You are on page 1of 7

Materials Science & Engineering A 674 (2016) 171–177

Contents lists available at ScienceDirect

Materials Science & Engineering A


journal homepage: www.elsevier.com/locate/msea

Spatial correlation between local misorientations and nanoindentation


hardness in nickel-base alloy 690
Rickard R. Shen a,n, Valter Ström b, Pål Efsing a,c
a
Department of Solid Mechanics, KTH Royal Institute of Technology, Teknikringen 8D, Stockholm SE-100 44, Sweden
b
Department of Materials Science, KTH Royal Institute of Technology, Brinellvägen 23, Stockholm SE-100 44, Sweden
c
Ringhals AB, Ringhalsverket, Väröbacka SE-432 85, Sweden

art ic l e i nf o a b s t r a c t

Article history: Misorientation increases with plastic strain in metals, and this observation has been used as an empirical
Received 3 May 2016 assessment of plastic strain in recent years. The method has been validated for a sample area corre-
Received in revised form sponding to a 100 mm  100 mm square, but on the micrometer scale misorientations no longer seem to
13 July 2016
correlate with plastic strain. Misorientations are however not dependent on plastic strain but rather on
Accepted 29 July 2016
dislocation density, which means it should also be related to hardness. Therefore, we have in this work
Available online 30 July 2016
compared maps of predicted hardness calculated from misorientation determination with maps of actual
Keywords: hardness measured by nanoindentation.
EBSD It was shown that the predicted and measured hardness maps do indeed correlate spatially in nickel-
Nanoindentation
base Alloy 690, although the measured values have a significantly smaller hardness variation. This is
Local misorientation
explained by a presumably high and uniform density of statistically stored dislocations, which contribute
Hardening
Plasticity to hardness but do not affect the misorientation determination from electron backscatter diffraction.
Nickel based superalloys Thus local misorientation can be used to qualitatively map the local effective plastic strain distribution,
for example to identify regions of increased hardness.
& 2016 Elsevier B.V. All rights reserved.

1. Introduction image correlation (DIC) have been performed, although with un-
satisfactory results in copper [14,15] and aluminum [16].
In recent years electron backscatter diffraction (EBSD) has been Sasaki et al., however, discussed that DIC measures the actual
used to characterize changes in polycrystalline metals that result strain, while misorientations between neighboring points reflect
from plastic deformation [1–12]. Specifically the change in mis- the local dislocation density, and that these two quantities do not
orientations, i.e. the change in crystal lattice orientation between necessarily correlate on a micrometer scale [15]. They illustrated
different points in the material, has become popular to focus on. A that plastic deformation can accumulate where dislocations had
scalar misorientation quantity can be defined in different ways, passed through, but that misorientations only increased where
and the most common methods have been summarized in litera- dislocations accumulated. Misorientations may thus be closer re-
lated to effective plastic strain, which is a measure of strain hard-
ture [13]. No matter how misorientations are defined though,
ening, i.e. dislocation density buildup, rather than plastic strain,
when averaged over a larger area they tend to increase with
which is a measure of geometric shape change.
average plastic strain, and this empirical observation has allowed
Polycrystalline metals are plastically inhomogeneous at the
the use of EBSD for plastic strain estimation.
microscopic size scale, thus an inhomogeneous hardness dis-
This misorientation based plastic strain assessment technique
tribution can be expected as the material deforms. As such, maps
has been validated for quantification of plastic strain using a
of local hardness would be relevant for processing and monitoring
sample area corresponding to roughly a 100 mm  100 mm square degradation of materials where local material properties are im-
[1,2]. Spatial variations in local misorientation values can be seen portant, e.g. recrystallization, stress corrosion cracking (SCC) and
on a micrometer scale as well, and attempts at correlating these fatigue.
local misorientations with strain measurements based on digital If misorientations do indeed correlate with effective plastic
strain, it can be used as a tool to predict the hardness profile.
n
Corresponding author.
Therefore, nanoindentation hardness mapping has been used in
E-mail addresses: rshen@kth.se (R.R. Shen), valter@kth.se (V. Ström), this work as a means to investigate the correlation with those
pal.efsing@vattenfall.com (P. Efsing). predicted by misorientation mapping.

