is then
3.4.2
Limitations of the current category system and free surface biasing effects
It seems convenient to directly apply the concepts that wer
e used in TEM thin film
studies (in
-
situ slip transfer studies) of slip transfer in polycrystal heterogeneous deformation.
-
66
Figure 24a) 3
-
D geometry of slip system i
nteractions at a grain boundary. The two slip systems are considered
as well
-
correlated slip systems since they intersect at the same point at the grain boundary on the free surface.
Although slip systems intersect at the same black point at the surface,
it can be clearly seen that they do not
meet below the surface since the geometric orientations of the two slip systems are different, with
0
o
.
b)
Another 3
-
D geometry of slip bands interactions in the vicinity of a grain boundary. The two slip systems are
defined as non
-
correlated since they do not meet at the surface. However, they may meet at the grain
boundary plane somewhere below the surfa
ce. Although the slip transfer mechanism may be different
between a) and b), current studies of slip transfer ignore this potential difference and directly use the slip
transfer criteria to evaluate the strain accommodating events at the grain boundary.
67
interactions at the surface, nor TEM thin
-
film [79
-
86] observations are able to fully represent
-
Figure 24a&b
show the
ideal 3
-
-
-
slip systems on opposite
sides of a grain boundary plane. The two slip bands from the two neighboring grains will only
meet at one point at the grain boundary plane, since generally
o
, except in very limited
cases. If this is true, then it is hard
to rationalize that slip transfer only occurs at the common
points. Thus, the deformation shear transfer in the regions away from the common point is
currently unclear. In addition to this, not only in this study, but also others [45, 74, 75], have
found
-
-
slip systems. This may suggest the free surface is biasing the observations by creating more
-
sm. Nonetheless, the
alignment of slip bands and the slip transfer among slip systems (
o
) below the surface will
be revealed by ECCI after removing approximately five microns of material through
electropolishing [171, 175].
3.4.3
Comparison of slip system
alignment at and below surface
-
Figure 25a
.
Despite the large angle between the slip traces, slip traces from the two slip systems
consistently meet at the same points on the grain boundary. ECCI facilitated slip trace analysis
reveals the slip system in the upper grain is the (0
10)[
110] prism <
a> slip system with a
higher Schmid factor of 0.48, and its well
-
correlated counterpart in the lower grain is the
(10
0)[
2
0] prism slip system (lower M=0.36). It seems dislocation slip is more easily
activated in the upper grain than in the lower gr
ain due to a higher Schmid factor M with the
68
Figure 25a) SE image showing two slip traces that are well
-
correlated at the surface. The despite a large
angle between the slip traces, they intersect at the same points on a grain boundary. ECCI facilit
ated slip
trace analysis indicates the slip system in the upper grain is the (0
10)[
110] prismatic slip system with
the Schmid factor 0.48. The slip system in the lower grain is (10
0)[
2
0] prismatic slip system, with a
lower Schmid factor of 0
.36. The white arrows mark one of the well
-
correlated slip traces for comparison. b)
ECC image of the same region, but 5 µm below the surface. The observed slip bands are now due to
dislocation contrast, rather than topography. It can be clearly seen th
at the slip bands are misaligned in the
electropolished area, which is indicated by the relative position change of the white arrows. Nevertheless,
the relative spacing and distributions of the slip bands remains unchanged. c) high magnification ECC imag
e
of the area in the red boxed area in figure (b). It shows the activation of (
011)[
2
0] pyramidal
secondary slip systems (M =0.45) in the lower grain from an intersection point of the slip system form the
upper grain at the grain boundary. The sec
ondary slip system propagates a short distance and appears to
merge into the primary slip system in the lower grain. There might be multiple activation of such secondary
slip system, possibly from different sources at different depth in the grain boundary
plane, which is indicated
by the small arrows.
69
same CRSS value, which is also consistent with the observation that deformation slip only
propagates a short distance in the lower grain. It also appears the slip system in the lower grain
is nucleated fr
om the intersection points between the slip system in the upper grain at the
grain boundary to accommodate the induced deformation shear, although it is not clear where
the slip system in the upper grain is originated. Despite the correlation of these s
lip bands (i.e.
two big arrows), the two slip systems do not align perfectly as they might appear to. The
the Burgers vectors is 15.4
o
), suggesting the two
slip systems are still misaligned. Since the slip
normal is 50.4
o
), as shown in
Figure 24a
, it can be imagined that unless
= 0, there will be
significant mi
salignment of the slip planes below the surface.
Figure 25b
is an ECC image of the same area after electropolishing. Different from
Figure 25a,
where the slip bands are observed due to topographical contrast, the observation
of the slip bands come fr
om the contrast of dislocations at certain channeling condition (in this
case
g
= [
]). By comparing
Figure 25a&b
, it appears that the spatial distribution of the slip
bands remains the same (i.e. the spacings and distributions of the slip bands on eit
her side of
the boundary are the same at and below the surface). It also appears that the propagation of
all slip bands in each grain are consistent in the two images (slip bands move to the right in the
upper grain, while the slip bands move to the left
in the lower grain). But critically, despite the
spatial distribution of the slip bands remaining the same between surface and subsurface
observation, the intersection points of the slip bands at the grain boundary are now offset by
approximately 7
m (i
ndicated by big arrows). This is consistent with the expectation that the
70
slip systems will no longer be well
-
correlated below surface. Detailed ECCI analysis of the red
boxed area is shown in
Figure 25c
. A number of secondary slip systems in the lower
grain are
found to be well
-
correlated with the primary slip system in the upper grain (indicated by small
arrows). These secondary slip systems appear to nucleate from where the slip system in the
upper grain intersects the grain boundary. The secondary
slip systems then propagate a short
distance and merge into the primary slip system in the lower grain. ECCI contrast analysis
reveals the Burgers vectors of the dislocations associated with the secondary slip system are
the same as those in the primary
slip system, [
2
0]. However, these secondary slip bands are
on the (
011) pyramidal plane, different from the (10
0) plane of the primary slip system.
Nevertheless, the shear accommodation at grain boundary appears to be facilitated by the
activation of
primary slip system in the upper grain is still achieved by the secondary slip system in the lower
grain. By correlating the surface and subsurface images, the approximate 3
-
D geometry of the
slip systems and the grain boundary can be reconstructed (
Appendix H
), and the angle
between the intersection lines of slip systems at the grain boundary (and the grain boundary
inclination angle
) can be calculated. This allows the i
nteractions between slip systems at grain
boundaries to be assessed in a more comprehensive way. In this particular case (
Figure 25
),
based on the seven
microns
offset between the intersections points of the slip systems at the
grain boundary at five micr
ons below the surface, the grain boundary plane inclination angle
is around 20
o
. The angle
p
-
p
o
, while the
p
-
s
71
around 17
o
the two
activity of this secondary pyramidal slip system is supp
ressed due to a much larger CRSS
value to the prism slip system (1.3~8:1 [25
-
28,24
-
37]). It is reasonable to conclude that the
local activation of secondary slip system facilities strain accommodation at the grain boundary.
The much stronger alignmen
system helps to compensate the incompatibility between the primary slip systems within the
grain boundary below the surface. Since dislocation propagation on the pyramidal plane is not
favora
ble, these dislocations begin to cross
-
slip and glide onto the prism plane at a short
distance away from the boundary, where the propagation is much easier under global stress
state. In conclusion, different from the overall deformation of grain interior,
which is strongly
controlled by the primary slip systems with high Schmid factor/CRSS, local accommodation at
the grain boundary generally requires the activation of secondary slip systems that are better
The local
accommodation at the grain boundary below surface can also be found
-
boundary. One example is shown in
Figure 26
. The prism slip system (M = 0.49) in the
upper
grain appears to be well
-
correlated with the basal slip system (M = 0.40) at the grain
boundary in the as
-
deformed sample. Comparison of
Figures 26a&b
reveals that the relative
spacing and distribution of the slip lines remain the same before and aft
er electropolishing,
72
represents the trace of the grain boun
dary at the intersection point with the strongest band in the upper grain.
b) ECC image of the same area after removing 5
electropolishing. The grain boundary orientation
is significantly changed, which is indicated by the change of the red da
shed tangent line around the same
strongest activated slip band. The slip band spacings and relative distributions remain the same before and
after surface removal. c) Higher magnification ECC image of red tangent line area. A ~50 nm misalignment
betwee
n the two slip systems below the surface has been observed. d) ECC image of the red boxed area in
figure (c). A small number of secondary slip system is observed in the upper grain in this case.
73
although the grain boundary orientation has changed signi
ficantly, as is reflected by the change
-
grain boundary reveals an approximately 50 nm offset between the intersection points at the
grain boundary after elec
tropolishing (
Figure 26c
). This very small deviation between the
primary slip systems across the grain boundary is consistent with the relative high geometric
diffi
limited number of pyramidal dislocations in the lower grain, as well as one pyramidal
secondary slip system in the upper grain. Particularly, as shown in
F
igure 26d
, despite the
Burgers vector of the secondary slip system being the same with the primary slip system in the
upper grain, ECCI contrast analysis shows the dislocations are in different contrast. This
indicates the dislocations have different lin
e directions and different edge and screw
components. Although not shown specifically, the dislocations associated with the secondary
slip system also have the same Burgers vector with the primary slip system in the lower grain.
The angle
p
-
p
between th
e two primary slip system is relatively small, ~15
o
, which is consistent
exist in both grains, it is necessary to calculate the
p
-
s
separately
16
. The
p
-
s
betwe
en
pyramidal slip system in the upper grain is ~ 37
o
, while the
p
-
s
16
to
complete the other half of the analysis.
74
slip in the lower grain is ~22
o
. Other secondary slip systems combinations have also been
studied, including
p
-
s
between primary and secondary
slip systems in the upper grain (~ 25
o
),
p
-
s
limited to the lower grain (~36
o
), and
s
-
s
(~60
o
). These suggest the secondary slip system is
primarily accommodating the shear from the primary slip system across the boundary, rather
than the shear within
the own grain. Additionally, it might not be necessary for the secondary
slip systems to be well
-
correlated with each other. Although the detailed mechanism of this
t
hat local accommodation can still occur by the activation of secondary slip systems despite the
two primary slip systems being well
-
aligned. Atomic studies of such slip well correlated surface
slip transfer with varying
would be insightful for understan
ding the details of the dislocation
mechanisms at these boundaries, but are beyond the scope of the present study.
-
is shown in
Figure 27
. At the
surface, the (0
10)[
110] prism slip bands (M= 0
.30) in the upper grain and the
(10
0)[
2
0] prism slip bands (M= 0.28) in the lower grain do not meet at the same points
on the grain boundary (
Figure 27a
). The lack of correlation between the two primary slip
system is also consistent with the low m
p
-
p
=58
o
), despite the slip systems being readily
active within their respective grains. Following electropolishing, it can be clearly seen that the
relative spacing and distribution of the slip bands remain the same in each grain below the
surfa
ce, suggesting the slip bands are confined in their own slip planes (
Figure 27b
). However,
as indicated by the small white arrows, the relative positions of the intersection points at the
grain boundary are significant deviated, by ~20
µ
m. This suggests
the observed slip traces
across the grain are not associated with each other and the slip transfer between them is
75
Figure 27a) SE image of non
-
correlated slip systems on the surface. The small white arrows indicate the
relative positions of the selec
ted slip bands on the surface. It can be clearly seen that these slip bands do not
intersect at the same point on the grain boundary.
b) ECC image of the same region below the surface. The
relative spacing and distributions of the slip lines remain unchanged within either grain, but the relative
intersection points on the grain boundary changes, as indicated by the small white arrows.
c) ECC image of
the grain boundary area in the red box. Pyramidal secondary slip system is activated from the intersection
points of the incoming prism slip system in the upper grain at the grain boundary.
d) ECC image of the
grain boundary area
in the blue box. Secondary pyramidal slip system is activated to accommodate the
strain induced by the prism slip system in the lower grain.
76
limited. The closer examination at the grain boundary area within the lower grain (red) and the
upper gr
ain (blue) is shown in
Figures 27c&d
, in order to further investigate the local
10)[
110] prism slip system in the
011)[1
10] secondary slip system
(M = 0.32)
p
-
s
= 46
o
). Similarly,
0)[
2
0] prism slip system is also accommodated by a limited number
11) [
110] secondary slip system (M= 0.31) in
p
-
s
= 38
o
). Compared to the poor compatibility between the two
p
-
p
= 58
o
), the induced shear from each
primary slip system is resolved by the activation of better aligned secondary slip systems.
These secondary slip systems also have the same Burgers vector as the primary slip systems
within the respective grains, but on dif
ferent slip planes. But due to a larger CRSS value of the
secondary slip systems, the activity of the dislocation slip is found localized around the grain
boundary. It is speculated that if the sample were to be strained further, these regions might
be s
ources for easier primary slip system activation, due to dislocation cross
-
slip.
Figure
28
0)[
2
0] prism slip system (M=0.45) in the uppe
r grain
is efficiently
p!
other side of the grain boundary (
Figure 28a
). However, close examination on the other side of
the grain boundary has found the local activation of mult
iple dislocation slip systems in the
lower grain (limited to a small zone), with the majority of the dislocations being associated with
the (0001)[1
p
-
s
= 38
o
). A number of
77
dislocations on the pyramidal sl
=0.67. These dislocations appear to be activated to accommodate the strain at the grain
not able to
propagate out of the grain boundary area due to a low Schmid factor. This
observation is consistent with a latest work by Gioacchino et al. [189] that strain can
be
partially transferred out in a more compatible way through the localized crystal lattice rotation.
and there will always be some strain relaxation at the gra
in boundary area through dislocation
activations, or alternatively fracture at the grain boundary.
3.4.4
Comparison of slip transfer parameters among slip system interaction types
With all the crystal orientation information from EBSD and the geometry of slip s
ystems
at grain boundaries, it is able to assess the slip accommodation events at the grain boundary, as
Figure 28a) SE image of the third interaction type. The slip system appears to be blocked by the grain
boundary and no active slip system is observed in the lower grain. b)
ECC image in the red boxed area. A
number of dislocations within a limited zone is observed, with majority belong to the basal plane, and some on
the pyramidal plane.
78
outlined
Section 3.4.3
. This geometric analysis is shown in
Figures 29a&b
, which involves
numeric factors, namely: the angle
between intersection lines of different slip planes on a
grain boundary plane (a key factor in the LRB criteria [83
-
86]), the geometric compatibility
Figure 29 Comprehensive analysis of slip trans
fer parameters between well
-
correlated slip (top 10 lines), non
-
correlated slip (middle 9 lines), and
blocked
slip (bottom 3 lines). a) comparison of the angles
between
incoming primary and outgoing primary (black)/ secondary (red) slip system interse
ction lines at the grain
boundary plane, and the geometric compatibility factor m
between the incoming primary and outgoing
primary (black)/ secondary (red) slip systems. b) Comparison of the global Schmid factors M of the outgoing
primary (black)/ secon
dary (red) slip systems, angle
between the Burgers vectors of incoming and outgoing
slip systems, and angle
between the incoming primary and outgoing primary (black)/ secondary (red) slip
plane normal. Only one value of
is given because the outgoing
primary and secondary slip systems have the
same Burgers vectors. Schmid factors for the outgoing primary slip systems are shown in italics in the cases
where there is no correlated slip observed between the incoming primary and outgoing slip systems. O
nly one
Schmid factor is given for the
blocked
cases because only limited slip was observed in the outgoing grains in
the vicinity of the primary incoming slip systems. The fine dotted line between the non
-
correlated and
blocked
slip band interactions
indicates the similarity in the accommodating behavior.
