the effect of grain orientation on nanoindentation

9
Full length article The effect of grain orientation on nanoindentation behavior of model austenitic alloy Fe-20Cr-25Ni * Tianyi Chen a, * , Lizhen Tan a , Zizhe Lu b , Haixuan Xu b a Oak Ridge National Laboratory, Oak Ridge, TN 37831, USA b Department of Materials Science and Engineering, University of Tennessee, Knoxville, TN 37996, USA article info Article history: Received 26 April 2017 Received in revised form 10 July 2017 Accepted 12 July 2017 Available online 26 July 2017 Keywords: Slip Twinning Creep Anisotropy 709 abstract Instrumented nanoindentation was used to investigate the hardness, elastic modulus, and creep behavior of an austenitic Fe-20Cr-25Ni model alloy at room temperature, with the indented grain orientation being the variant. The samples indented close to the {111} surfaces exhibited the highest hardness and modulus. However, nanoindentation creep tests showed the greatest tendency for creep in the {111} indented samples, compared with the samples indented close to the {001} and {101} surfaces. Scanning electron microscopy and cross-sectional transmission electron microscopy revealed slip bands and dis- locations in all samples. The slip band patterns on the indented surfaces were inuenced by the grain orientations. Deformation twinning was observed only under the {001} indented surfaces. Microstruc- tural analysis and molecular dynamics modeling correlated the anisotropic nanoindentation-creep behavior with the different dislocation substructures formed during indentation, which resulted from the dislocation reactions of certain active slip systems that are determined by the indented grain orientations. © 2017 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved. 1. Introduction Nanoindentation is a robust technique for studying localized deformation behavior at the nanoscale and micron scale; it pro- vides an efcient and economical approach to developing a fundamental understanding of the deformation mechanisms that can contribute to multi-scale modeling [1e 7]. Small-volume deformation is often required to assess the mechanical properties of size-limited structures such as nanostructured materials, thin lms, and ion irradiationedamaged regions [7e11]. Anisotropic hardness and elastic modulus have been observed in some mate- rials at nanoindentation scales [2,3,8,12], which are related to the activation of different deformation mechanisms when the material is stressed along different crystallographic orientations [3e5,13]. Besides hardness and modulus, nanoindentation can probe creep properties from a very small volume of materials [7,8,14e18]. While holding the compression load during nanoindentation, materials experience continuous displacements, mimicking the conventional primary and secondary creep stages, from which similar creep parameters are extractable [14e16]. However, indentation-induced signicant plastic deformation prior to indentation creep can be a problem and requires serious consideration in tting nanoindentation-creep results to the power-law equation [7,17,18]. In addition, the indentation direction with respect to the indented grain orientation inuences the activation of different slip systems, leading to possible variations in indentation creep behavior. To the authors' best knowledge, a systematic investigation of the anisot- ropy in nanoindentation creep, combining microstructural analysis and modeling efforts, has never been carried out. Austenitic stainless steels (SSs) are widely used in different components of nuclear and fossil power plants. Newly developed and commercialized heat-resistant austenitic SSs, targeting * This manuscript has been authored by UT-Battelle, LLC under Contract No. DE-AC05-00OR22725 with the U.S. Department of Energy. The United States Government retains and the publisher, byaccepting the article for publication, acknowledges that the United States Government retains a non-exclusive, paid-up, irrevocable, worldwide license to publish or reproduce the published form of this manuscript, or allow others to do so, for United States Government purposes. The Department of Energy will provide public access to these results of federally sponsored research in accordance with the DOE Public Access Plan (http://energy.gov/downloads/doe-public-access-plan). * Corresponding author. E-mail address: [email protected] (T. Chen). Contents lists available at ScienceDirect Acta Materialia journal homepage: www.elsevier.com/locate/actamat http://dx.doi.org/10.1016/j.actamat.2017.07.028 1359-6454/© 2017 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved. Acta Materialia 138 (2017) 83e91