http://dx.doi.org/10.1016/j.msea.2016.07.123
0921-5093/& 2016 Elsevier B.V. All rights reserved.
172 R.R. Shen et al. / Materials Science & Engineering A 674 (2016) 171–177

An inherent uncertainty with the misorientation technique is ε = ln( L/L 0), (1)
that misorientations are in principle only dependent on the in-
homogeneous straining, which induces geometrically necessary where L0 was the length of a region before deformation, and L is
dislocations (GNDs), whereas the total dislocation density also the current length. The specimens were unloaded before failure
consists of statistically stored dislocations (SSDs), which are in- and retained a plastic strain, εp, of 0.029, 0.062, 0.087, 0.116 and
duced by homogeneous straining [17]. Since homogeneous strains 0.144 respectively. The heat treated material exhibited a yield
do not contribute to misorientations, but do contribute to strain strength of 273 MPa.
hardening, misorientation distribution does not necessarily reflect
the local hardness distribution accurately on the micrometer size 2.2. Prediction of local effective plastic strains
scale, which necessitates a clarification of the actual relation be-
tween misorientation and hardness. Material from both the unstrained and strained specimens was
In practice, strains on a larger scale are often estimated from used to prepare specimens for EBSD. The material was carefully
misorientation maps (acquired with EBSD), which suggests that a sectioned along a plane normal to the tube's axial direction using a
verified extension of this method to the micrometer scale would Struers Accutom-5 precision cut-off machine. The sampled mate-
constitute a significant advance. rial was mounted in conductive resin and ground using SiC paper,
In this work we have shown the existence of a correlation be- followed by polishing using diamond paste. Final polishing was
tween local hardness and local kernel average misorientation performed using 0.06 mm colloidal silica.
(KAM) which appears useful, but since deformation mechanisms An EBSD system from HKL Technology was used within a LEO
can differ between materials, and even for the same material if 1530 Gemini field emission scanning electron microscope (SEM) to
deformation temperature or rate is significantly changed, we obtain crystal orientation maps on the sampled materials. The
cannot a priori infer that the findings in this work are generally accelerating voltage was set to 15 kV, and the aperture was opened
applicable. to 120 mm. The working distance differed slightly between sam-
ples, but was between 8–10 mm. On each specimen, an area of
610 mm  460 mm was mapped using a square grid with step
length 0.9 mm. HKL software was used for acquisition of the or-
2. Experimental
ientation maps and phase identification, but in-house scripts were
used for post-processing of the orientation data. The indexing
2.1. Test material
rates for the specimens were between 98.0–99.9%, and no cleanup
of non-indexed points was performed.
Alloy 690, a solution strengthened high chromium nickel-base
An orientation smoothening filter proposed by Kamaya was
alloy with a face-centered cubic (FCC) crystal structure, was used
applied to each orientation map [19]. This filter is essentially a
in this work. It is highly ductile, has significant strain hardening,
grain-wise moving average of the orientation, and tends to in-
and deforms by dislocation slip at the temperature and strain rates
crease the precision in the orientation map at the cost of a slight
involved in this work.
loss of spatial resolution. A square kernel of 3  3 data points was
The material used in this work came from a tube that has been
used for the moving average. A grain was defined as a region en-
characterized in a previous work [18]. In the as-delivered condi-
closed by boundaries containing more than 10 points, and a
tion, the microstructure consisted of an austenitic matrix phase,
boundary was defined between two neighboring points when the
fine intergranular M23C6 carbides, and large intragranular Ti(C,N)
misorientation between them exceeded 5°. This threshold was in
precipitates. This particular tube was chosen for having shown a
general high enough to allow deformation-induced misorienta-
homogeneous microstructure. The chemical composition of this
tions, but low enough to identify grain boundaries.
tube is given in Table 1.
Each point in the EBSD map is directly connected to four
In order to minimize the overall dislocation density in the
neighboring points, and diagonally connected to four additional
material, it was recovery heat treated at 1050 °C for 4 h and slowly
points. In this work, KAM with a kernel consisting of only the four
cooled down to room temperature inside the furnace. This heat
closest neighboring points was used to assign a scalar value to
treated material has been considered as being free from effective
each point, representing its local misorientation level. The KAM
plastic strains in this work. The heat treatment caused the average were lognormally distributed, and the average KAM, KAMave, over
grain size, measured by the intercept method, to increase from a sampled region was calculated as
around 50 mm to 70 mm. The Ti(C,N) precipitates remained un-
changed, but the intergranular M23C6 carbides had changed from a ⎡ N ⎤
1
KAMave = exp⎢ ∑ ln( KAML, i)⎥,
semi-continuous network of fine carbides to coarse and discreet ⎢⎣ N ⎥⎦
i=1 (2)
ones, but were still located exclusively at the grain boundaries. No
intragranular M23C6 carbides were observed either before or after where KAML, i is the local KAM at point i, and N is the number of
the heat treatment. points in the sampled area. Such an average was calculated for
Cylindrical uniaxial tensile specimens were manufactured from each of the specimens with different level of strain. As shown in
the heat treated material along the tube's axial direction. The Fig. 1, KAMave increased approximately proportionally with uni-
nominal diameter was 5.00 mm, and the gauge length was axial plastic strain, and a linear function was fitted through the
12.55 mm. All specimens were strained at room temperature using data points to serve as a calibration curve. The level of mis-
a servo-hydraulic testing machine at a strain rate of approximately orientation uniformity in the specimens was assessed by splitting
3  10 4 s 1, with strain, ε, defined as the logarithmic strain as the full sampled area into 15 subregions of the same size. Their