79
the angle
q}
between slip plane normals, and the global Schmid factors M [45, 46, 74]. Overall,
despi
systems in all cases are mostly prism slip systems, with some basal slip systems.
ed with the
pyramidal slip systems, with a limited number of basal slip systems. It should be noted
that there are indeed other types of dislocations observed during the ECCI analysis, however,
sults.
Cumulative distribution plots of
, and M are provided in
Figure 30
in order to
discover the influences of these parameters on the accommodation behavior. The cumulative
Figure 30a
) shows the importance of alignment
of slip systems in the
primary slip system when they are well
-
when the slip systems are not correlated. Similarly,
it also appears to have a good correlation
-
to
-
primary and primary
-
to
-
sec
ondary slip systems. In general, it is expected that significant slip
transfer will occur between well
-
correlated slip bands, with relatively good alignment between
slip systems. As the primary slip systems become more poorly aligned, it is possible that
secondary slip systems will be more active to compensate the lack of slip from the primary
systems.
-
80
Figure 30 Cumulative distribution plots of the a) the compatibility factors m
between the various slip systems,
b) the angles
between the various outgoing slip systems and the incoming primary slip system, and c) global
Schmid factors M of the outgoing
slip systems.
81
which is consistent with the lack of interactions between the primary slip systems. Meanwhile,
-
primary slip system to accommodate the accumulat
ed strain at grain boundaries. This trend is
clearly shown in
Figure 29a
-
-
co
-
-
can also be interpret
primary slip system across the grain boundary being accommodated by the limited activation of
secondary slip systems in the opposite grain.
The analysis of
in
Figure 30b
shows a sim
-
p
-
p
(and
p
-
s
) than those that
-
p
-
s
is generally smaller than
p
-
p
, but in some cases larger.
It appears the
strong correlation is related to the strong alignment of the slip plane
intersections at the grain boundary and may directly affect the ease of slip transfer. In one case
(the third row in
Figure 29a
),
One hypothesis is the
82
angle
Figure 30c
is the cumulative distribution plot of the global Schmid factor M. The global
-
-
ses. This
-
primary (majority prism ) and secondary (majority pyramidal )
slip systems are not so
significant, it is generally believed these secondary slip systems are more difficult to be
activated due to much higher CRSS values [25
-
28,24
-
37]. This indicates the global Schmid
factor is not an important element in the local ac
tivation of secondary slip systems, although it
controls the activation and propagation of the primary slip systems.
3.4.5
C
onclusions on grain boundary local accommodation activities
The correlation of (ECCI) images on and below the surface reveals the 3
-
D geom
etry of
the slip systems at the grain boundaries, allowing the comprehensive assessment of slip
accommodation behavior between slip systems at the grain boundaries. In general, the slip
bands are confined within their respective slip planes on and below s
urface, with the relative
83
spacing and distributions remaining the same. Despite the detailed mechanism unknown, the
free surface is indeed biasing the behavior of slip systems at the grain boundary region. The
-
he surface lose the well
-
correlation, and the
resulting offsets between slip systems locally activate the secondary slip system to
accommodate the divergency, trying to restore the integrity between the slip systems. The
Burgers vectors of these secondary
slip systems are generally the same as the primary slip
systems in the same grain, however, are associated with different slip planes. Due to higher
CRSS values, the secondary slip systems are not readily activated, and cross
-
slipping into the
primary sl
ip systems at a short distance from the grain boundaries. Nevertheless, the apparent
-
lip plane/grain
boundary intersection lines (small
).
Consequently, as the slip systems become less and less
correlated, local activation of these secondary slip systems become more prominent. Overall,
the local shear accommodation at the grain boundary
in the heterogeneous deformation are
more complicated than it appears. Thus extra care should be taken in the study of slip transfer
due to free surface biasing effect.
3.4.6
Short discussion on the tangential continuity theory
The frequent observations of t
he local activation of secondary slip systems at grain
Livingston and Chalmers [71, 80]. The observation in this study agree with the tangential
continuity that the accu
mulated strain at the grain boundary cannot be fully accommodated
between two (primary) slip systems. Although ECCI analyses in this study have found some
84
other types of dislocations that are not associated with the primary or secondary slip systems
aroun
d the grain boundary, it is not known if they have contributed to the strain
accommodation activities and their densities are usually very low. Nevertheless, the constraint
(
Section 1.1.4
) cannot be fully fulfilled since it requires four slip systems to f
ully accommodate
the strain at the grain boundary. These observations reveal the fact that current understanding
of the grain boundary accommodation activities in polycrystal deformation is still incomplete
17
[104]. The slip transfer studies that only in
volve the interactions between two primary slip
systems are far from realistic to fully represent the plasticity behavior at the grain boundaries.
3.5
The direction of slip propagation within polycrystals
The direction of slip transferring within a grain and across a grain boundary is always
ignored by plastic models and majority slip transfer studies. Most studies are insensitive to the
direction of slip transfer, and assume the shear is always following
one direction [75
-
77, 87
-
89].
This may be reasonable in ideal experiments such as bicrystals since slip transfer parameters
do not have a direction vector [83
-
86, 95]. However, deformation is not always in
one direction in real
-
life deforma
tion of polycrystals. There might be multiple nucleation
sources, either within a grain, or at a grain boundary [190]. For example, one may expect two
primary slip systems nucleated from a left grain boundary and a right boundary will propagate
in a diff
erent direction. Yet no research has been done on the propagation direction of a
dislocation slip. This long
-
ignored element help capture the overall flow of deformation shear
17
The author carried out analysis based on tangential continuity model, trying to find the number of active slip
syste
most of the accommodating systems were typically predicted to be in the
grain
.
Furthermore, the primary observed accomm
odating systems were typically not consistent with these
predictions
. It seems the model prefers self
-
accommodation if no external shear is applied as a driving force.
85
travelling within polycrystals, and may further improve the plasticity model i
n the deformation
prediction. Luckily, during the throughout ECCI analysis across several neighboring grain
patches in sample 3 after the electropolishing, it is found that ECCI is able to reveal the
potential travelling direction of individual slip syste
m.
After the EBSD
-
based slip trace analysis on the as
-
deformed sample (
Figure 17b
), ECCI
was applied on the electropolished sample for the identification of Burgers vectors (
Figure 31
).
Despite the change in the overall grain shape during the electro
polishing (
Figure 31b
), crystal
orientation before and after electropolishing is not changed [96]. Dislocations on the slip
bands
18
can be identified through the contrast analysis as outlined in
section 3.2.2
(shown in
Figure 31d
-
i
). With the informatio
n from the slip trace analysis, the primary slip system in grain
1 is identified to be (1
)[11
0] prism slip system (M= 0.15), with additional cross
-
slip
dislocations on (1
) pyramidal plane (M=0.23). Likewise, the primary slip system in grain 2
a
nd 3 is (1
)[11
0] prism slip system (M=0.41) and (10
0)[1
10] prism slip system
(M=0.48), respectively. Based on the observation on the as
-
deformed sample (
Figure 31a
), the
deformation slip propagation within these grains is linked by the slip
bands that propagate
through three grains. Some of the slip bands are quite distinct in grain interior (grain 2&3),
while some slip bands in grain 2 become less distinct as they approach the grain boundary with
grain 3. Additionally, despite the distinct
slip bands in grains 2&3, they do not appear to meet
at the same points on the grain boundary in between. Meanwhile, it appears the slip bands in
grains 1 and 2 meet at the grain boundary, however, the slip propagation in grain 1 is not easy,
18
The correlation of the slip bands at and below surface is possible since the relati
ve spacing and distributions of
the slip bands remain the same, as outlined in
Section 3.4.3
. Thus it is able to find the specific slip system after
electropolishing.
86
Figure
31a) SE image of grains 1
-
3 in the as
-
deformed sample 3. The slip systems are identified by the ECCI
facilitated slip trace analysis, which are marked by dash lines in different colors and labelled in different
colored fonts. It appears the slip lines i
n grain 2 and 3 are not well
-
correlated. The slip systems in grain 1
appear to meet the slip systems in grain 2 at the boundary, while the slip propagation in grain 1 is difficult. b)
General BSE image of the same grains 1
-
3 in the electropolished sample
3. All the slip traces are removed by
the electropolishing and no clear contrast is seen because the sample is not under a channeling condition. c)
One example of ECCI analysis on the electropolished surface. At g = [1
], the slip line contrasts are d
istinct.
These contrasts come from dislocation contrasts rather than topography. d
-
i) High Mag ECC images taken
from the red boxed area in grain 3 for the Burgers vector identification. The Burgers vector is [1
10]. With the
information from the slip t
race analysis, the slip system that is active in grain 3 is
(10
0)[1
10]
prism slip
system. ECCI identification of slip system in grain 1 and 2 can be found in the
Appendix
F
, following the same
approach.
87
as reflected by the wavy slip bands. Overall, the deformation evolution within the grain patch is
quite complicated.
by the neighboring grain to maintain the overall integrity. It will be easy to study the
accommodation behavior if the direction of deformation flow within the patch is known.
3.5.1
Slip transfer direction identified by ECCI contrast analysis
The ECCI analysis reveals that dislocation morphologies appear differently when
following a certain direction. One example can be found in
Figure 32
. After electropolishing, at
a certain channe
ling condition, the contrast of slip bands is a direct consequence of the
channeling contrast of the individual dislocation in the slip band. Slip bands that show strong
contrast reflect a large number of dislocations in the slip band, suggesting this sli
p band is
carrying a large shear (
Figure 32a
). It appears that dislocations are nucleated from the grain
boundary between grain 2 and grain 3, based on the following argument. The slip
band width
is very small near the grain boundary (~0.02
µ
m in
Figure
32b
), but the slip line broadens
and become even
cross
-
slip, although the mechanism is not
clear. Once the dislocation slip band reaches the
grain boundary with grain 4, dislocations become widely distributed along the grain boundary
Despite the change in slip band width, it is interesting to note that the contra
st of
dislocations appears different at different positions along the slip line. When the dislocations
are nucleated from the grain boundary with grain 2 (
Figure 32a
), they appear as dots,
suggesting the dislocations are close to perpendicular to the surf
ace at this channeling
88
Figure 3
2
a) electropolished surface of grain 3. b
-
f) ECC images taken at the same channeling condition at
different positions along a slip line. Dislocations are limited in a sharp and narrow line from the nucleation
point at
the grain boundary with grain 2, and spreading out as going deeper into the grain and will finally
distributed around the grain boundary with grain 4.
89
condition. Based on the sense of black/white contrast of the dislocations, almost all of the
dislocation
s along the slip band at the grain boundary area have the same Burgers vector sign.
As the dislocations propagate into the grain interior, more dislocations show opposite contrast,
suggesting they have the opposite sign of Burgers vectors (10% at position
c, ~30% at position
d, ~ 50% at position e, and ~65% at position f at the grain boundary with grain 4). Additionally,
in the center of grain 3 at positions c and d, ECCI also reveals more dislocations appear as lines,
This indicates the dislocations are aligned more parallel (or
elastic relaxation [140
-
144]. The observation of more dislocations with opposite sign Burgers
ve
-
dislocations cross
-
slip become more prevalent.
Based on the change of slip band width, and the observations of more cross
-
slip activity,
it is reasonable to beli
eve that once a slip system is nucleated from a source, the slip band is
sharp and all the dislocations are limited strictly in one plane. As the dislocation slip propagates
forward into the grain interior, the slip band becomes broader due to dislocation
cross
-
slip.
Once the slip system reaches the grain boundary with grain 4, more cross
-
slip event is expected
to happen
a
s the dislocations pile up at the boundary, resulting in a wider spread of dislocations
in the grain boundary region. This indicates
the overall slip transfer direction within grain 3 is
from the upper left boundary with grain 2 to the bottom right boundary with grain 4. Similar
analysis has been done on other primary slip bands showing strong contrast (
Appendix I
), all
the slip transf
er direction in grain 3 is from the upper left to the bottom right.
However, there are indeed some slip bands with weak contrast propagating in the
90
Figure 3
3
a) Low mag ECC image of grain 3,
the four
arrows indicate the slip transfer directions of sl
ip bands
with strong ECC contrast. The dashed arrow indicates the opposite transfer direction of a slip band with weak
-
f) ECC images following the slip band from the boundary with
grain 4 into grain interi
or. Overall dislocation density is low compared with the case in Figure 31. As the slip
band broadening effect further dilutes the dislocation concentration, the dislocation contrast of the slip band is
hardly detectable.
91
opposite direction. One typica
l example is shown in
Figure 33
. A low mag ECC image (
Figure
33a
) reveals this apparent weak slip band nucleate from the boundary with grain 4 and
propagate towards the upper left. The contrast of the slip band becomes even weaker during
its propagation,
and finally fades around position f,
magnification ECC image at the grain boundary confirms the observation that dislocation
nucleated from the grain boundary as a narrow band (
Figure 33b
). Additionally, the contrast
ana
lysis also reveals these dislocations are dot
-
like, with the majority showing opposite
contrast with those in
Figure 32b
. This indicates these dislocations have Burgers vectors of
opposite sign (and opposite propagating direction) than those in the primar
y, high contrast slip
bands. Consequentially, dislocation cross
-
slip occurs during the propagation towards the upper
left of the grain. As the dislocation have not propagated as far, the slip band finally become
nciteful reminder that a slip band that shows weak
SEM topography contrast does not always mean the slip band is less activated. The apparent
reaching the grain boun
dary, it might be a result of broadening effects that dissipates the
dislocations in a wider region along the grain boundary.
3.5.2
Comprehensive analysis of deformation within grain 1
-
3 patch
By and large, the complete ECCI analysis provides an understanding
of
the deformation
evolution of the grain patch.
The analysis suggests the majority deformation shear is carried
away from the upper left boundary with grain 2 to the lower right boundary with grain 4
through dislocation slip.
However, deformation also
occurs from the lower right boundary and
moves in an opposite direction, although the shear is not significant.
This might be a result of
92
accommodation to the shear from the grain 4.
By the application of similar analyses along most
of the slip bands in
grains 1&2, it is possible to outline the potential deformation evolution
history within the grains of the
polycrystal patch.
As indicated by the wavy slip band at the surface (
Figure 17b&31a
), the deformation
shear in grain 1
appears difficult, as indi
cated by the low Schmid factor (M=0.15). This is
consistent with the observation of wide spread of dislocations along the slip band after
electropolishing (
Figure 34a
), as the slip band propagates to the upper left into the grain
interior.