Upload: others

Post on 04-Feb-2022

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: The effect of grain orientation on nanoindentation

lable at ScienceDirect

Acta Materialia 138 (2017) 83e91

Contents lists avai

Acta Materialia

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

Full length article

The effect of grain orientation on nanoindentation behavior of modelaustenitic alloy Fe-20Cr-25Ni*

Tianyi Chen a, *, Lizhen Tan a, Zizhe Lu b, Haixuan Xu b

a Oak Ridge National Laboratory, Oak Ridge, TN 37831, USAb Department of Materials Science and Engineering, University of Tennessee, Knoxville, TN 37996, USA

a r t i c l e i n f o

Article history:Received 26 April 2017Received in revised form10 July 2017Accepted 12 July 2017Available online 26 July 2017

Keywords:SlipTwinningCreepAnisotropy709

* This manuscript has been authored by UT-Battellretains and the publisher, by accepting the article for plicense to publish or reproduce the published form oprovide public access to these results of federally spon* Corresponding author.

E-mail address: [email protected] (T. Chen).

http://dx.doi.org/10.1016/j.actamat.2017.07.0281359-6454/© 2017 Acta Materialia Inc. Published by E

a b s t r a c t

Instrumented nanoindentation was used to investigate the hardness, elastic modulus, and creep behaviorof an austenitic Fe-20Cr-25Ni model alloy at room temperature, with the indented grain orientationbeing the variant. The samples indented close to the {111} surfaces exhibited the highest hardness andmodulus. However, nanoindentation creep tests showed the greatest tendency for creep in the {111}indented samples, compared with the samples indented close to the {001} and {101} surfaces. Scanningelectron microscopy and cross-sectional transmission electron microscopy revealed slip bands and dis-locations in all samples. The slip band patterns on the indented surfaces were influenced by the grainorientations. Deformation twinning was observed only under the {001} indented surfaces. Microstruc-tural analysis and molecular dynamics modeling correlated the anisotropic nanoindentation-creepbehavior with the different dislocation substructures formed during indentation, which resulted fromthe dislocation reactions of certain active slip systems that are determined by the indented grainorientations.

© 2017 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

1. Introduction

Nanoindentation is a robust technique for studying localizeddeformation behavior at the nanoscale and micron scale; it pro-vides an efficient and economical approach to developing afundamental understanding of the deformation mechanisms thatcan contribute to multi-scale modeling [1e7]. Small-volumedeformation is often required to assess the mechanical propertiesof size-limited structures such as nanostructured materials, thinfilms, and ion irradiationedamaged regions [7e11]. Anisotropichardness and elastic modulus have been observed in some mate-rials at nanoindentation scales [2,3,8,12], which are related to theactivation of different deformation mechanisms when the materialis stressed along different crystallographic orientations [3e5,13].Besides hardness and modulus, nanoindentation can probe creepproperties from a very small volume of materials [7,8,14e18]. While

e, LLC under Contract No. DE-AC0ublication, acknowledges that thef this manuscript, or allow otherssored research in accordance with

lsevier Ltd. All rights reserved.

holding the compression load during nanoindentation, materialsexperience continuous displacements, mimicking the conventionalprimary and secondary creep stages, from which similar creepparameters are extractable [14e16]. However, indentation-inducedsignificant plastic deformation prior to indentation creep can be aproblem and requires serious consideration in fittingnanoindentation-creep results to the power-law equation [7,17,18].In addition, the indentation direction with respect to the indentedgrain orientation influences the activation of different slip systems,leading to possible variations in indentation creep behavior. To theauthors' best knowledge, a systematic investigation of the anisot-ropy in nanoindentation creep, combining microstructural analysisand modeling efforts, has never been carried out.

Austenitic stainless steels (SSs) are widely used in differentcomponents of nuclear and fossil power plants. Newly developedand commercialized heat-resistant austenitic SSs, targeting

5-00OR22725 with the U.S. Department of Energy. The United States GovernmentUnited States Government retains a non-exclusive, paid-up, irrevocable, worldwideto do so, for United States Government purposes. The Department of Energy willthe DOE Public Access Plan (http://energy.gov/downloads/doe-public-access-plan).