Table 1
Composition of the alloy 690 tube.

Element Ni Cr Fe C Si Mn P S N Ti Al

Weight-% Bal. 29.5 10.0 0.020 0.28 0.31 0.007 0.001 0.040 0.35 0.18
R.R. Shen et al. / Materials Science & Engineering A 674 (2016) 171–177 173

n1
σf = K1⋅εep (
+ K2⋅exp n2εep , ) (4)

where K1, n1, K2 and n2, are fitting parameters, matched well to the
stress-strain curve up to necking, which occurred at around 0.37
strain.
By adding the strain from the indentations to the effective
plastic strain map that was predicted based on the local mis-
orientations, Eqs. (3) and (4) were used to predict the hardness
distribution over the surface.
This approach, at least on a larger size scale, is supported by the
work of Overman et al. [3]. They showed that hardness in Alloy
690 can be linked to average misorientation through Tabor's re-
lation, under the assumption that average misorientation is pro-
portional to plastic strain. A similar verification in this work using
Vickers hardness tests on the pre-strained samples is shown in
Fig. 2. A load of 295 N was used for these indentations, with a 20 s
dwell time.
Fig. 1. Average KAM increased approximately proportionally with increasing
plastic uniaxial strain.
2.4. Nanoindentation
respective standard deviations of average misorientation are re-
presented by the error bars in Fig. 1. The EBSD specimen with material subjected to 0.144 plastic
Such a calibration curve is normally used to estimate the plastic strain was selected for nanoindentation since it exhibited the lar-
gest variation in intra-grain misorientation. Extra care was taken
strain averaged over a sampled area. However, in our situation we
during the grinding of the specimens to keep the top and bottom
use the same relation to estimate local effective plastic strain from
surfaces of the specimen parallel to ensure compatibility with the
local misorientation.
nanoindentation. Ti(C,N) precipitates, naturally present in the
material and visible on the as-polished surface, were used as na-
2.3. Prediction of local hardness vigation markers to find the region previously mapped by EBSD.
Nanoindentation was performed at several chosen sites within
By performing hardness indentation, additional strains would the region already mapped by EBSD. The indents were made using
be introduced into the surface in addition to the pre-existing ones. a Nano Indenter IIs (Nano Instruments in Oak Ridge, Tennessee),
Tabor showed that an additional effective plastic strain of 0.08 was fitted with a Berkovich tip. Indentations were made with a loading
representative for Vickers indenters [20]. A Berkovich tip, used for rate of 0.05 mN/s until a nominal depth of 150 nm was reached.
nanoindentation in this work, has the same equivalent cone angle The load was then held constant for a dwell time of 20 s. The
and thus the same representative strain. Tabor further showed subsequent unloading was also set to a rate of 0.05 mN/s. At each
that a material's hardness, H, can be approximated to site where the hardness was mapped, the indents were placed
along three parallel rows 2.6 mm apart, with a 2.6 mm center to
H = 3⋅σf , (3) center spacing.
The projected contact areas of the indents were measured in a
where sf is the material's flow stress when the additional strain
JSM-7800F field emission SEM. The images were acquired using an
from the indentation is taken into account to the pre-existing
accelerating voltage of 15 kV and a working distance of around
strains. 10 mm. All contact area determinations in the SEM micrographs
The flow stress as a function of effective plastic strain, εep, was were made solely by one of the authors in order to avoid com-
obtained by performing a uniaxial tensile test in the same way as parison inaccuracies from subjective interpretations. These SEM
for the pre-strained materials, but all the way to failure. Fig. 2 images were also used in conjunction with the EBSD data to de-
shows that Ludwigson's equation [21] termine the exact location of the indents relative the EBSD map.

Fig. 2. A material model was fitted to the curve in order to extrapolate the hard-
ening behavior to strains beyond necking. Tabor's relation was verified using Fig. 3. Example of an SEM image of indents used for measuring the indent contact
Vickers hardness tests. area.
174 R.R. Shen et al. / Materials Science & Engineering A 674 (2016) 171–177

Fig. 4. A small region of the predicted hardness map is shown with a) the indent locations in a grain, b) the comparison between the measured and predicted hardness
profiles, and c) the indentation order.

3. Results and discussion and the predicted hardness values in this region, and the order of indent
placement. The trends in the predicted hardness profile correlated well
The indents were roughly of the same size, and pile up was ob- with those in the measured one. Both profiles agreed on a diagonal
served for all indents. An example of an SEM image of indents is band of higher hardness on the right side, lower hardness near the
shown in Fig. 3. The indent spacing is short, and was intentionally set upper corners, and hardness values in-between on the lower left side.
shorter than that recommended by ASTM E2546-15 [22]. This was a The two sets of hardness maps however showed a significant
carefully chosen balance between the requirement of a high spatial difference in relative variance. The highest measured hardness
resolution and being able to measure the projected indent area with values were roughly a factor 1.2 of the lowest ones, while the
sufficient precision. As a consequence each indent has most likely prediction indicated that the same factor is around 3 in the same
been affected by the previous indent and the previous row of in- area. The hardness prediction did not account for crystal orienta-
dents. This should theoretically cause a systematic increase in the tion or for the indent size effect (ISE), but both of these affect the
hardness values of the first row, and a larger systematic increase in hardness by a multiplicative factor. Since the crystal orientation is
both subsequent rows. While the accuracy of the measured absolute roughly constant within a grain, and the ISE was kept constant by
hardness values can certainly be questioned, the systematic nature of only making indents of the same size, the errors from these two
the error still allows an unambiguous statement of the spatial sources would only affect the absolute hardness values, but not
hardness profile. A similar reasoning was used by Marra to study the their trends in spatial distribution or their degree of relative var-
hardness profile near grain boundaries of Alloy 690 [23]. iance. The large discrepancy in relative variance in hardness thus
The indent locations of a hardness map placed within a selected shows that the method for estimating effective plastic strains is
grain is shown in Fig. 4 along with a comparison between the measured not fully applicable at the microstructural size scale.