However, due to the highly active slip system in grain 2 (M=0.41) nucleated from the
grain boundary, grain 1 needs to accommodate the shear by the activation of dislocation slip
that is better aligned
19
= 32
o
)(
Figure 34b
), although
the further propagation of the accommodated slip system is difficult in grain 1. It is worthy to
point out that none of the high Schmid factor deformation systems ( slip systems and
twinning) are observed. This suggests grain
1 is passively activated in the heterogeneous
deformation.
accumulation at the grain boundary, shear can be carried away through dislocations to the
lower right into grain
2 by the easy activated slip system (M =0.41). Again, the initial slip band
in grain 2 nucleated from the boundary with grain 1 is sharp and has a small band width (
Figure
34c
). Due to slip band broadening effect, the slip band width is increasing as
it is propagating
inside grain 2 (
Figure 34d
). Once the slip band in grain 2 reaches the lower right grain boundary
19
The slip transfer parameters can also be calculated in the situation that a grain boun
dary kicks out two
direction of slip transfer.
93
Figure 3
4
a) A wide spread of dislocations in grain 1 from where a dislocation slip in grain 2 is nucleated at the
grain boundary. b) Wide spread of dislocations in grain 1 interior. c) a sharp slip band is nucleated from the
boundary in grain 2. d) The slip band
is wider in grain interior than it is at the grain boundary. e) Intersection
of dislocation slip at the grain boundary with grain 3, slip systems in grain 2 and 3 are not correlated.
Dislocation density is high in grain 2. F) Accommodating dislocations
nucleate from the intersection point of a
slip system in grain 3, and propagate only a short distance with significant broadening effect.
94
with grain 3, dislocations are diffused along the grain boundary area in grain 2 (left side of the
boundary in
Fig
ure 34e
) to partially relief strain accumulation [75].
Despite the strong activity of the slip system in grain 3 (M =0.48), the two primary slip
= 40
o
), as reflected by the offset
between intersecti
on points of the two slip bands at the grain boundary (
Figure 34e
).
This
suggests the shear from grain 2 is not efficiently transferred to grain 3, thus the grain boundary
needs extra efforts to accommodate the accumulated shear to maintain its integrity.
-
=28
o
) with the highly active slip system in grain 3 (
Figure 34f
). This
observation indicates that slip systems can be nucleated from the same point at the grain
boundary and propagated in an opposite direction into their respective grains. This can be
interpreted as slip systems in both grains are activated to accom
modate the shear at the grain
boundary.
3.5.3
Conclusions of polycrystal deformation identification
As would be expected, the detailed deformation evolution of polycrystal patch is
complicated. Nonetheless, slip bands are found to spread out during their prop
agation, which
gives a key to understanding the deformation evolution. With additional information of the
change of dislocation contrasts and morphologies at different positions following the slip band,
it is possible to identify the slip transfer directi
on within a certain grain. Based on the
observations of the nucleation of different slip systems, it is very interesting to point out that in
this very study, it appears most dislocation slip systems are nucleated from the grain boundary,
rather than some
where within grain interiors. This agrees well with arguments that grain
95
boundary can be dislocation sinks and sources [52
-
55, 190]. As there might be more defects on
the grain boundary plane due to irregular atomic arrangement in that region, the grain
deformation. This might be insightful to the plasticity modeling that both grain boundary and
grain interior can be sources for slip activations. It is also interesting to
find out that the slip
systems of the same type within one grain may not propagating towards the same direction,
this depends on where the dislocation is nucleated (i.e. slip systems nucleated from the left
grain boundary will propagate in a opposite dire
ction with those nucleated from the right
boundary within the same grain). Additionally, grain boundary accommodation can happen
the same boundary and propagate into their respective
grains. Although the slip transfer parameter is not affected by the direction to slip transfer,
understanding the transfer direction and the deformation sequences within polycrystal is
critical for the
comprehensive understanding of polycrystal engineering material (i.e. local work
hardening, crack nucleation, etc.).
96
4.
Summary
In order to correctly understand the heterogeneous deformation of hcp titanium, post
-
deformation analysis need to be precise an
d comprehensive. The preciseness refers to the
correct identification of slip systems activated during the deformation, and the collectiveness
requires the analytical method to identify all the slip systems activated in the accommodation.
This comprehens
ive analysis cannot be simply achieved through one approach and needs
assistance from other analytical techniques. Slip system identification is mostly based on the
crystal orientation data from EBSD and the SEM observation of traces at the surface. With
the
improvement of EBSD and cross
-
correlation techniques that detect local lattice rotations [185,
188], and the AFM and DIC [162
-
165] that measure the displacements
resulting from
dislocation slip, slip system identification and quantification is becomin
g more precise.
Nevertheless, the current analyses are not able to identify the slip systems that do not
contribute to the slip traces. On the other hand, ECCI can precisely identify the Burgers vector
of the slip systems based on dislocation contrast an
alysis [144
-
determinant analysis and avoids the calculation of potential deviation factor in DIC and cc
-
EBSD
[185, 186, 188]. Implementation of the ECCI technique in concert with the current approaches
moves the characterizati
on of local slip
behavior in heterogeneous deformation substantially
forward.
Slip accommodation at grain boundary regions in polycrystal deformation is quite
complicated. By correlating SEM images at the surface with subsurface ECCI images, the 3
-
D
geome
try of the slip systems at the grain boundary area is revealed, allowing the assessments
. Additionally, subsurface
97
-
surface, and reveals the
locally activation of secondary slip systems that help compensate the incompatibility between
the primary slip systems at the grain boundary area in the subsurface. This observation agrees
well with the long
-
forgotten tangential
continuity theory [71, 80], that indicates that the
interactions between two primary slip systems are not enough to accommodate the strain at
the grain boundary. Although this study only finds the local activation of secondary slip
systems, this does not
mean there cannot be more dislocations involved in the accommodation,
especially at higher strain levels. Nevertheless, the current research extends the understanding
of the complete nature of the accommodating mechanisms at the grain boundary, and again,
suggests the slip transfer mechanisms developed based on the slip trace analysis may be
limited.
In addition, the slip band broadening effect is clearly revealed through ECCI analysis in
electropolished hcp titanium samples.
The comparison of dislocati
on morphologies, dislocation
contrast, and dislocation density at different positions in slip bands suggests that the vast
majority of dislocations are nucleated from the grain boundary, rather than sources within the
grains, at least in the early stages o
f deformation.
However, this does not suggest dislocation
sources cannot be found in grain interiors, as evidence shows that as the slip bands propagates,
more dislocations with opposite sign are observed in the slip bands.
It should be noted,
however, i
t is not known if this broadening effect is unique to hcp metals due to the lower
crystal symmetry, or is a universal phenomenon in all metals.
Regardless, this discovery does
help to establish an understanding of the deformation evolution in polycrystals
.
98
5.
Outlook and Future Direction of ECCI
In the latest future, one of the easiest things that can be achieved is whether slip band
broadening effect is present in cubic materials or other hcp metals.
This effect will be valuable
if it can be widely appli
ed on different materials.
It will also be interesting to investigate the slip/twin interactions by ECCI in hcp titanium
since deformation twin are also common in the plastic deformation in some other titanium
materials. One example is shown in
Figure 35
. In this figure, deformati
on twin is observed in
the upper grain, whereas two types of dislocation slip systems are observed in the lower grain.
It is interesting to investigate how the slip/twin interactions are at the grain boundary.
Additionally, it is interesting to find dislocations kicking out of the tip of the deformation
twin and propagate into grain interior. It is also interesting to investigate the relationship
between dislocation slip and the twin, and how dislocations become a part of twinning during
the defor
mation (in
-
situ if necessary). Sequential electropolishing of the sample may also
Figure 35
Deformation slip in the lower grain interacting with the deformation twin in the upper grain. ECCI
analysis in the red boxed area shows the dislocation propagating out of the tip of the deformation twin.
99
reveal the geometry of the twin in the subsurface, which may also bring more insights in the
twin evolution.
Since 3
-
D printing of titanium gears or other
consumables are becoming more and more
important in the aerospace industries, it will be an interesting and short project to correlate the
mechanical behaviors of 3
-
D printed titanium samples (using different methods, such as
powders or wires) during diffe
rent stages of processing with the dislocation densities, phase
changes, etc. This may help guide the industry to improve the overall quality of 3
-
D printed
titanium materials.
By and large, ECCI is a strong SEM near
-
surface
-
based analysis technique that
is
complementary to many other techniques.
This non
-
destructive technique is especially useful
for the in
-
situ study of continuous polycrystal deformation without destructively damaging the
sample.
Thus far, this study only reveals that post
-
deformation
analysis is able to provide some
clues to deformation history.
With careful design in future, one may be able to observe the
deformation evolution from the initiation to the final structure of a slip band.
This may be
useful for understanding how the dis
location density, morphology, and contrast changes with
increasing strain, and thus provide the opportunity to link the macroscopic deformation with
the dislocation
-
scale activities simultaneously on the same target area.
This approach is very
advantageou
s over the other destructive studies, such as FIB
-
lift
-
out TEM, since it is extremely
hard to find two same grains with even similar grain and boundary orientation characteristics in
assessable polycrystalline groups.
Unfortunately,
the electropolishing
in this study is hard to control precisely.
Removing
the surface topography typically requires about 2 µm removal, and establishing precise uniform
100
removal rates can be difficult.
As a result, the surface removal in this study are typically around
5 µm o
r more.
It is anticipated that it will not be possible to remove material with enough
precision to track individual dislocations with electropolishing.
An alternative approach may
be the Xe
+
plasma FIB technique [191, 192], which is able to precisely c
ontrol the surface
removal.
This approach has the potential to allow high resolution real 3
-
D reconstruction of a
full dislocation structures below the surface (One should note the potential artifact induction
and titanium hydride precipitation during pla
sma FIB).
With proper coding facilitation, in future
ECCI may be extended to automated identification of
Burgers vectors, slip line directions, slip
planes and the edge/screw component of a dislocation.
Currently, as ECCI is effectively a
manual techniq
ue, the identification of all of these parameters is done tediously by collecting
five or more ECC images at different channeling conditions and long scan times (10 minutes per
scan).
101
APPENDICES
102
APPENDIX A
Electropolishi
ng mechanisms and parameters used in this study
103
The sample was electropolished in a proper setup, consisting a power supply (can switch
between 0 ~ 120 V and 0 ~ 30V), a cathode (6 x 6 x 2 mm stainless steel plate) and anode
(sample with the polished area facing the plate), a magnetic stir (50 mm in le
ngth) with a
magnetic stir plate, electrolyte (in a 1000 ml baker) and cold bath (200 ~ 300 ml methanol
cooled by liquid nitrogen/dry ice or a more stable control of temperature during
electropolishing). Two different electrolyte compositions were used in
the study with
correlated electropolishing parameters, as listed in
Table A1
. It is worthwhile to mention that
the experimental parameters and composition of electrolytes may be different if the minor
element components of titanium are different. Nevert
heless, the parameters in the table are
based on freshly made electrolyte, which can be safely used for cumulative 10 ~ 20 times
without changing the electropolishing result.
Table A1
.
Electropolishing parameters
104
Figure A1 A scheme shows the general four stages of electropolishing with respect to different voltage and
current density ratio.
Figure A2
Left) Below 29 V results in etching of metal with rough surface
under optical & electron microscope.
Middle) A good polishing zone results in shinning & smooth surface with good contrast under electron
microscope. Right) Above 40 V results in a dimmer surface in optical microscope. Under electron microscope,
pitting
occurs, especially at grain boundaries, with slightly worse SACPs.
105
The electropolishing outcome is quite complicated based on the applied voltage and
current (
Figure A1
).
The general process usually falls into several stages [175, 176, 177], which
include: I. The etching of metal through a direct dissolution at low voltage; II. The passivation of
the metal surface by creating an oxidized layer at slightly elevated voltage;
III. The polishing of
the metal by the consistent dissolution & diffusion of anions through the stabilized passivated
layer; IV. Pitting and gas evolution that induces imperfections on the surface beyond the
polishing voltage. The comparisons of optical,
microstructures under scanning electron
microscopy, and electron channeling patterns are shown in
Figure A2
, indicating the perfect
surface finish after optimization of electropolishing parameters. As it is shown in the
Table A1
,
electrolyte A is speciall
y used for the controlled removal of surface material during
electropolishing with 2 µm/min (the surface removal is calculated in
Figure A3
) and the
electropolishing can be finished within 2 minutes (it needs 10 ~ 20 s to reach a steady
-
state that
the poli
sh process is homogeneous throughout the sample), however, some grains are suffered
with hydride precipitation (around 1 out of 100 grains). On the contrary, electrolyte B is free
from hydride
formation [174], but it is almost impossible to control the el
ectropolishing rate (~
8 µm/min).
Figure A3 Calculation of the surface removal by
the Vickers indent. Several Vickers indents are placed at the
surface before electropolishing and final result is the average of each calculated one.
106
APPENDIX B
Removal of polymer film and gold nanoparticles (AuNPs) for DIC patterning
107
Introduction:
As indicated from recent papers [161, 162], with the addition of gold (III) chloride
2
O) and trisodium citrate dihydrate (C
6
H
5
Na
3
O
7
2
O) to produce AuNPs
as the patterning material, organosilanes such as (3
-
aminopropyl) trimethoxysilane (APTMS) or
(3
-
mercaptopropyl)methyl) dimethoxysilane (MPMDMS) was added as to cova
lently bonding to
the dangling hydroxyl group on the metal surface to form a monolayer that was able to
immobilize AuNPs through coupling reactions. Although the patterned DIC technique offers
precise strain measurement during deformation without relying
on complicated and much
expensive experimental set
-
ups, the only limitation of this technique is the coating itself. Due
to overshadowing of the organic coating and nanoparticle which disturb electron interactions
with the sample, and because of artifacts
created on the sample surface, no other studies have
been accomplished after DIC patterning, such as CC
-
EBSD and AFM that are also capable of
monitoring strain development, or perform surficial analysis such as ECCI. The general
approach is to mechanical
ly polish off the surface polymer within a short time, however in
reality, this approach is nearly impossible to peel off the layers without damaging the surface of
the sample. As a result, using a chemical reaction which is selectively targeting only the
nanoparticles and polymer without touching the metal is the ultimate way to perfectly address
this issue. This short paragraph is provided, describing how the patterning layer is removed
through chemical reactions.
Based on many synthetic papers [178
-
182
], the best way of removing AuNP as well as
cleaving off the polymeric silyl ether layer was to use strong halogen reagents, such as F
-
. A
general concern is to use hydrofluoric acid (HF), however upon consideration, the reagent has
108
to be able to penetrat
e the polymer layer and react with the AuNPs but blind to the titanium,
which is also vulnerable in the acid, a mild organic fluorine source should be selected. In this
research, tetra
-
n
-
butylammonium fluoride (TBAF
O and 1mol/L TBAF in THF) is used (
from
Sigma
-
Aldrich) since this is also considered as a phase transfer catalyst which can bring water
and immiscible organic solvents together. Although TBAF is not the only reagent or the best
among all alternatives (fluorotrimethylsilane, referred as the
TMS
-
F, may also work but much
expensive), picking the best reagent is not the main purpose of this paper.