Page 2: The effect of grain orientation on nanoindentation

T. Chen et al. / Acta Materialia 138 (2017) 83e9184

promising thermal and mechanical properties and corrosionresistance in harsh environments, are likely to benefit life extensionand improve the safety margin and thermal efficiency of powerplants. Recently, the NF709 (20Cr-25Ni-1.5MoTiNb) class of alloywas selected for development as a candidate structural material forsodium-cooled fast reactors because of its superior creep andcorrosion resistance compared with 300-series austenitic SS.Research efforts on NF709 have focused on creep behavior [19],thermal stability [20,21], and radiation resistance [22]. There is stillno fundamental understanding of the deformation mechanisms ofNF709, which would be of great value to the comprehension ofstrengthening-failure mechanisms, material development, andmulti-scale modeling efforts.

In this study, we investigated the anisotropic nanoindentationhardness, elastic modulus, and creep behavior of a model alloy ofNF709. The deformation mechanisms were revealed throughelectron microscopic observations and molecular dynamics (MD)simulation. The methodology explored in this study will be appli-cable to studies of similar materials. In particular, the results andunderstandings obtained from this study will serve as a criticalbaseline and guidance for further experimental and theoreticalstudies of the engineered alloy NF709 and alloys with radiationdamage.

2. Materials and methods

Amodel austenitic alloy of NF709 with a measured compositionof Fe-20Cr-24.5Ni-0.0013C in weight percentages was used in thisstudy. The alloy was cast using vacuum arc melting and solution-annealed at 1050 �C for 1 h, followed by hot rolling at 900 �Cwith an 85% thickness reduction to 3.8 mm and water quenching. Asample of the alloy was mechanically polished to a mirror finishwith colloidal silica as the final preparation step.

Nanoindentation tests were conducted on the polished sampleusing an Agilent G-200 nanoindenter with a Berkovich tip at roomtemperature. Four indentation matrixes (8 � 10) with 20 mminterspacing between the indents were performed on the sampleusing the continuous stiffness measurement with the hardness andelastic modulus calculated using Oliver and Pharr's method [23].Each nanoindentation test ran at a constant loading rate of0.053 mN/s to a maximum depth of 1000 nm into the indentedsurface. The hardness and modulus values were averaged over theindentation depth of 800e900 nm. After the maximum indentationdepth was reached, the load was held for 600 s to probe the creepbehavior of the material.

Electron backscatter diffraction (EBSD: EDAX OIM 6.0) wasperformed on the indentation matrixes to analyze the grain sizes ofthe material and to identify the grain orientation under eachindent. Results for the indentations, measured near the centers ofthe grains larger than 30 mm, were included in the statistical ana-lyses. Presumably, there was negligible influence from grainboundaries because of the small indentation size compared withthe grain size. Indentations exerted on sample surfaces close to{001}, {101}, and {111} were selected for detailed microstructuralcharacterization. Scanning electron microscopy (SEM: JEOL 6500 Fat 5 kV) investigated the indentation-induced surface morphologychanges; cross-sectional transmission electron microscopy (TEM:JEOL 2100 F at 200 kV) in either conventional TEM or scanning TEM(STEM) modes investigated the microstructural evolution under-neath the indents. A focused-ion-beam (FIB: FEI Versa 3D) instru-ment was used to lift out the lamella underneath the selectedindents for TEM characterization. Progressive final cleaning stepsusing a 5 to 2 kV ion beam were employed to minimize possibleartifacts on the lamellas from FIB processing. The localized defor-mation mechanisms were understood by correlating the

mechanical properties with the microstructures.In addition to the experimental efforts, the large-scale atomic/