Fig. 5. A region with a) indents crossing a twin boundary, b) the comparison between the measured and predicted hardness profiles in this region with the twin boundary
marked by the diagonal line, and c) the indentation order with excluded points in gray.
R.R. Shen et al. / Materials Science & Engineering A 674 (2016) 171–177 175

Fig. 6. A region with a) the locations of two sets of indents marked out, b) the comparison between the measured and predicted hardness profiles in the set lying within one
grain, d) the comparison of the set crossing a grain boundary, and c) and e) the indentation order of respective map with excluded points in gray. The grain boundary is
marked by a diagonal line.

Similar results can be seen in another region shown in Fig. 5,


where the indents were placed across a twin boundary. Indents
placed directly on grain boundaries and twin boundaries were
excluded so that all results are representative of the matrix. Again,
the spatial trends in hardness were captured by the prediction,
and again, there was a significant difference in the relative var-
Fig. 7. The band contrast is relatively uniform in the grain on the right side.
iance in predicted and measured hardness.
Another two sets of maps are shown in Fig. 6. The set of indents
shown in Fig. 6b) was placed within one grain. This set again showed higher background SSD density. This is supported by a map of the
a spatial correlation in the trends between the measured and the EBSD pattern band contrast showing a relatively uniform band
predicted hardness maps, along with the same large difference in contrast in the right grain, which is shown in Fig. 7.
relative variance. However, the set of indents shown in Fig. 6d), which Band contrast is a value calculated by the HKL EBSD software as
was placed across a grain boundary, shows a very different result. The a way of quantifying the image quality of the EBSD pattern, which
hardness profile on the left side of the grain boundary showed similar deteriorates with increasing dislocation density. While this has
results as the previously shown maps, but the right side did not. The been proposed as an alternative method for plastic strain esti-
measured hardness was consistently high on the right side of the mation [24], it has not attracted a wide spread usage due to its
grain boundary, and did not show the softer region further away from sensitivity to many other factors aside from dislocation density,
the grain boundary as the predicted map had shown. including crystal orientation. The pattern quality can at times
While the change in crystal orientation across the grain differ significantly even across a low angle boundary, which makes
boundary could explain a jump in absolute hardness, it cannot the method challenging to use on the micrometer scale.
account for the discrepancy in the trends between the predicted To better understand the overall higher hardness in the right
and measured hardness. Interestingly though, the discrepancy in side of Fig. 6d), this region was re-mapped by EBSD after na-
Fig. 6d) is instead an observation where the hardness prediction noindentation. A shorter EBSD step length of 0.2 mm was used, and
fails, and a further study appears motivated for a better under- a KAM map is shown in Fig. 8, which was calculated using points
standing. The uniform level of measured hardness suggests that 0.9 mm from the kernel center like the previous maps.
the total dislocation density may be more uniform than what is The result shows that the local misorientations around the
expected based on the KAM distribution, which would imply a indents on the right side are higher and extends further. As
176 R.R. Shen et al. / Materials Science & Engineering A 674 (2016) 171–177

effect of variations in GND density. Since misorientations depend


on the GND density, and hardness on total dislocation density,
both methods would still show the same spatial trends. Thus local
KAM can give a good indication of the qualitative hardness profile,
or effective plastic strains, across the surface. Using the macro-
scopic relationship between misorientations and strains to quan-
tify local hardness or strain on the micrometer scale is however
discouraged. Future work on other materials, strain rates and
temperatures are nonetheless required to clarify how far the re-
sults of this work extend.
Fig. 8. An indented region was re-mapped to assess the deformation field around
the indents. The cubes represent the crystal orientations in the two grains.