Experimental procedures:
The as
-
deformed Ti
-
7Al (
Figure 13 a
) dog
-
bone tensile sample was provided by
terns coated on surface and strain map collected.
The final ingredients used in this experiment were 10: 1: 1 weight ratio of TBAF, chloroform,
and ethylene glycol. The detailed procedure of uncoating is as follows:
1.
Merge the sample completely in the so
lution at around 30 ~ 35
o
C for 1 hour.
2.
Take the sample out, clean by 5
-
10 mins sonication (20
-
40 kHz) in a baker of soap
water (pH ~ 8), flash with ethanol
-
DI water
-
ethanol, air dry and track the progress.
3.
Repeat 1 and 2 until the surface is cleaned
enough.
4.
Place hand soap on the surface, which is then swept off by cotton stick, go through final
ethanol
-
DI water
-
ethanol washing and air dry.
Results and discussion:
-
free, with
little residual AuNP remained. Nonetheless, the detailed uncoating progress with time was
109
Figure A4a1) Surface condition of as received sample. a2) SE image of AuNPs taken at high magnification from
the red box, with weak SACPs due to interference.
b1) Surface condition after 1 hour at 30
o
C, showing
removal of majority patterning material. b2) SE image showing one of the unremoved clusters of AuNPs, with
sharp SACPs. Particles in these areas are the focused point during DIC data
acquisition. Long time exposure of
beam may condense the NPs into the material or strengthen the bonding interaction, with detailed mechanism
unknown. The patterns c1) At 4
th
hour, all nanoparticles are consumed, including the clusters. c2) SE image of
the same area with b2, showing clean sample, with sharp slip traces and grain boundaries. d1) 24
-
hour
reaction time of an undeformed control sample 1, showing the slightly etching of material. d2) SE image of the
etched area in the red box, showing line
type etched marks and small etched cavities. e1) Surface condition of
an undeformed control sample 2 after the 3
rd
hour in solution at 50
o
C, also showing large areas being etched.
e2) SE image of the red box area, showing surface material has been etche
d away
.
110
shown in
Figure A4
, with extra uncoating information (i. e. uncoating at elevated temperature,
longer reaction time, etc.) acquired from control samples. In
Figure A4 b1&2
, after one hour, it
can be seen that major particles have been washed away,
with clusters of AuNPs (and some
fiducial marks, although not shown in the image) left on the surface. Those were possibly
segregated due to long exposure to the high energy electron beam during DIC data acquisition.
The reagents are hard to get access t
o the clusters as the surface to ratio was diminished after
segregation. Meanwhile, dramatic improvement of SACPs before and after uncoating (
Figure
A4
a2
&
b2
) suggests the removal of particles as well as the polymer layer. With longer reaction
time up to
the 4
th
hour, all particles have disappeared, even the robust clusters, suggesting the
completion of the uncoating process, which is shown in
Figure A4 c1&2
. It should be noted
that although F
-
anions are mostly locked in the organic environment so that
they are only
-
are still considered to
be aggressive to Ti metal. This has been proved that longer reaction time (i. e. 24 hours as
shown in
Figure A4 d1&2
) and higher temperature (i.
e. 50
o
C as shown in
Figure A4 e1&2
).
Although not shown specifically in this manuscript, similar progress can be achieved by simply
Figure A5a) SE image of a random area, showing clear slip traces on surface after the uncoating. b) 2
-
D AFM
map showing topographic information of the same area. c) 3
-
D AFM map showing clear steps from the sli
p
band and the grain boundary (concaved).
111
using 1 mol/L TBAF in Tetrahydrofuran (THF), with little residual AuNPs clusters left on the
surface (7 hours), thus t
his can serve as alternative reagent if not asking for complete extinction
of AuNPs. Neither approach harms the surface within the reaction time and the surface is able
to perform ECCI and AFM analysis on the sample with little interference, as shown in
F
igure A5
.
Conclusion:
A weight ratio of TBAF: CHCl
3
proved to be efficient in removing the pattern after DIC with almost no harm to the surface if
using properly.
112
APPENDIX C
Strain measurement after four
-
point bending
113
As the sample surface between supporting pins is
experiencing uniaxial tensile stress, at
low strain level that the sample does not have too much bend curvature, the tensile strain is
approximately measured from the distance change between triple points of different grains
along/close to the tensile dire
ction before and after deformation using Image J
TM
or other
image processing software, as shown in
Figure A6
.
Figure A6
L
eft) The distance between triple points before deformation.
R
ight) The distance between same
triple points after deformation.
114
APPENDIX D
The calibration of MIAR III FEG
-
SEM with SACPs module for ECCI and CC
-
EBSD analysis
115
As it is of critical importance to ensure the calibrated status of the SEM for the accuracy
of ECCI and CC
-
EBSD analysis, this section provides a detailed procedure on the calibration of
MIRA III FEG
-
SEM. Before proceeding, it is highly recommended to rea
d the manual from MIRA
III and understand the terms that are frequently used in electron microscopes: focus,
magnification, stigmation, wobble, etc.
Proper alignment of column and gun:
There are several modes provided in MIRA III FEG
-
SEM in the CMSC center of MSU,
namely: resolution mode, depth mode, field mode, wide filed mode, and channeling mode.
ECCI images are taken in the resolution mode, while the SACPs are collected in field m
ode with
beam rocking and shown in the channeling mode for the establishment of a diffraction vector
g
.
Thus, it is critical to ensure the beam alignment in each mode. The general alignment usually
to the conditions for ECCI analysis. First, in resolution mode, repeat focusing and increasing
image magnification until an out
-
of
-
focus target (
Figure A7a
) is in focus (
Figure A7b
) at the field
of view 2 ~ 3
µ
m. Ast
igmatism at high magnification is corrected during focusing using the
Figure A
7
Secondary electron (SE) images taken at the field of
view of 3.86
µ
m that shows: a) The particle is
out of focus. b) The dirt is in focus, but astigmatic. c) The dirt is stigmatic and in perfect focus.
116
Figure A7c
). During this adjustment, the
swinging
when the electron beam is moving back and forth across the focus.
Aft
er the beam alignment in the resolution mode is finished, the second alignment is
done in the field mode. Additional focusing operation in this mode is not necessary since this
operation is already done in the resolution mode, but the aperture should be w
ell aligned using
Figure A
8
a) SE image in resolution mode (field of view 3.86
µ
m) that shows the opti
c axis (black cross) is on a
particle. b) SE image in field mode (field of view 65.3
µ
m) that shows the deviation of optical axis from the
resolution mode. Although it appears blurry, this particle is already in focus in field mode since the resolution
b
etween modes is different. c) SE image in field mode that shows the optical axis is moved back to the same
Figure A
9
a) An asymmetric SACP aperture with overlapping of patterns from surrounding grains. b) A
symmetric, perfect round aperture with interference signals from surrounding grains. c) A perfect circular
aperture within pattern only from the target grain.
117
targeting the same position both in the resolution mode and field mode, so that the pattern is
collected from the target area (
Figure A8c
). This ste
p is critical when the ECCI analysis is
performed near grain boundaries, at high tilt conditions (> 10
o
),
or in small grains (grain size
larger than 20
µm for a perfect SACP in this MIRA SEM
) for the diffraction condition set
-
up. The
final SACP
After the aperture alignment, the aperture should appear to be a perfect circle, with an un
-
overlapped pattern (
Figure A9c
). Crisp ECC images with precise pattern informati
on can be
achieved after all these alignments were done properly.
118
APPENDIX E
Procedures of ECC image acquisition and data analysis
119
This section provides detailed information on the establishment of channeling
condition, obtain
dislocation contrast by following a Kikuchi band, and the identification of
channeling bands using T. O. C. A. software.
Establishing a channeling condition:
The fundamental mechanism for setting
-
up of an ECCI channeling condition is analogous
to the estab
lishing of two
-
beam diffraction condition to transmission electron microscopy
(TEM), by moving the optic axis/un
-
tilted electron beam (indicated by the black cross in
Figure
A10
) approaching the edge of the channeling band through proper tilt and rotation [141, 147].
It is noticeable that the deviation from the Bragg condition
s
= 0 is the imaging condition to get
maximum contrast in ECCI analysis [140] (as illustrate in
Figure
A10b
). It is more challenging to
correctly set up
s
= 0 channeling condition if the target area is highly textured or having an
orientation gradient due to higher strain level deformation, one have to manually adjust the
relevant position of the optic axi
s and the band edge for a best ECC image, which may need
several attempts. It should be mentioned that the orientation deviation after plastic
deformation can be solved by using a higher resolution SACP module that can collect the
Figure A
10
a) ECC image taken at s <0, the contrast of dislocations in the lower grain is not perfect. b) ECC
image taken at s = 0, a crisp image of sharp dislocations in pe
rfect contrast. c) ECC image taken at s > 0, the
dislocations in the same area are badly resolved with poor contrast.
120
accurate from smaller ar
eas (currently SACP module in MIRA III requires at least a 10 µm
diameter area in order to set up a channeling condition, and ~ 20 µm to get a perfect, un
-
overlapped pattern).
To get a crisp ECC image on a proper channeling condition, it is necessary to
take
multiple ECC images as the optic axis is traveling following the same channeling band as shown
in
Figure A11
. Some dislocations appear as dots with strong black contrast on one side and
white contrast on the other side, suggesting the dislocations are more inclined near the surface,
while some
dislocations appear as lines, indicating these dislocations are more parallel close to
the surface. It can be easily recognized that the dislocations labeled by small white arrows have
opposite black & white contrast with those marked by small black arrow
s, a reflection of the
opposite Burgers vectors. By comparing the same dislocations taken near different zone axis, it
Figure A1
1
a) The dislocation tails are in weak contrast, as
indicated by the small black arrows while some other
dislocation tails are in strong contrast, as indicated by the small white arrows when ECCI is taken close to the
upper zone axis. b) The dislocations indicated by the small black arrow are in strong c
ontrast, on the contrary,
the tails marker by the small white arrows are in weak contrast when the optic axis moves closer to the lower
zone axis.
121
is easy to find that dislocations labeled by the small black arrows have shorter tails with weaker
contrast in
Figure A11 a)
than in
b)
,
while the ones with opposite black & white contrast,
indicated by the white arrows, show slightly longer tails and better contrast in
a)
. This suggests
the dislocations with opposite Burgers vector also have opposite line directions or, dislocation
incli
nations, near the free surface. Nevertheless, the difference in length of the dislocation tails
with respect to the position on the channeling band also provides a clue to identify the near
-
surface dislocation line directions of these dislocations.
Identi
fication of the channeling band:
version No. 2.2, developed by Dr. Stefa
n Zaefferer in the year 2010. T. O. C. A. is specifically
used for the identification of channeling bands in both TEM with the interface shown in
Figure
A12
. The channeling band identification is done in the SEM mode, although the difference
between simu
lated patterns in TEM and SEM mode is minimal. The procedure for getting the
simulated pattern of a target crystal after loading the
-
Ti crystal dataset is as followed:
I. Input the Euler angle, which is acquired from the EBSD
-
-
a
module.
II. Manually rotate the crystal 90
o
along the Z
-
122
Figure A1
2
The interface of T. O. C. A. software shows 3 components in the simulation display, including the
simulated patterns shown on the left pop
-
up window, the corresponding pole figure shown on the upper
-
right
window, and the crystal orientation shown on the b
ottom
-
right window. The most useful parameters input in
the control panel are listed as: 1. The Euler angle; 2. The 90
o
rotation; 3. The acceleration voltage at which ECCI
is taken; 4. The magnification at which ECCI is taken.
123
IV. Input the magnificatio
V. (Optional) Adjust the width of the channeling band by changing the number in the
The simulated pattern should look similar to the real patterns collec
ted from the
channeling mode in MIRA III SEM, which is shown in
Figure A13
.
The dislocation Burgers vector
b
can be determined through
g
b
= 0 invisibility analyses as
g
is correctly labeled by T. O. C. A.
Figure A1
3
Left) Simulated channeling patterns with these bands labeled. Right) The real patterns collected
from channeling mode in MIRA III SEM, with the optic axis labeled as black cross.
124
APPENDIX F
Dislocation identifications
125
This section
only includes high quality ECC images that were not shown in the
manuscript, some rough (poor quality) ECC images that were used to quickly identify
Figure A14a) Grain 1 as deformed. b)
grain 1 after electropolishing. c
-
h) ECC images from the red box area taken
at different
g
vectors, dislocations are (1
)[11
0] and (1
)[11
0].
126
Figure A15a) Grain 2 as deformed. b) Grain 2 after electropolishing. c
-
h) ECC images from the red box
area taken
at different
g
vectors. Dislocations are
(1
)[11
0].
127
Figure A16a) Grain 3 as deformed. b) Grain 3 after electropolishing. c
-
h) ECC images from the rex boxed area at
different channeling conditions. Dislocations are
(10
0)[1
10].
128
Figure A17a) Grain 4 as deformed. b) Grain 4 after electropolishing.
c
-
h) Dislocations at the boxed area (grain
boundary between grains 3 and 4) taken at different channeling conditions. Dislocations are majority
(10
0)[
2
0].
129
Figure A18 One example of dislocations identification in sample 2 after electropolishing. a) ECC image of slip traces
after electropolishing. The contrast is not due to topography but from the contrast of dislocations. b
-
f)
Dislocations taken at differe
nt
g
vector. Dislocations are (10
0) [
2
0].
130
Figure A19 One example of dislocations identification in sample 2 after electropolishing on the other side of the
grain. Dislocations are (01
0) [
110].
131
Figure A20 Neighboring grains in Sample 2. Fir
st six ECC images) Identified dislocations are (1
00)[11
0]. Second
six ECC images) Identified dislocations are (1
00)[11
0] and (1
01)[11
0].
132
Figure A21 Neighboring grains in Sample 2. First six) Identified dislocations are (1
00)[
20] and (
101)[
11
0].
Second six) Identified dislocations are (1
00)[
20] and (1
01)[11
0].
133
Figure A22 Neighboring grains in Sample 2. First six) Identified dislocations are (10
0)[
0] and (
011)[1
10].
Second six) Identified dislocations are (0
10)[2
0]
and (01
1)[
110].
134
Figure A23 Neighboring grains in Sample 2. First six) Identified dislocations are (10
0)[
0] and (
011)[1
10].
Second six) Identified dislocations are (0
10)[2
0].
135
Figure A24 Neighboring grains in Sample 2. First six) Ident
ified dislocations are (10
0)[
0] and (
011)[1
10].
Second six) Identified dislocations are (0001)[
20].
136
Figure A25 Neighboring grains in Sample 2. First six) Identified dislocations are (1
00)[
20] and (1
01)[11
0].
Second six)
Identified dislocations are (
00)[
20].
137
Figure
A26 Neighboring grains in Sample 2. First six) Identified dislocations are (0
10)[
110]. Second six)
Identified dislocations are (
0)[
0].
138
Figure
A27 Neighboring grains in Sample 2. First six)
Identified dislocations are (1
00)[11
0]. Second six)
Identified dislocations are (
0)[
0].
139
Figure
A28 Neighboring grains in Sample 2. First six) Identified dislocations are (1
00)[11
0] and(1
01)[11
0].