molecular massively parallel simulator (LAMMPS, http://lammps.sandia.gov) [24] was used for MD simulations of the indentationsin three Fe-20Cr-25Ni systems with surface orientations of (001),(101), and (111). An embedded atom method potential developedby Bonny et al. [25] was used. The simulation cells had dimensionsof ~40� ~40� ~40 nm3with a periodic boundary condition appliedin the x- and y-axes. In the z-axis, six layers of atoms at the bottomwere fixed. The top surface in the z-axis was a free surface that wasindented. The radius of the rigid spherical indenter was 5 nm. Theinteractions between the indenter and the sample were describedby the Ziegler-Biersack-Littmark (ZBL) part of the potential fromBonny. Before the nanoindentation simulation, the systems wereequilibrated at 300 K using a constant-temperature, constant-vol-ume ensemble (NVT). The OVITO software was used to visualizeand analyze the simulation results [26].

3. Results and discussion

3.1. Hardness and elastic modulus

The EBSD characterization indicated equiaxed grains with sometwins in the material, suggesting complete dynamic recrystalliza-tion occurred during the hot rolling. The grain size, analyzed byarea, averaged 40 ± 20 mm in diameter, disregarding the twins.Among the 320 nanoindentation tests, 70 indents were found nearthe central areas of the 70 large grains. The elastic modulus andhardness results of the 70 indentation tests are plotted in thestandard stereographic triangles (SST) with color-coded values inFig. 1aeb. Each data point represents the indented grain crystal-lographic orientation referring to the surface normal, i.e., the di-rection of the indentation. The statistic modulus and hardnessdistributions of the indentations close to {001}, {101}, and {111} areplotted in Fig. 1ced, with the indented grain orientations within asolid angle of 0.14 steradian from the pole directions. To bettervisualize the distributions, the Gaussian fitting curves weresuperimposed on the statistical data. The modulus results in Fig. 1aand c show a strong anisotropy depending on the crystallographicorientation of each grain. The grains close to {111} have the highestmoduli on average, followed by the grains close to {101} and then{001}. The elastic moduli range from 197 to 234 GPa, slightly higherthan the values for 316L SS (155e190 GPa) and for the modelaustenitic alloy Fe-15Cr-15Ni (171 GPa) [12,13]. The dependence ofhardness on grain orientation in Fig. 1b and d is not strong; yet thegrains near {111} have the highest hardness, whereas the hardnessnear {001} and {101} is comparably lower. The hardness rangesfrom 1.85 to 2.21 GPa, which is close to the nanoindentationhardness range of 316L (2.0e2.2 GPa) but higher than that of alloyFe-15Cr-15Ni (1.46 ± 0.02 GPa) [12,13]. The anisotropy in elasticmodulus and hardness is in agreement with values for 316L andHadfield steels reported earlier [12,27].

The anisotropy in moduli was generally observed in uniaxialexperiments on single crystals and is attributed to the anisotropy inelastic constants. The anisotropy in hardness can be qualitativelycorrelated with the resolved shear stresses on slip and twinningsystems estimated from Schmid's law for uniaxial compression.Table 1 lists the maximum Schmid factors for face-centered cubic(fcc) slip and twinning systems under compressions along the<001>, <101>, and <111> directions and the numbers of systemshaving the maximum Schmid factors. The maximum Schmid factorfor the <111> compression is considerably lower than those for the<001> and <101> compressions; that difference could be thereason for the higher measured hardness near the <111 > pole inthe SST [8,28]. This correlation suggests that during the loading

Page 3: The effect of grain orientation on nanoindentation

Fig. 1. Color-coded SST showing elastic modulus (a) and hardness (b) from the indentations exerted on individual grains with different crystallographic orientations with respect tothe indentation direction. Distributions of elastic modulus (c) and hardness (d) of the indents close to {001}, {101}, and {111} grains, symbolized with triangles, squares and crosses,respectively, superimposed with Gaussian curves.

Table 1Schmid factors for slip and twinning in fcc systems under uniaxial compressionsalong <001>, <101> and <111>.