Acknowledgments
expected, overlaps in indent deformation fields are evident in
many cases, but do not appear to explain the overall higher We are grateful and for the funding provided from Vattenfall,
hardness of the right side alone. The first indents were laid in the Fortum and Eon/Uniper and for the material used in this work
bottom right of Fig. 8 and show limited overlap, yet still exhibited provided by the Electric Power Research Institute (EPRI). We also
high hardness values. The overall higher hardness on the right side wish to acknowledge Dr. C.F.O. Dahlberg for his assistance with the
is more likely to be explained by a different crystal orientation in-house EBSD post processing script, and M. Öberg and K.
with respect to the loading, or the previously discussed elevated Lindqvist for their help with the uniaxial tensile testing.
background SSD density.
The significantly smaller variations in the measured hardness
compared to the predicted hardness observed in all maps could be Appendix A. Supplementary material
explained by the presence of a relatively high and uniform SSD
density in the background. This can be interpreted as a significant Supplementary data associated with this article can be found in
portion of the plastic strain being homogeneous, and that this the online version at http://dx.doi.org/10.1016/j.msea.2016.07.123.
homogeneous portion of the plastic strain is roughly spatially
uniform. This would dilute the effect of GND density variations on
the total dislocation density. Since hardness is a function of the References
total dislocation density, the relative variation of measured hard-
ness would be small, but still follow the same trends as the GND
[1] E.M. Lehockey, Y.-P. Lin, O.E. Lepik, Mapping residual plastic strain in materials
density. Misorientations on the other hand are only based on the using electron backscatter diffraction, in: A.J. Schwartz, M. Kumar, B.L. Adams
GND part of the total dislocation density, and thus effective plastic (Eds.), Electron Backscatter Diffraction in Materials Science, Kluwer Academic-
strain-, and hardness-, predictions based on local misorientations Plenum Publishers, New York, 2000, pp. 247–264.
[2] M. Kamaya, Assessment of local deformation using EBSD: quantification of
would show a much larger relative variance, and thus the method accuracy of measurement and definition of local gradient, Ultramicroscopy 111
should not be used for quantification of local hardness. (2011) 1189–1199.
The generally good spatial correlation found in this work be- [3] N.R. Overman, M.B. Toloczko, M.J. Olszta, S.M. Bruemmer, Strain correlation in
Alloy 690 materials using electron backscatter diffraction and Vickers hard-
tween misorientations and hardness, and the lack of correlation ness, in: Corrosion 2014, March 9–13, 2014, San Antonio, TX, Paper No. NACE-
between misorientation and local plastic strain observed in other 2014–4453. NACE International, Houston, TX.
works [14–16], suggests that plastic strain and effective plastic [4] D.P. Field, P.B. Trivedi, S.I. Wright, M. Kumar, Analysis of local orientation
gradients in deformed single crystals, Ultramicroscopy 103 (2005) 33–39.
strain are not synonymous at the micrometer scale. While these
[5] A. Kundu, D.P. Field, Influence of plastic deformation heterogeneity on de-
previous works have shown that it is inappropriate to use local velopment of geometrically necessary dislocation density in dual phase steel,