Second six) Identified dislocations are
(
0)[
0].
140
Figure
A29 Neighboring grains in Sample 2. First six) Identified dislocations are (0001)[1
0] and(
011)[1
0].
Second six) Identified dislocations are (
0)[
0].
141
APPENDIX G
AFM, slip trace
analysis and ECCI of grains 3&5 in sample 1
142
Figure A30a) AFM color
-
scale topography map of grain 3 in sample 1. b) Slip systems identified by the trace
analysis are the (10
0)[
2
0] and (0
0)[2
0] prism slip systems. c) ECC image of the red box a
rea in b).
Contrast analysis also reveals two additional slip systems, which are (1
00)[
20] prism slip system and the
(0
11)[2
0] pyramidal slip system. The pyramidal dislocations are responsible for the curvy slip traces.
d) AFM color
-
sc
ale topography map of grain 5 in sample 1. e) (1
00)[11
0] prism slip system is identified by
the slip trace analysis. f) ECC images of the red box area in e). Contrast analysis also reveals a significant number
of (0
10)[2
0] prism dislocation
s within the observed area.
143
APPENDIX H
Calculation of the geometry of slip systems at a grain boundary
144
The correlation of
Figures 25 a and b
is used in this section as a particular example to
show how 3
-
D geometry of the slip systems at the grain boundary is revealed. With the 3
-
D
geometry, it is able to characterize the angle
between the intersection lines of slip planes on
both sides of t
he grain boundary with the grain boundary plane.
Figure A31
(left) is a sketch of
the geometry of the grain boundary plane and slip planes of the two interacting primary slip
systems. The upper right figure is the surface image which is associated with t
he upper plane in
the left sketch, and the lower right figure is corresponding to the subsurface image, which is
sketched as the bottom plane to the left figure. In order to correctly correlate the two images,
an arbitrary coordinate system is established
, with the z direction perpendicular to the surface,
x direction pointing down, and y direction pointing right on the surface. The origin O (0, 0, 0) is
set to the intersection between the two slip bands at the grain boundary, as shown in the upper
right
figure. The depth of material removal, d, is the vertical distance between the surface and
subsurface. It can be calculated from the electropolishing current and time, and this value can
be verified using microindentation removal measurements [96]. The
overall electropolished
surface is flat and smooth, with d varied between 4.8 and 5.2 µm across the examined area.
Thus, values of 4.8, 5.0, and 5.2 µm were used in the calculations to determine the variability of
the results.
On the subsurface, the gra
-
boundary traces in the subsurface since all the parameters can be calculated during the
145
construction of 3
-
D geometry. However, in this calculation, it is a
ssumed that the each of the
slip band is confined within its own slip plane, and there is no he
ight variation on the
subsurface at a given d. The calculation of the inclination of the grain boundary plane
and
more importantly,
, is carried out as follows:
The grain boundary trace on the electropolished surface can be expressed as a straight
li
ne:
y = k*x
Figure A3
1
3
-
D geometry of the well
-
correlated slip systems at the grain boundary by the correlation of surface
and subsurface images. (Amended from [171])
146
where k is the slope of the grain boundary trace (assume the trace is close to a straight
line). In some cases, the grain boundary line orientation may vary between the as
-
deformed
surface and
the electropolished surface. Thus, k is the averaged value between the k
surface
and
k
sub
-
surface
that are measured on each image.
-
b,
-
, 0,
-
d), are the intersection points of the grain boundary
1
, k*x
1
b,
-
2
, k*x
2
b,
-
d)
L =
|
|=
The directions of
and
are the respective intersections of the incoming (red)
and outgoing (yellow) slip planes with the grain boundary plane (
Figure A31
left). These will be
perpendicular to
the respective slip plane normal,
N
1
and
N
2
, which are readily available from
the MATLAB codes based on the crystal orientation information from the EBSD software:
N
1
N
2
Simultaneous solution of equations 1
-
4 allows determination of
and
. The
angle
between slip plane intersections on the grain boundary plane is then simply expressed
as:
= cos
-
1
(
Once
, and the corresponding grain boundary inclination angle,
, is given by:
147
= tan
-
1
(
)
The value of
reported in this study is
avg
, which is calculated on the basis of d = 5 µm
and k
average
. The true value of
has a variation range of ~3
o
to ~5
o
since different values of d =
4.8, 5.2 µm and k = k
surface
, k
subsurface
have been used. The uncertainty can be calculated by:
uncertain
=
max
min
148
Figure A32 All cases slip interactions at grain boundaries with calculated results.
149
APPENDIX I
Slip band broadening in grain 3 of sample 3
150
Figure A33 Slip band broadening effect observed along two different slip
bands.
151
BIBLIOGRAPHY
152
BIBILOGRAPHY
[1] K. Wang, The use of titanium for medical applications in the USA, Material Science and
Engineering: A, 213 (1
-
2), 1996, 134
-
137,
https://doi.org/10.1016/0921
-
5093(96)10243
-
4
.
[2] C. Leyens, M
. Peters, Titanium and titanium alloys: Fundamentals and Applications, Wein
-
heim: Wiley
-
VCH, 2003,
https://doi.org/10.1002/3527602119
.
[3] M. Geetha, A. K. Singh, R. Asokamani, A. K. Gogia, Ti based bioma
terials, the ultimate choice
for orthopaedic implants
A review, Progress in Material Science, 54 (3), 2009, 397
-
425,
https://doi.org/10.1016/j.pmatsci.2008.06.004
.
[4] D. Banerjee, J. C. Willi
ams, Perspectives on Titanium Science and Technology, Acta
Materialia, 61, 2013, 844
-
879,
https://doi.org/10.1016/j.actamat.2012.10.043
.
[5] M. R. Bache, W. J. Evans, Dwell Sensitive Fatigue Res
ponse of Titanium Alloys for Power
Plant Applications, Journal of Engineering for Gas Turbines and Power, 125 (1), 2003, 241
-
245,
https://doi.org/10.1115/1.1494094
.
[6] J. Hirsch, K. Lücke, Overview no. 76:
Mechanism of deformation and development of rolling
textures in polycrystalline f.c.c. metals
I. Description of rolling texture development in
homogeneous CuZn alloys, Acta Metallurgica, 36 (11), 1988, 2863
-
2882,
https://doi.org/10.1016/0001
-
6160(88)90172
-
1
.
[7] Y. Zhou, J. J. Jonas, J. Savoie, A. Makinde, S. R. MacEwen, Effect of texture on earing in FCC
metals: Finite element simulations, International Journal of Plasticity, 14 (1
3), 1998, 11
7
-
138,
https://doi.org/10.1016/S0749
-
6419(97)00044
-
2
.
[8] I. Samajdar, P. Ratchev, B. Verlinden, E. Aernoudt, Hot working of AA1050
relating the
microstructural and textural developments, Acta Materialia, 49 (10), 2001, 1759
-
1767,
https://doi.org/1
0.1016/S1359
-
6454(01)00083
-
0
.
[9] S. Suwas, R. K. Ray, A. K. Singh, S. Bhargava, Evolution of hot rolling textures in a two
-
phase
2
3
Al base alloy, Acat Materialia, 47 (18), 1999, 4585
-
4598,
https://doi.org/10.1016/S1359
-
6454(99)00327
-
4
.
[10] J. D. Picksley, F. R. Hall, I. Pillinger, Metal forming of titanium: Experimentation,
microstructure and computer simulation, Journal of Materials Processing Technology, 45 (1
-
4),
1994, 491
-
496,
https://doi.org/10.1016/0924
-
0136(94)90387
-
5
.
[11] Y. N. Wang, J. C. Huang, Texture analysis in hexagonal materials, Materials Chemistry and
Physics, 81 (1), 2003, 11
-
26,
https://doi.org/10.1016/S0254
-
0584(03)00168
-
8
.
153
[12] A. K. Singh, R. A. Schwarzer, Texture and Anisotropy of Mechanical Properties in Titanium
and Its Alloys, Zeitschrift f
r Metallkunde/Materials Research and Advanced Tec
hniques, 91(9),
2000, 702
-
716.
[13] H. Conrad, Effect of interstitial solutes on the strength and ductility of titanium, Progress in
Material Science, 26 (2
-
4), 1981, 123
-
403,
https://doi.org/1
0.1016/0079
-
6425(81)90001
-
3
.
[14] M. Ghazisaeidi, D. R. Trinkle, Interaction of oxygen interstitials with lattice faults in Ti, Acta
Materialia, 76 (1), 2014, 82
-
86,
https://doi.org/10.1016/j.ac
tamat.2014.05.025
.
[15] G. I. Taylor, Plastic strain in metals, Journal of the Institute of Metals, 62, 1938, 307
324,
http://pajarito.materials.cmu.edu/rollett/27750/Taylor_19
38.pdf
.
[16] P. V. Houtte, On the equivalence of the relaxed Taylor theory and the Bishop
-
Hill theory for
partially constrained plastic deformation of crystals, Materials Science and Engineering, 55 (1),
1982, 69
-
77,
https://doi.org/10.1016/0025
-
5416(82)90085
-
4
.
[17] Lei Bao, Contribution to the study of deformation twinning in titanium, Other, Université
Paul Verlaine
-
Metz, 2011, English, NNT: 2011METZ014Sff. fftel
-
01749050ff
[18] C. Zambaldi, Y. Yang,
T. R. Bieler, Orientation informed nanoindentation of a
-
titanium:
Indentation pileup in hexagonal metals deforming by prismatic slip, Journal of Materials
Research, 27 (1), 2011, 356
-
367,
https://doi.or
g/10.1557/jmr.2011.334
.
[19] P. G. Partridge, The crystallography and deformation modes of hexagonal close
-
packed
metals, Metallurgical Reviews, 12 (1), 1967, 169
-
194,
https://doi.org/10.1179/mtlr.
1967.12.1.169
.
[20] H. Numakura, M. Koiwa, Dislocations in metals and alloys with the hexagonal close
-
packed
structure, Metallurgical Science and Technology, 16 (1), 1998.
[21] M. P. Biget, G. Saada, Low
-
temperature plasticity of high
-
-
titanium sin
gle crystals,
Philosophical Magazine A, 59 (4), 1989, 747
-
757,
https://doi.org/10.1080/01418618908209818
.
[22] S. Naka, A. Lasalmonie, P. Costa, P. Kubin, L. P. Kubin, The low
-
temperature plastic
de
-
titanium and the core structure of a
-
type screw dislocations, Philosophical
Magazine A, 57 (5), 1988, 717
-
740,
https://doi.org/10.1080/01418618808209916
.
[23] S. Naka, L. P. Kubin, C.
Perrier, The plasticity of titanium at low and medium temperatures,
Philosophical Magazine A, 63 (5), 1991, 1035
-
1043,
https://doi.org/10.1080/01418619108213935
.
[24] F. Bridier, P. Villechaise, J.
Mendez, Analysis of the different slip systems activated by
Materialia, 53 (3), 2005, 555
-
567,
https://
doi.org/10.1016/j.actamat.2004.09.040
.
154
[25] J. C. Williams, R. G. Baggerly, N. E. Paton, Deformation behavior of HCP Ti
-
Al alloy single
crystals, Metallurgical and Materials Transactions A, 33, 2002, 837
-
850,
https://doi.org/10.1007/s11661
-
002
-
0153
-
y
.
[26] W. Hutchinson, M. Barnett, Effective values of critical resolved shear stress for slip in
polycrystalline magnesium and other hcp metals, Scripta Materialia, 63 (7), 2010, 737
-
740,
https://doi.org/10.1016/j.scriptamat.2010.05.047
.
[27] B. Barkia, V. Doquet, J. Couzinié, I. Guillot, E. Héripré, In situ monitoring of the deformation
mechanisms in titanium with different oxy
gen contents, Materials Science and Engineering: A,
636 (11), 2015, 91
102,
https://doi.org/10.1016/j.msea.2015.03.044
.
[28] H. Li, D. E. Mason, T. R. Bieler, C. J. Boehlert, M. A. Crimp, Methodolo
gy for estimating the
-
phase Ti using EBSD
-
based trace analysis, Acta
Materialia, 61 (20), 2013, 7555
7567,
https://doi.org/10.1016/j.actamat.2013.08.04
2
.
[29] J. Gong, A. J. Wilkinson, Micro
-
cantilever testing of prismatic slip in commercially pure
Ti, Philosophical Magazine, 91 (7
9), 2011, 1137
1149,
https://doi.org/10.1080/14786435.2010.
495359
.
[30] S. Zaefferer, A study of active deformation systems in titanium alloys: dependence on alloy
composition and correlation with deformation texture, Materials Science and Engineering: A,
344 (1
-
2), 2003, 20
-
30,
https://doi.org/10.1016/S0921
-
5093(02)00421
-
5
.
[31] A. Poty, J. M. Raulot, H. Xu, J. Bai, C. Schuman, J.
-
S. Lecomte, M.
-
J. Philippe, C. Esling,
Classification of the critical resolved shear stress in the hexagonalclose
-
packed mat
erials by
-
zirconium and
qf
-
titanium, Journal of Applied Physics, 110
(1), 2011, 014905,
https://doi.org/10.1063/1.3599870
.
[32] J. Gong, A. J. Wilkinson, A microcantilever
investigation of size effect, solid solution
strengthening and second
-
phase strengthening for prism slip in
qf
-
Ti, Acta Materialia, 59,
2011, 5970
5981,
https://doi.org/10.1016/j.actamat.201
1.06.005
.
[33] K. E. K. Amouzou, T. Richeton, A. Roth, M. A. Lebyodkin, T. A. Lebedkina, Micromechanical
-
titanium in tensile condition,
International Journal of Plasticity, 80, 2016, 222
-
240,
https://doi.org/10.1016/j.ijplas.2015.09.008
.
modelling of monotonic loading of CP a
-
Ti: Correlation between macroscop
ic and microscopic
-
Struct, 573, 2013, 222
233,
https://doi.org/10.1016/j.msea.2013.02.022
.
155
[35] J. Kacher, I. M. Robertson, In situ and tomographic
analysis of dislocation/grain boundary
-
titanium, Philosophical Magazine. 94, 2014, 814
-
829,
https://doi.org/10.1080/14786435.2013.868942
.
[36] J. L. W. Warwick, N. G. Jones, K.
M. Rahman, D. Dye, Lattice strain evolution during tensile
and compressive loading of CP Ti, Acta Materialia, 60 (19), 2012, 6720
-
6731,
https://doi.org/10.1016/j.actamat.2012.08.042
.
[37] D. Gl
oaguen, G. Oum, V. Legrand, J. Fajoui, S. Branchu, Experimental and theoretical
studies of intergranular strain in an alpha titanium alloy during plastic deformation, Acta
Materialia, 61 (15), 2013, 5779
-
5790,
https://doi.org/10.1016/j.actamat.2013.06.022
.
[38] A. Chakraborty, C. Zhang, S. Balachandran, T. R. Bieler, P. Eisenlohr, Assessment of surface
and bulk
-
dominated methodologies to measure critical resolved shear stresses in hexagonal
mate
rials, Acta Materialia, 184 (1), 2020, 241
-
253,
https://doi.org/10.1016/j.actamat.2019.11.023
.