Compression Slip Twinning

<001> 0.41, 8 systems 0.47, 4 systems<101> 0.41, 4 systems 0.24, 4 systems<111> 0.27, 6 systems 0.16, 6 systems

T. Chen et al. / Acta Materialia 138 (2017) 83e91 85

stage of indentation, as the stress increases and the stress fieldexpands in the material, the ease of continuous activation of slip/twinning systems plays an important role in determining thenanoindentation hardness.

As will be shown later, when the indentation direction deviatesfrom the pole directions, the distributions of resolved stress on thedeformation systems vary from the ideal case as provided in Table 1.This variance is sensitive to both the deviation angle from the idealpole direction and the deviation direction. For instance, as illus-trated in Fig.1b, the near-{001} indentations deviating toward {111}direction appear to have higher harnesses than those deviatingtoward {101} direction. As a result, the statistical data in Fig. 1d forthe near-{001} indentations, spanning from ~1.85 GPa to ~2.03 GPa,shows a “bimodal” hardness distribution, which would have beendiminished if more near-{001} indentations were available in thematerial so that only indentations with smaller deviation angleswere included in the statistic.

3.2. Nanoindentation creep

Fig. 2a shows the load (F) edisplacement (h) curves for sixindentation tests on individual grains having the surface normalclose to <111>, <101>, and <001>, denoted by dotted, solid anddashed lines, respectively. Under the same load, more displacementwas observed in the {101} and {001} samples than in the {111}samples. This observation agrees with the anisotropy in hardnessshown in Fig.1. After the loading stage, themaximum loadwas heldfor 600 s; the additional displacement was monitored and is shownin Fig. 2b as a function of holding time (t). In Fig. 2b, the load-

holding-induced additional displacement (y-axis) is denoted asDh, which is the difference between the instantaneous displace-ment of the indenter (h) and its displacement at the beginning ofthe holding stage (h0). The curves in Fig. 2b resemble general creepcurves containing the primary and the secondary regimes. Thedeformation takes place more rapidly in the primary regime thanthe secondary regime. The displacement rate decreases with timein the primary regime and approximately reaches a constant in thesecondary regime for each test. Overall, the constant strain rateswere observed to be the highest in the indentations close to {111}grains and the lowest close to {101} grains.

To evaluate the tendency to creep deformation, a simple pro-cedure was proposed to generate the parameter P using Eq. (1) [18]:

PðtÞ ¼ Dhdhdt

(1)

The P values were averaged over 210e360 s of the load-holdingstage and are presented in Table 2. The indentations on grains near{111} had the largest P values, suggesting the highest tendency tocreep, followed by the indentations on grains near {001} and then{101}.

The instantaneous strain rate was deduced from _ε ¼ 1h

dhdt [14]

and is plotted as a function of stress (s) in Fig. 2cee for the sixindentation tests during the holding stage, wheres ¼ F∕(3√3h2tan265.35) [29]. Correlating the data with the classicpower-law creep model of

_ε ¼ Kexp��QRT

�sn ; (2)

where K is a material dependent constant, Q the activation energyfor creep, R the gas constant, and T the absolute temperature, thestress exponent n can be obtained from the slopes of the ln _εelnsrelationship as shown in Fig. 2cee.

All the ln _εelns data plotted in Fig. 2cee show a transition be-tween two linear fitting segments. Depending on the indentationorientation, the transition point corresponds to the time point of

Page 4: The effect of grain orientation on nanoindentation

Fig. 2. (a) Load-displacement curves from loading to unloading; (b) load-holding-induced additional displacement (Dh) as a function of holding time; (cee) strain rate-stress curvesduring the load-holding stage for six samples representing indentations on grains near {111}, {001}, and {101}, respectively; and (f) the M709 stress exponents from the nano-indentation creep tests as shown in (c) compared to high-temperature creep data for alloy 800 [30].

Table 2The creep tendency parameter (P), stress exponent (n), and strain rate sensitivity (SRS) calculated from the data of the nanoindentation-creep tests.