misorientations for assessment of local plastic strain on the mi- Mater. Sci. Eng. A 667 (2016) 435–443.
crometer scale, the results in this work show that local mis- [6] L.N. Brewer, M.A. Othon, L.M. Young, T.M. Angeliu, Misorientation mapping for
visualization of plastic deformation via electron back-scattered diffraction,
orientations are still useful for obtaining a qualitative map of the Microsc. Microanal. 12 (2006) 85–91.
effective plastic strain. [7] P.L. Andresen, Stress corrosion cracking of current structural materials in
It should nonetheless be pointed out that misorientations and commercial nuclear power plants, Corrosion 69 (2013) 1024–1038.
[8] J. Hou, T. Shoji, Z.P. Lu, Q.J. Peng, J.Q. Wang, E.-H. Han, W. Ke, Residual strain
plastic strain are related through the storage of dislocations. A measurement and grain boundary characterization in the heat-affected zone
different strain rate, temperature or material may change the de- of a weld joint between Alloy 690TT and Alloy 52, J. Nucl. Mater. 397 (2010)
formation mechanism and thus potentially change the outcome. 109–115.
[9] A. Sáez-Maderuelo, L. Castro, G. de Diego, Plastic strain characterization in
Therefore, our results with clear spatial correlation between pre-
austenitic stainless steels in nickel alloys by electron backscatter diffraction, J.
dicted and measured hardness, for this material and deformation, Nucl. Mater. 416 (2011) 75–79.
suggest further studies for a broader applicability. [10] X. Zhang, K. Matsuura, M. Ohno, S. Suzuki, Quantification of local plastic strain
distribution beneath surface of deformed iron, Mater. Sci. Eng. A 564 (2013)
169–175.
[11] K. Radwański, Application of FEG-SEM and EBSD methods for the analysis of
4. Conclusions the restoration process occurring during continuous annealing of dual-phase
steel strips, Steel Res. Int. 86 (2015) 1379–1390.
[12] D. Dziaszyk, E.J. Payton, F. Friedel, V. Marx, G. Eggeler, On the characterization
It was found that the predicted hardness distribution based on of recrystallized fraction using electron backscatter diffraction: A direct
local KAM correlated well with the spatial trends in measured comparison to local hardness in an IF steel using nanoindentation, Mater. Sci.
hardness using nanoindentation in most cases in the nickel-base Eng. A 527 (2010) 7854–7864.
[13] L.N. Brewer, D.P. Field, C.C. Merriman, Mapping and assessing plastic de-
Alloy 690 material tested in this work. The KAM-based method
formation using EBSD, in: A.J. Schwartz, M. Kumar, B.L. Adams, D.P. Field (Eds.),
however vastly overestimated the variation in hardness, and for Electron Backscatter Diffraction in Material Science, Springer, US, 2009,
example suggested a ratio of 3 between maximum and minimum pp. 251–262.
hardness where the measured hardness only exhibited a ratio of [14] M. Kamaya, J.Q. da Fonseca, L.M. Li, M. Preuss, Local plastic strain measure-
ment by EBSD, Appl. Mech. Mater. 7–8 (2007) 173–179.
1.2. This could be explained by the presence of a high and uniform [15] K. Sasaki, M. Kamaya, T. Miura, K. Fukuya, Correlation between microstructural
level of SSD density in the background, which would dilute the scale plastic strain and misorientation, J. Jpn. Inst. Met 74 (2010) 467–474.
R.R. Shen et al. / Materials Science & Engineering A 674 (2016) 171–177 177