[39] R. V. Mises, Mechanik der festen Körper im plastisch deformablen Zustand, Göttin.
Nachvodeg Ma
thematics, 1913, 582
-
592,
http://www.digizeitschriften.de/dms/resolveppn/?PID=GDZPPN002503697
.
[40] R. A. Lebensohn, C. N. Tomé, A self
-
consistent anisotropic approach for
the simulation of
plastic deformation and texture development of polycrystals: Application to zirconium alloys,
Acta Metallurgica, 41 (9), 1993, 2611
-
2624,
https://doi.org/10.1016/0956
-
7151(93)90
130
-
K
.
[41] I. Karaman, H. Sehitoglu, A. J. Beaudoin, Y. I. Chumlyakov, H. J. Maier, C. N. Tomé, Modeling
the deformation behavior of Hadfield steel single and polycrystals due to twinning and slip, Acta
Materialia, 48 (9), 2000, 2031
-
2047,
https://doi.org/10.1016/S1359
-
6454(00)00051
-
3
.
[42] M. J. Philippe, M. Serghat, P. V. Houtte, C. Esling, Modelling of texture evolution for
materials of hexagonal symmetry
Acta Metallurgica, 43 (4), 1995, 1619
-
1630,
https://doi.org/10.1016/0956
-
7151(94)00329
-
G
.
[43] Y. Yang, L. Wang, C. Zambaldi, P. Eisenlohr, R. Barabash, W. Liu, M. R. Stoudt, M. A. Crimp,
T. R. Bieler, Characterization and modeling of hetero
geneous deformation in commercial purity
titanium, The Journal of The Minerals, Metals & Materials Society (TMS), 63 (9), 2011, 66
-
73,
https://doi.org/10.1007/s11837
-
011
-
0161
-
8
.
[44] Y. Yang, L. Wan
g, T. R. Bieler, P. Eisenlohr, M. A. Crimp, Quantitative atomic force
microscopy characterization and crystal plasticity finite element modeling of heterogeneous
deformation in commercial purity titanium, Metallurgical and Materials Transactions A, 42,
201
1, 636
-
644,
https://doi.org/10.1007/s11661
-
010
-
0475
-
0
.
156
[45] T. R. Bieler, M. A. Crimp, L. Yang, L. Wang, P. Eisenlohr, D. E. Mason, W. Liu, G. E. Ice, Strain
heterogeneity and damage nucleation at g
rain boundaries during monotonic deformation in
commercial purity titanium, Journal of The Minerals, Metals & Materials Society (TMS), 61 (12),
2011, 45
-
52,
https://doi.org/10.1007/s11837
-
009
-
0180
-
x
.
[46] T. R. Bieler, P. Eisenlohr, F. Roters, D. Kumar, D. E. Mason, M. A. Crimp, D. a, The role of
heterogeneous deformation on damage nucleation at grain boundaries in single phase metals,
International Journal of Plasticity, 25 (9), 2009, 1655
-
1683,
https://doi.org/10.1016/j.ijplas.2008.09.002
.
[47] F. C. Frank, The Frank
-
Read source, Proceedings of the Royal Society A, 371 (1744), 1980,
136
-
138,
https://doi.org/10.1098/rspa.1980.0069
.
[48] F. C. Frank, W. T. Read Jr., Multiplication Processes for Slow Moving Dislocations, Physical
Review, 79 (4), 1950, 722
-
723,
https://doi.org/10.1103/
PhysRev.79.722
.
[49] (a) E. O. Hall, The deformation and ageing of mild steel: III discussion of results,
Proceedings of the Physical Society. Section B, 64 (9), 1951, 747
-
753,
https://doi.org/10.1
088/0370
-
1301/64/9/303
. (b) N. J. Petch, The cleavage strength of
polycrystals, The Journal of the Iron and Steel Institute, 173 (5), 1953, 25
28.
[50] V. Doquet, B. Barkia, Combined AFM, SEM and crystal plasticity analysis of grain boundary
sliding in ti
tanium at room temperature, Mechanics of Materials, 103, 2016, 18
-
27,
https://doi.org/10.1016/j.mechmat.2016.09.001
.
[51] T. Matsunaga, T. Kameyama, S. Ueda, E. Sato, Grain boundary sliding duri
ng ambient
-
temperature creep in hexagonal close
-
packed metals, Philosophical Magazine, 90 (30), 2010,
4041
-
4054,
https://doi.org/10.1080/14786435.2010.502883
.
[52] M. W. Grabski, R. Korski, Grain
boundaries as sinks for dislocations, The Philosophical
Magazine: A Journal of Theoretical Experimental and Applied Physics, 22 (178), 1970, 707
-
715,
https://doi.org/10.1080/14786437008220941
.
[53]
H. Gleiter, Grain boundaries as point defect sources or sinks
Diffusional creep, Acta
Metallurgica, 27 (2), 1979, 187
-
192,
https://doi.org/10.1016/0001
-
6160(79)90095
-
6
.
[54] A. Sarce, E. J. Savi
no, Grain Boundary Sink Strength in Anisotropic Materials, Physica Status
Solidi b, 159 (2), 1990, 557
-
570,
https://doi.org/10.1002/pssb.2221590205
.
[55] B. P. Uberuaga, L. J. Vernon, E. Martinez, A.
F. Voter, The relationship between grain
boundary structure, defect mobility and grain boundary sink efficiency, Scientific Reports, 5,
2015, 9095,
https://doi.org/10.1038/srep09095
.
[56] C. Goux, Structur
e des joints de grains: considérations cristallographiques et méthodes de
calcul des structures, The Canadian Journal of Metallurgy and Materials Science, 13 (1), 1974, 9
-
31,
https://doi.org/10.1179/c
mq.1974.13.1.9
.
157
[57] M. A. Fortes, Tilt and Twist Boundaries, Acta Crystallographica A, 29, 1973, 68
-
70,
https://doi.org/10.1107/S0567739473000148
.
[58] M. A. Tschopp, S. P. Coleman, D. L. McDowel
l, Symmetric and asymmetric tilt grain
boundary structure and energy in Cu and Al (and transferability to other fcc metals), Integrating
Materials and Manufacturing Innovation, 4, 2015, 176
-
189,
htt
ps://doi.org/10.1186/s40192
-
015
-
0040
-
1
.
[59] R. C. Pond, D. A. Smith,V. Vitek, A model for grain boundary structure based on random
close packing, Scripta Metallurgica, 12 (8), 1978, 699
-
702,
https://doi.org/10.1016/0036
-
9748(78)90311
-
3
.
[60] D. Smith, V. Vitek, R. Pond, Computer simulation of symmetrical high angle boundaries in
aluminium, Acta Metallurgica, 25 (5), 1977, 475
-
483,
ht
tps://doi.org/10.1016/0001
-
6160(77)90187
-
0
.
[61] K. L. Merkle, D. Wolf, Low
-
energy configurations of symmetric and asymmetric tilt grain
boundaries, Philosophical Magazine A, 65 (2), 1992, 513
-
530,
https://doi.org/10.1080/01418619208201536
.
[62] M. D. Sangid, T. Ezaz, H. Sehitoglu, I. M. Robertson, Energy of slip transmission and
nucleation at grain boundaries, Acta Materialia, 59 (1), 2011, 283
-
296,
https://doi.org/10.1016/j.actamat.2010.09.032
[63] R. J. Kurtz, R. G. Hoagland, J. P. Hirth,
Computer simulation of extrinsic grain
-
boundary
defects in the
11,
101
{131} symmetric tilt boundary
, Philosophical Magazine A, 79 (3),
1999, 683
-
70
3,
https://doi.org/10.1080/01418619908210325
.
[64] A. P. Sutton, V. Vitek, On the structure of tilt grain boundaries in cubic metals I.
Symmetrical tilt boundaries, Philosophical Transactions of th
e Royal Society of London, Series A:
Mathematical, Physical and Engineering Sciences, 309 (1506), 1983, 1
-
68,
https://doi.org/10.1098/rsta.1983.0020
.
[65] H. Grimmer, W. Bollmann, D. H. Warrington, Coi
ncidence
-
site lattices and complete
pattern
-
shift in cubic crystals, Acta Crystallographica Section A, 30 (2), 1974, 197
-
207,
https://doi.org/10.1107/S056773947400043X
.
[66] V. Thaveepringsriporn, G
. S. Was, The role of CSL boundaries in creep of Ni
-
16Cr
-
9Fe at
360°C, Metallurgical and Materials Transactions A, 28, 1997, 2101
-
2112,
https://doi.org/10.1007/s11661
-
997
-
0167
-
6
.
[67] V. Y. Gertsman
, S. M. Bruemmer, Study of grain boundary character along intergranular
stress corrosion crack paths in austenitic alloys, Acta Materialia, 49 (9), 2001, 1589
-
1598,
https://doi.org/10.1016/S1359
-
6454(01)00064
-
7
.
158
[68] K. Glowinski, A. Morawiec, Twist, tilt, and symmetric grain boundaries in hexagonal
materials, journal of Materials Science, 49, 2014, 3936
-
3942,
https://doi.org/10.1007/s108
53
-
013
-
7958
-
5
.
[69] P. Delavignette, T. Karakostas, G. Nouet, F. W. Schapink, Grain Boundary Orientation
Determination in Transmission Electron Microscopy, physica status solidi A, 107 (2), 1988, 551
-
558,
https://doi.org/10.1002/pssa.2211070209
.
boundaries, International Journal of Plasticity, 24 (2), 2008, 191
-
217,
https://doi.org/10.1016/j.ijplas.200
7.02.010
.
[71] E. Bayerschen, A. T. McBride, B. D. Reddy, T. Boehlke, Review on slip transmission criteria
in experiments and crystal plasticity models, Journal of Materials Science, 51, 2016, 2243
-
2258,
https://doi.org/10.1007/s10853
-
015
-
9553
-
4
.
[72] L. Priester, Grain Boundaries: From theory to engineering, Springer Series in Materials
Science, 2013, ISBN 978
-
94
-
007
-
4969
-
6
[73] S. Zaefferer, J.
-
C. Kuo, Z. Zhao, M. Winning, D. Raabe, On the influen
ce of the grain
boundary misorientation on the plastic deformation of aluminum bicrystals, Acta Materialia, 51
(16), 2003, 4719
-
4735,
https://doi.org/10.1016/S1359
-
6454(03)00259
-
3
.
[74] T. R. Bi
eler, P. Eisenlohr, C. Zhang, H. J. Phukan, M. A. Crimp, Grain boundaries and
interfaces in slip transfer, Current Opinion in Solid State and Materials Science, 18 (4), 2014,
212
-
226,
https://doi
.org/10.1016/j.cossms.2014.05.003
.
[75] Y. Guo, T. B. Britton, A. J. Wilkinson, Slip band
grain boundary interactions in commercial
-
purity titanium, Acta Materialia, 76 (1), 2014, 1
-
12,
https://
doi.org/10.1016/j.actamat.2014.05.015
.
[76] Z. Li, C. Hou, M. Huang, C. Ouyang, Strengthening mechanism in micro
-
polycrystals with
penetrable grain boundaries by discrete dislocation dynamics simulation and Hall
Petch effect,
Computational Materials Scien
ce, 46 (4), 2009, 1124
-
1134,
https://doi.org/10.1016/j.commatsci.2009.05.021
.
[77] Z. Shen, R. H. Wagoner, W. A. T. Clark, Dislocation and grain boundary interactions in
metals, Acta Metallurgica, 36, 1988, 3231
-
3242,
https://doi.org/10.1016/0001
-
6160(88)90058
-
2
.
[78] S.
Kondo, T. Mitsuma, N. Shibata, Y. Ikuhara, Direct observation of individual dislocation
interaction processes with grain boundaries, Science Advances, 2, 2016, e1501926,
https://doi.org/10.1126/sciadv.
1501926
.
159
[79] D. E. Spearot, M. D. Sangid, Insights on slip transmission at grain boundaries from atomistic
simulations, Current Opinion in Solid State and Materials Science, 18 (4), 2014, 188
-
195,
https://doi.org/10.1016/j.cossms.2014.04.001
.
[80] J. D. Livingston, B. Chalmers, Multiple slip in bicrystal deformation, Acta Metallurgica. 5,
1957, 322
327,
http://dx.doi.org/10.1016/0001
-
6160(57)90044
-
5
.
[81] Z. Shen, R. H. Wagoner, W. A. T. Clark, Dislocation pile
-
up at grain boundary interactions in
304 stainless steel, Scripta Materialia. 20, 1986, 921
926,
https://doi.org/10
.1016/0036
-
9748(86)90467
-
9
.
[82] Z. Shen, R. H. Wagoner, W. A. T. Clark, Dislocation and grain boundary interactions in
metals, Acta Metallurgica, 36, 1988, 3231
-
3242,
https://doi.org/10.1016/00
01
-
6160(88)90058
-
2
.
[83] T. C. Lee, I. M. Robertson, H. K. Birnbaum, Prediction of slip transfer mechanisms across
grain boundaries, Scripta Materialia. 23, 1989, 799
80,
https://doi.org/10.1016/
0036
-
9748(89)90534
-
6
.
[84] T. C. Lee, I. M. Robertson, H. K. Birnbaum, TEM in situ deformation study of the interaction
of lattice dislocations with grain boundaries in metals, Philosophical Magazine A. 62, 1990, 131
153,
http://dx.doi.org/10.1080/01418619008244340
.
[85] T. C. Lee, I. M. Robertson, H. K. Birnbaum, An in situ transmission electron microscope
deformation study of the slip transfer mechanisms in metals, Metallurgical Transactions A, 2
1,
1990, 2437
-
2447,
https://doi.org/10.1007/BF02646988
.
[86] T. C. Lee, I. M. Robertson, H. K. Birnbaum, Anomalous slip in an FCC system,
Ultramicroscopy, 29 (1
-
4), 1989, 212
-
216,
https://doi.org/10.1016/0304
-
3991(89)90248
-
9
.
[87] W. A. T. Clark, R. H. Wagoner, Z. Y. Shen, T. C. Lee, I. M. Robertson, H. K. Birnbaum, On the
criteria for slip transmission across interfaces in polycrystals, Scripta M
etallurgica, 26, 1992,
203
-
206,
https://doi.org/10.1016/0956
-
716X(92)90173
-
C
.
[88] R. S. Davis, R. L. Fleischer, J. D. Livingston, B. Chalmers, Effect of orientation on the plastic
deformation of
aluminum single crystals and bicrystals, The Journal of The Minerals, Metals &
Materials Society, 9, 1957, 136
-
140,
https://doi.org/10.1007/BF03398471
.
[89] R. C. Pond, W. A. T. Clark, R. H. Wagoner, On t
he crystallography of slip transmission in
hexagonal metals, Scripta Metallurgica, 20 (9), 1986, 1291
-
1295,
https://doi.org/10.1016/0036
-
9748(86)90051
-
7
.
[90] Th. Kehagias, Ph. Komninou, R. Grigo
riadis, G. Dimitrakopulos, J. G. Antonopoulos, Th.
Karakostas, Pyramidal slip in electron beam heated deformed titanium, Interface Science, 3
(3),1996, 195
-
201,
https://doi.org/10.1007/BF00191046
.
160
[91] Th.
Kehagias, Ph. Komninou, G. P. Dimitrakopulos, J. G. Antonopoulos, Th. Karakostas, Slip
transfer across low
-
angle grain boundaries of deformed Titanium, Scripta Metallurgica, 33 (12),
1995, 1883
-
1888,
https://doi.org/10.1016/0956
-
716X(95)00351
-
U
.
[92] J. Shirokoff, I. M. Robertson, H. K. Birnbaum, The Slip Transfer Process Through Grain
Boundaries in HCP Ti, Symposia CB
Defect
-
Interface Interactions, 319, 1993, 263,
https://doi.org/10.1557/PROC
-
319
-
263
.
[93] T. B. Britton, F. P. E. Dunne, A. J. Wilkinson, On the mechanistic basis of deformation at the
microscale in hexagonal close
-
packed metals, Proceedings of The Royal Society A Mathematical
P
hysical and Engineering Sciences, 471 (2178), 2015,
https://doi.org/10.1098/rspa.2014.0881
.
[94] T. B. Britton, A. J. Wilkinson, Stress fields and geometrically necessary dislocation density
distributi
ons near the head of a blocked slip band, Acta Materialia, 60 (16), 2012, 5773
-
5782,
https://doi.org/10.1016/j.actamat.2012.07.004
.
[95] J. Luster, M. A. Morris, Compatibility of deformation in two
-
phase Ti
-
Al alloys: Dependence
on microstructure and orientation relationships, Metallurgical and Materials Transactions A. 26,
1995, 17451756,
https://doi.org/10.1007/BF02670762
.
[96] S. Han, P. Eisenlohr, M. A. Crimp, ECCI based characterization of dislocation shear in
polycrystalline arrays during heterogeneous deformation of commercially pure titanium,
Materials Characterization. 142, 201
8, 504
-
514,
https://doi.org/10.1016/j.matchar.2018.06.003
.
[97] R. Alizadeh, M. Peña
-
Ortega, T. R. Bieler, J. LLorca, A criterion for slip transfer at grain
boundaries in Al, Scripta Materialia
, 178, 2020, 408
-
412,
https://doi.org/10.1016/j.scriptamat.2019.12.010
.
[98] T. R. Bieler, R. Alizadeh, M. Peña
-
Ortega, and J. LLorca, An analysis of (the lack of) slip
transfer between near
-
cube oriented grains in pure Al, International Journal of Plasticity, 118,
2019, 269
-
290,
https://doi.org/10.1016/j.ijplas.2019.02.014
.
[99] L. Nervo, A. King, A. Fitzner, W. Ludwig, M. Preuss, A study of deformation twinning in a
titanium alloy by X
-
ray diffraction
contrast tomography, Acta Materialia, 105, 417
-
428,
https://doi.org/10.1016/j.actamat.2015.12.032
.
[100] M. G. Wang, A. H.W. Ngan, Indentation strain burst phenomenon induced by grain
boundaries
in niobium, Journal of Materials Research, 19 (8), 2004, 2478
-
2486,
https://doi.org/10.1557/JMR.2004.0316
.
[101] E. Werner, W. Prantl, Slip transfer across grain and phase boundaries, Acta Metallurgica
,
38 (3), 1990, 533
-
537,
https://doi.org/10.1016/0956
-
7151(90)90159
-
E
.
161
[102] H. Wang, C. J. Boehlert, Q. D. Wang, D. D. Yin, W. J. Ding, In
-
situ analysis of the slip activity
during tensile defor
mation of cast and extruded Mg
-
10Gd
-
3Y
-
0.5Zr (wt.%) at 250 °C, Materials
Characterization, 116, 2016, 8
-
17,
https://doi.org/10.1016/j.matchar.2016.04.001
.
[103] T. Tsuru, Y. Shibutani, and T. Hi
rouchi, A predictive model for transferability of plastic
deformation through grain boundaries, American Institute of Physics (AIP) Advances, 6 (1),
2016, 015004,
https://doi.org/10.1063/1.4939819
.
[104] Y.
Su, S. Han, P. Eisonholr, M. A. Crimp, Slip trace accommodation at grain boundaries in
alpha titanium, In preparation.
[105] R. Ding, J. Gong, A. J. Wilkinson, I. P. Jones, A study of dislocation transmission through a
grain boundary in hcp Ti
-
6Al using m
icro
-
cantilevers, Acta Materialia, 103, 2016, 416
-
423,
https://doi.org/10.1016/j.actamat.2015.10.023
.
[106] J. A. Venables, C. J. Harland, Electron back
-
scattering patterns
A new technique for
o
btaining crystallographic information in the scanning electron microscope, The Philosophical
Magazine: A Journal of Theoretical Experimental and Applied Physics 8, 27 (5), 1973,
https://doi.org/10.1
080/14786437308225827
.
[107] A. J. Schwartz, M. Kumar, B. L. Adams, Electron backscatter diffraction in materials
science, Springer
-
Verlag US, Boston, MA, 2000,
https://doi.org/10.1007/978
-
1
-
4757
-
3
205
-
4
.
[108] M. N. Alam, M. Blackman, D. W. Pashley, High
-
angle Kikuchi patterns, Proceedings of the
Royal Society A Mathematical, Physical and Engineering Sciences, 221 (1145), 1954, 224
-
242,
https://
doi.org/10.1098/rspa.1954.0017
.
[109] F. J. Humphreys, Grain and subgrain characterisation by electron backscatter diffraction,
Journal of Material Science, 36, 2001, 3833
-
3854,
https://doi.org/10.102
3/A:1017973432592
.
[110] D. Dingley, Progressive steps in the development of electron backscatter diffraction and
orientation imaging microscopy, Journal of Microscopy, 213 (3), 2004, 214
-
224,
https://doi.org/10.1111/j.0022
-
2720.2004.01321.x
.
[111] S. Zaefferer, On the formation mechanisms, spatial resolution and intensity of backscatter
Kikuchi patterns, Ultramicroscopy, 107 (2
-
3), 2007, 254
-
266,
https://doi.org/10.1016/j.ultramic.2006.08.007
.
[112] S. Wright, M. Nowell, J. Basinger, Precision of EBSD based Orientation Measurements,
Microscopy and
Microanalysis, 17 ,2011, 406
407,
https://doi.org/10.1017/S143192761100290X
.
[113] K. Kunze, S. I. Wright, B. L. Adams, D. J. Dingley, Advances in automatic EBSP single
orientation measurements, Proceedings of the Symposium Microscale Textures of Material, 20,
1993, 41
-
54,
https://doi.org/10.1155/TSM.20.41
.
162
[114] N. C. K. Lassen, Source point calibration from an arbitrary electron backscattering pattern,
Journal of Microscopy, 195 (3), 1999, 204
-
211,
https://doi.org/10.1046/j.1365
-
2818.1999.00581.x
.
[115] J. Kacher, C. Landon, B. L. Adams, D. Fullwood, Bragg's Law diffraction simulations for
electron backscatter diffraction analysis, Ultramicroscopy, 109 (9), 2009,
1148
-
1156,
https://doi.org/10.1016/j.ultramic.2009.04.007
.
[116] J. Basinger, D. Fullwood, J. Kacher, B. Adams, Pattern center determination in electron
backscatter diffraction microscopy, Mic
roscopy and Microanalysis, 17 (3), 2011, 330
-
340,
https://doi.org/10.1017/S1431927611000389
.
[117] D. Fullwood, T. Ruggles, B. Adams, Mesoscopic Characterization of Heterogeneous Plastic
Deformation
, Reference Module in Materials Science and Materials Engineering, 2016,
https://doi.org/10.1016/B978
-
0
-
12
-
803581
-
8.02884
-
8
.
[118] K. Z. Troost, P. V. D. Sluis, D. J. Gravesteijn,
Microscal
e elastic
oÀ
strain determination by
backscatter Kikuchi diffraction in the scanning electron microscope
, Applied Physics Letters, 62
(10), 1993, 1110
-
1112,
https://doi.org/10.1063/1.108758
.
[119] A. J. Wilkins
on, G. Meaden, D. J. Dingley, High
-
resolution elastic strain measurement from
electron backscatter diffraction patterns: New levels of sensitivity, Ultramicroscopy, 106 (4
-
5),
2006, 307
313,
ht
tps://doi.org/10.1016/j.ultramic.2005.10.001
.
[120] A. J. Wilkinson, A new method for determining small misorientations from electron back
scatter diffraction patterns, Scripta Materialia, 44 (10), 2001, 2379
-
2385,
https://doi.org/10.1016/S1359
-
6462(01)00943
-
5
.
[121] D. Fullwood, M. Vaudin, C. Daniels, T. Ruggles, S. I. Wright, Validation of kinematically
simulated pattern HR
-
EBSD for measuring absolute strains and lattice tetragonality, Material
s
Characterization, 107, 2015, 270
-
277,
https://doi.org/10.1016/j.matchar.2015.07.017
.
[122] A. J. Wilkinson, T. B. Britton, J. Jiang, P. S. Karamched, A review of advances and
challenges in EBS
D strain mapping, IOP Conf. Series: Materials Science and Engineering, 55,
2014, 012020,
http://dx.doi.org/10.1088/1757
-
899X/55/1/012020
.
[123] A. Winkelmann,
Principles of depth
oÀ
resolved Kiku
chi pattern simulation for electron
backscatter diffraction
, Journal of Microscopy, 239 (1), 2010, 32
-
45,
https://doi.org/10.1111/j.1365
-
2818.2009.03353.x
.
[124] P. B. Hirsch, M. J. Whelan, A
kinematical theory of diffraction contrast of electron
transmission microscope images of dislocations and other defects, Philosophical Transactions of
the Royal Society of London. Series A, Mathematical and Physical Sciences, 252 (1070), 1960,
499
529,
https://doi.org/10.1098/rsta.1960.0013
.
163
[125] A. Howie, M. J. Whelan, Diffraction contrast of electron microscope images of crystal
lattice defects. III. Results and experimental confirmation of the dyna
mical theory of dislocation
image contrast, Philosophical Transactions of the Royal Society of London. Series A,
Mathematical and Physical Sciences, 267 (1329), 1962, 206
-
230,
https://doi.org/10.1098/r
spa.1962.0093
.
[126] P. B. Hirsch, R. W. Horne, M. J. Whelan, LXVIII. Direct observations of the arrangement
and motion of dislocations in aluminum, Philosophical Magazine, 1 (7), 1956, 677
-
684,
https://doi.org/1
0.1080/14786435608244003
.
[127] M. Knoll, J. Kügler, Subjective light pattern spectroscopy in the encephalographic
frequency range, Nature (London), 184, 1959, 1823
1824,
https://doi.org/10.1038/1841823a0
.
[128] R. Erni, M. D. Rossell, C. Kisielowski, U. Dahmen, Atomic
-
resolution imaging with a sub
-
50
-
pm electron probe, Physical Review Letters, 102 (9), 2009,
https://doi.org/10.1103/PhysRevLett.1
02.096101
.
[129] C. Kisielowski, B. Freitag, M. Bischoff, H. van Lin, S. Lazar, G. Knippels, P. Tiemeijer, M. van
der Stam, S. von Harrach, M. Stekelenburg, M. Haider, S. Uhlemann, H. Müller, P. Hartel, B.
Kabius, D. Miller, I. Petrov, E. A. Olson, T. Donchev, E. A.
Kenik, A. R. Lupini, J. Bentley, S. J.
Pennycook, I. M. Anderson, A. M. Minor, A. K. Schmid, T. Duden, V. Radmilovic, Q. M. Ramasse,
M. Watanabe, R. Erni, E. A. Stach, P. Denes, U. Dahmen, Detection of single atoms and buried
defects in three dimensions by
aberration
-
corrected electron microscope with 0.5
-
Å
information limit, Microscopy and Microanalysis, 14 (5), 2008, 469
-
477,
https://doi.org/10.1017/S1431927608080902
.
[130]
M. H. F. Overwijk, F. C.
van den Heuvel, C. W. T. Bulle
oÀ
Lieuwma
, Novel scheme for the
preparation of transmission electron microscopy specimens with a focused ion beam, Journal of
Vacuum Science & Technology B: Microelectronics and Nanometer Structures Processing,
Measurement, an
d Phenomena, 11 (6), 1993, 2021
-
2024,
https://doi.org/10.1116/1.586537
.
[131] F. A. Stevie, T. C. Shane, P. M. Kahora, R. Hull, D. Bahnck, V. C. Kannan, E. David,
Applications of focused ion beams in microe
lectronics production, design and development,
Surface and Interface Analysis, 23 (2), 1995, 61
-
68,
https://doi.org/10.1002/sia.740230204
.
[132] L. A. Giannuzzi, F. A. Stevie, A review of focused ion be
am milling techniques for TEM
specimen preparation, Micron, 30 (3), 1999, 197
-
204,
https://doi.org/10.1016/S0968
-
4328(99)00005
-
0
.
[133] A. J. Leslie, K. L. Pey, K. S. Sim, M. T. F. Beh, G. P. Go
h, TEM sample preparation using FIB:
Practical problems and artifacts, 21
st
International Symposium for Testing and Failure Analysis,
ASM International, 1995, 53,
https://www.osti.gov/servlets/purl
/1426988
.
164
artefacts induced by FIB in austenitic stainless steel, Journal of Microscopy, 237 (3), 2010, 439
-
442,
https://doi.org/10.1111/j.1365
-
2818.2009.03288.x
.
[135] F. Lagattu, F. Bridier, P. Villechaise, J. Brillaud, In
-
plane strain measurements on a
microscopic scale by coupling digital image correlation and an in situ SEM technique, Materia
ls
Characterization, 56 (1), 2006, 10
-
18,
https://doi.org/10.1016/j.matchar.2005.08.004
.
[136] T. J. Ruggles, D. T. Fullwood, Est
imations of bulk geometrically necessary dislocation
density using high resolution EBSD, Ultramicroscopy, 133, 2013, 8
-
15,
https://doi.org/10.1016/j.ultramic.2013.04.011
.
[137] B. E. Dunlap, T.
J. Ruggles, D. T. Fullwood, B. Jackson, M. A. Crimp, Comparison of
dislocation characterization by electron channeling contrast imaging and cross
-
correlation
electron backscattered diffraction, Ultramicroscopy, 184, 2018, 125
-
133,
https://doi.org/10.1016/j.ultramic.2017.08.017
.
[138] T. J. Ruggles, Y. S. J. Yoo, B. E. Dunlap, M. A. Crimp, J. Kacher, Correlating results from high
resolution EBSD with TEM
-
and ECCI
-
based dislocation microscopy: A
pproaching single
dislocation sensitivity via noise reduction, Ultramicroscopy, 210, 2020, 112927,
https://doi.org/10.1016/j.ultramic.2019.112927
.
[139] P. Morin, M. Pitaval, D. Besnard, G. Fon
taine, Electron
-
channelling imaging in scanning
electronmicroscopy, Philosophical Magazine A, 40(4), 1979, 511
-
524,
https://doi.org/10.1080/01418617908234856
.