[16] S. Mukherjee, S.K. Mishra, I. Samajdar, P. Pant, Local strain calculations using [20] D. Tabor, The Hardness of Metals, Clarendon Press, Oxford, 1951.
electron backscatter diffraction (EBSD) measurements and digital image pro- [21] D.C. Ludwigson, Modified stress-strain relation for FCC metals and alloys,
cessing, Mater. Sci. Forum 702–703 (2012) 562–565. Metall. Trans. 2 (1971) 2825–2828.
[17] M.F. Ashby, The deformation of plastically non-homogeneous materials, Phi- [22] ASTM, E2546 15 Standard Practice for Instrumented Indentation Testing,
los. Mag. 21 (1970) 339–424. ASTM International, West Conshohocken, PA, 2015.
[18] R.R. Shen, B. Kaplan, P. Efsing, Experimental and theoretical investigation of [23] J.J. Marra, Relationship of grain boundary structure and mechanical properties
three Alloy 690 mockup components: Base metal and welding induced of Inconel 690 (Thesis), MIT, MA, 2009.
changes, Int. J. Nucl. Energy 2014 (2014) 504927. [24] A.J. Wilkinson, D.J. Dingley, Quantitative deformation studies using electron
[19] M. Kamaya, A smoothening filter for misorientation mapping obtained by back scatter patterns, Acta Metall. Mater. 39 (1991) 3047–3055.
EBSD, Mater. Trans. 51 (2010) 1516–1520.

You might also like