[140] B. A. Simkin, M. A. Crimp, An exp
channeling contrast imaging, Ultramicroscopy, 77 (1
-
2), 1999, 65
-
75,
http://dx.doi.org/10.1016/S03043991(99)00009
-
1
.
[141] M. A. Crimp, Scanning
electron microscopy imaging of dislocations in bulk materials, using
electron channeling contrast, Microscopy Research and Technique, 69 (5), 2006, 374
-
381,
https://doi.org/10.1002/jemt.20293
.
[142] S. Za
-
N. Elhami, Theory and application of electron channeling contrast imaging
-
50,
http://dx.doi.org/10.1016/j.actam
at.2014.04.018
.
84
-
85, 2014, 11
-
14,
http://dx.doi.org/10.1016/j.scriptamat.2014.03.001
.
[144] M. A. Crimp, B. Simkin, B. Ng, Demonstration of the g · b x u = 0 edge dislocation
invisibility criterion for electron channeling contrast imaging, Philosophical Magazine Letters, 81
(12), 2001, 833
-
837,
https://doi.org/10.1080/09500830110088755
.
165
[145] Y. N. Picard, M. E. Twigg, J. D. Caldwell, C. R. Eddy Jr., M. A. Mastro, R. T. Holm, Resolving
the Burgers vector for individual GaN dislocations by electron channeling contrast imaging,
Scripta
Materialia, 61 (8), 2009, 773
-
776,
https://doi.org/10.1016/j.scriptamat.2009.06.021
.
[146] D. C. Joy, D. E. Newbury, D. L. Davidson, Electron channeling patterns in the scanning
electron micr
oscope, Journal of Applied Physics, 53 (8), 1982, 81
-
122,
https://doi.org/10.1063/1.331668
.
[147] J. Guyon, H. Mansour, N. Gey, M. A. Crimp, S. Chalal, N. Maloufi, Sub
-
micron resolution
selected area electro
n channeling patterns, Ultramicroscopy. 149 (2015) 34
44.
https://doi.org/10.1016/j.ultramic.2014.11.004
.
[
148
] R. D. Kerns, S. Balachandran, A. H. Hunter, M. A. Crimp, Ultra
-
high spatial reso
lution
selected area electron channeling patterns, Ultramicroscopy, 210, 2020, 112915,
https://doi.org/10.1016/j.ultramic.2019.112915
.
[149] I. G. Urrutia, S. Zaefferer, D. Raabe, Electron chan
neling contrast imaging of twins and
dislocations in twinning
-
induced plasticity steels under controlled diffraction conditions in a
scanning electron microscope, Scripta Materialia, 61 (7), 2009, 737
-
740,
https://doi.org/10.1016/j.scriptamat.2009.06.018
.
[150] J. T. Czernuszka, N. J. Long, E. D. Boyes, P. B. Hirsch, Imaging of dislocations using
backscattered electrons in a scanning electron microscope, 62 (4), 1990, 227
-
232,
https://doi.org/10.1080/09500839008215127
.
[151] A. J. Wilkinson, G. R. Anstis, J. T. Czernuszka, N. J. Long, P. B. Hirsch, Electron channeling
contrast imaging of interfacial defects in strained silicon
-
germanium
layers on silicon,
Philosophical Magazine A, 68 (1), 1993, 59
-
80,
https://doi.org/10.1080/01418619308219357
.
[152] L. Han, B. Gamm, Optimizing electron channeling contrast imaging condition in scan
ning
electron microscope, European Microscopy Congress 2016: Proceedings,
https://doi.org/10.1002/9783527808465.EMC2016.6188
.
[153] H. Kriaa, A. Guitton, N. Maloufi, Fundamental and experimental aspects of diffraction for
characterizing dislocations by electron channeling contrast imaging in scanning electron
microscope, Scientific Report, 7, 2017, 9742,
https://doi.org/10.1038/s41598
-
017
-
09756
-
3
.
[154] P. Morin, M. Pitaval, D. Besnard, G. Fontaine, Electron
-
channeling imaging in scanning
electron microscopy, Philosophical Magazine A, 40 (4), 1979, 511
-
524,
https://doi.org/10.1080/01418617908234856
.
[155] M. A. Sutton, W. J. Wolters, W. H. Peters, W. F. Ranson, S. R. McNeill, Determination of
displacements using an improved digital correlation method, Image and Visio
n Computing, 1
(3), 1983, 133
-
139,
https://doi.org/10.1016/0262
-
8856(83)90064
-
1
.
166
[156] H. A. Bruck, S. R. McNeill, M. A. Sutton, W. H. Peters III, Digital image correlation using
Newton
-
Raphson m
ethod of partial differential correction, Experimental Mechanics, 29, 1989,
261
-
267,
https://doi.org/10.1007/BF02321405
.
[157] Y. L. Loo, R. L. Willett, K. W. Baldwin, J. A. Rogers, Additive, nanoscale pat
terning of metal
films with a stamp and a surface chemistry mediated transfer process: Applications in plastic
electronics, Applied Physics Letters, 81 (3), 2002, 562,
https://doi.org/10.1063/1.1493226
.
[15
8] M. A. Sutton, N. Li, D. C. Joy, A. P. Reynolds, X. Li, Scanning electron microscopy for
quantitative small and large deformation measurements part i: sem imaging at magnifications
from 200 to 10,000, Experimental Mechanics, 47, 2007, 775
-
787,
https://doi.org/10.1007/s11340
-
007
-
9042
-
z
.
[159] W. A. Scrivens, Y. Luo, M. A. Sutton, S. A. Collette, M. L. Myrick, P. Miney, P. E. Colavita, A.
P. Reynolds, X. Li, Development of patterns for digital image
correlation measurements at
reduced length scales, Experimental Mechanics, 47, 2007, 63
-
77,
https://doi.org/10.1007/s11340
-
006
-
5869
-
y
.
[160] M. Fan, A. G. Brolo, Silver nanoparticles self
-
assembly as SERS substrates with near single
molecule detection limit, Physical Chemistry Chemical Physics, 11 (34), 2009, 7381
-
7389,
https://doi.org/10.1
039/B904744A
.
[161] I. Liascukiene, N. Aissaoui, S. J. Asadauskas, J. Landoulsi, J.
-
F. Lambert, Ordered
nanostructures on a hydroxylated aluminum surface through the self
-
assembly of fatty acids,
Langmuir, 28 (11), 2012, 5116
-
5124,
https://doi.org/10.1021/la2051542
.
[162] A. D. Kammers, S. Daly, Self
-
assembled nanoparticle surface patterning for improved
digital image correlation in a scanning electron microscope, Experimental Mechanics, 53, 2013,
1333
-
1341
,
https://doi.org/10.1007/s11340
-
013
-
9734
-
5
.
[163] Z. Chen, W. Lenthe, J. C. Stinville, M. Echlin, T. M. Pollock, S. Daly, High
-
resolution
deformation mapping across large fields of view using scan
ning electron microscopy and digital
image correlation, Experimental Mechanics, 58, 2018, 1407
-
1421,
https://doi.org/10.1007/s11340
-
018
-
0419
-
y
.
[164] F. Ohnesorge, G. Binnig, True atomic resolution
by atomic force microscopy through
repulsive and attractive forces, Science, 260 (5113), 1993, 1451
-
1456,
https://doi.org/10.1126/science.260.5113.1451
.
[165] F. J. Giessibl, Atomic resolution o
f the silicon (111)
-
(7x7) surface by atomic force
microscopy, Science, 267 (5194), 1995, 68
-
71,
https://doi.org/10.1126/science.267.5194.68
.
e artifacts, Applied Surface Science, 304,
2014, 11
-
19,
https://doi.org/10.1016/j.apsusc.2014.01.149
.
167
[167] B. Voigtländer, Artifacts in AFM. In: Atomic Force Microscopy, NanoScience and
Techno
logy, Springer, Cham, 2019,
https://doi.org/10.1007/978
-
3
-
030
-
13654
-
3_8
.
[168] T. Strahlendorff, G. Dai, D. Bergmann, R. Tutscha, Tip wear and tip breakage in high
-
speed
atomic force microscopes,
Ultramicroscopy, 201, 2019, 28
-
37,
https://doi.org/10.1016/j.ultramic.2019.03.013
.
[169] C. Blochwitz, J. Brechbühl, W. Tirschler, Analysis of activated slip systems in fatigue nickel
polycryst
als using the EBSD
-
technique in the scanning electron microscope, Materials Science
and Engineering: A, 210 (1
-
2), 1996, 42
-
47,
https://doi.org/10.1016/0921
-
5093(95)10076
-
8
.
[170] J. Veselý, M. C
ieslar, C. Coupeau, J. Bonneville, Quantitative numerical method for
analysing slip traces observed by AFM, Surface Topography: Metrology and Properties, 1, 2013,
015002,
http://iopscience.iop.
org/2051
-
672X/1/1/015002
.
[171] S. Han, M. A. Crimp, ECCI analysis of shear accommodations at grain boundaries in
commercially pure alpha titanium, International Journal of Plasticity, in press, 2020,
https://doi.org/10.1016/j.ijplas.2020.102731
.
[172] Y. Su, C. Zambaldi, D. Mercier, P. Eisenlohr, T. R. Bieler, M. A. Crimp, Quantifying
deformation processes near grain boundaries in a titanium using nanoindentation and crystal
plasticity modeling,
International Journal of Plasticity, 86, 2016, 170
-
186,
https://doi.org/10.1016/j.ijplas.2016.08.007
.
[173] D. E. J Armstrong, A. J. Wilkinson, S. G. Roberts, Measuring local mechanical propertie
s
using FIB machined microcantilevers, Symposium II
Probing Mechanics at Nanoscale
Dimensions, 1185, 2009, 1185
-
I102
-
08,
https://doi.org/10.1557/PROC
-
1185
-
II02
-
08
.
[174] D. E. J Armstrong, A. J. W
ilkinson, S. G. Roberts, Micro
-
mechanical measurements of
fracture toughness of bismuth embrittled copper grain boundaries, Philosophical Magazine, 92,
2012, 394
-
400,
https://doi.org/10.1080/0950
0839.2011.573813
.
[175] G. Thomas, M. J. Goringe, Transmission Electron Microscopy of Materials, John Wiley &
Sons, New York, 1979.
[176] I. Epelboin, Amincissement électrolytique très poussé et ses applications, Rev. Met. Paris,
49 12 (1952) 863
-
875,
https://doi.org/10.1051/metal/195249120863
.
[177] P. A. Jacquet, On the anodic behavior of copper in aqueous solutions of orthophosphoric
acid, Trans. Electrochem. Soc., 1936, 69, 629
655.
https://doi.org/10.1149/1.3498234
.
[178] G. Yang, B. Wang, K. Tawfiq, H. Wei, S. Zhou, G. Chen, Electropolishing of surfaces: theory
and applications, Surface Engineering, 2017, 33:2, 149
-
166,
https://doi.org/10.1080/02670844.2016.1198452
.
168
[179] A. Chandra, On the Mechanism of Niobium Electropolishing, Thesis, 2012, Ohio State
University
[180] H. Oguri, S. Hishiyama, T. Oishi, M. Hirama
, Enantio
-
controlled synthesis of the ab ring
moiety of ciguatoxin, Synlett, 12, 1995, 1252
-
1254,
https://doi.org/10.1055/s
-
1995
-
5259
.
[181] E. J. Corey, A. Venkateswarlu, Protection of hydroxyl groups as tert
-
butyldimethylsilyl
derivatives, Journal of the American Society, 94 (17), 1972, 6190
-
6191,
https://doi.org/10.1021/ja00772a043
.
[
182] S. Higashibayashi, K. Shinko, T. Ishizu, K. Hashimoto, H. Shirahama, Selective deprotection
of t
-
butyldiphenylsilyl ethers in the presence of t
-
butyldimethylsilyl ethers by
tetrabutylammonium fluoride, acetic acid, and water, Synlett, 9, 2000, 1306
-
13
08,
https://doi.org/10.1055/s
-
2000
-
7158
.
[183] B. F. Bonini, M. C. Franchini, M. Fochi, G. Mazzanti, A. Ricci, Desilylation of (Z)
-
á
-
dimethylphenylsilyl vinyl sulfides with fluorideion: revised mechanism
for phenyl group
migration in substrates containing an electron
-
Chemical Society, Perkin Transactions 1, 21, 1997, 3211
-
3217,
https://doi.org/10.1039/A703662K
.
[
184] T. W. Greene, P. G. M. Wuts, Protective groups in organic synthesis, 3rd ed.; John Wiley &
Sons: New York, 1999,
https://doi.org/10.1002/0471220574
.
[185] R. Sperry, S. Han, Z. Chen, S. H. Daly, M. A.
Crimp, D. Fullwood, Comprehensive study and
comparison of various deformation measurement methods in Ti7Al alloy, in preparation.
[186] Z. Chen, S. H. Daly, Active Slip System Identification in Polycrystalline Metals by Digital
Image Correlation (DIC), Ex
perimental Mechanics 57 (1), 2017, 115
-
27.
https://doi.org/10.1007/s11340
-
016
-
0217
-
3
.
[187] GitHub
-
BYU
-
MicrostructureOfMaterials/OpenXY,
https://github.com/BYUMicrostructureOfMaterials/OpenXY
.
[188] R. Sperry, A. Harte, J. Q. da Fonseca, E. R. Homer, R. H. Wagoner, D. T. Fullwood, Slip band
characteristics in the presence of grain boundaries in nickel
-
based superalloy, Acta Mate
rialia,
193, 2020, 229
-
238,
https://doi.org/10.1016/j.actamat.2020.04.037
.
[189] F. D. Gioacchino, T. E. J. Edwards, G. N. Wells, W. J. Clegg, A new mechanism of strain
transfer in polycrystals,
Scientific Reports, 10, 2020, 10082,
https://doi.org/10.1038/s41598
-
020
-
66569
-
7
.
[190] L. E. Murr, Dislocation ledge sources: dispelling the myth of Frank
Read source
importance, Metallurgical and
Materials Transactions A, 47, 2016, 5811
5826,
https://doi.org/10.1007/s11661
-
015
-
3286
-
5
.
169
[191] T. L. Burnett, R. Kelley, B. Winiarski, L. Contreras, M. Daly, A. Gholinia, M. G. Burke, P. J.
Wither
s, Large volume serial section tomography by Xe Plasma FIB dual beam microscopy,
Ultramicroscopy, 161, 2016, 119
-
129,
https://doi.org/10.1016/j.ultramic.2015.11.001
.
[192] L. F. Allard, D. N. L
eonard, J. D. Poplawsky, M. F. Chisholm, B. D. Eckhart, A. Shyam, F. S.
Walden, B. B. Larson, R. Kelly, A. Stokes, W. C. Bigelow, The Utility of Xe
-
Plasma FIB for
Preparing Aluminum Alloy Specimens for MEMS
-
based In Situ Double
-
Tilt Heating Experiments,
Mi
croscopy and Microanalysis, 25, 2019, 1442
-
1443,
https://doi.org/10.1017/S1431927619007943
.