Indented plane (-5 -6 6) (-5 6 6) (3 -4 13) (-3 -2 19) (3 -17 16) (-1 5 6)

P (nm2s�1) 0.34 ± 0.13 0.34 ± 0.13 0.20 ± 0.09 0.17 ± 0.08 0.03 ± 0.01 0.06 ± 0.02n, 2nd 18 ± 11 34 ± 13 87 ± 63 81 ± 44 759 ± 488 345 ± 301SRS, 1st ( � 10�3) 6.2 ± 0.9 6.5 ± 0.7 4.8 ± 2.0 6.3 ± 0.7 5.9 ± 1.2 5.4 ± 0.7

T. Chen et al. / Acta Materialia 138 (2017) 83e9186

60e120 s during the load-holding stage, which separates the pri-mary and secondary regimes of the nanoindentation creep shownin Fig. 2b. The transitions in slope from the primary to the sec-ondary regime depend on the indentation orientations. From ahigher strain rate (primary regime) to a lower strain rate (second-ary regime), the slopes of the {111} group data decrease while theslopes of the {110} group data increase. Table 2 lists the stress ex-ponents n fitted from the second regime, which mimics the steady-state regime of conventional creep. The values of n are the lowestnear {111} and the highest near {110}. As will be detailed in thefollowing sections, the different indentation-induced microstruc-tures lead to the difference in n. The data plotted in Fig. 2cee arescattered, resulting to the significant standard deviations of thefitted n. That is because, at room temperature, the strain rate duringnanoindentation creep is extremely low, leading to a very smallspan of stress of about 0.05 GPa during the secondary regime.

Within such a small span of data points, any tiny disturbance, suchas thermal drifting during nanoindentation, can lead to strongfluctuations in the data. Particularly for the indents near {110},because the deformation during the secondary regime is so small,there are almost no changes in the stress during this regime.Nevertheless, the anisotropy in the nanoindentation creep behavioris strong enough to be distinguished by comparing Fig. 2cee andthe P and n values in Table 2.

As a reference, the stress exponent n for high-temperature creepof alloy 800 (Fe-21Cr-32Ni-TiAl) was fitted from the reportedminimum creep rate and stress data [30]. Fig. 2f shows the fitted nas a function of temperature. Except for the n at 750 �C fitted withnoticeably fewer data, n increases approximately linearly with thedecreasing temperature. The value of n at room temperature foralloy 800 was estimated to be ~30 by linear extrapolation, which iswithin the range of the values of n, plus/minus the standard

Page 5: The effect of grain orientation on nanoindentation

T. Chen et al. / Acta Materialia 138 (2017) 83e91 87

deviations, from the secondary regime of the near {111} nano-indentation creep tests. As will be discussed in the remainder ofthis paper, the indentation creep tests near {111} were less affectedby the indentation-induced defect substructures and thus present abetter simulation of conventional creep tests, despite the signifi-cant error bars introduced by the very small span of stress for thenanoindentation creep test.

Historically, the power-law model has been applied to severalindentation-based tests, showing good agreement between theexperimental results and literature [14e16]. However, theassumption of steady-state creep during indentation tests has beenquestioned by Goodall and Clyne [18]. Our results indicate that thevalue of n fitted from the primary stage of nanoindentation creeptests should not be treated as the steady-state stress exponent.

To understand the n values fitted from the primary regime, theBailey-Norton model [17], which takes primary creep into account,was used:

ε ¼ Asn0tm

0(3)

where t is the load-holding time; A, n0and m

0are temperature-

dependent parameters. This formula can be derived into a strain-rate equation of

_ε ¼ m0A

1m0 s

n0

m0ε

m0 �1m0 (4)

Given that an ideal Berkovich indenter has an effective strainindependent of the indentation depth [1], and that the additionaldisplacement during the load-holding stage is much smaller thanthe displacement at the beginning of this stage, the parameter nfitted by the power law in Eq. (2) from the primary creep regimeapproximates the n

0∕m

0in Eq. (4). The inverse of n

0∕m

0in Eq. (4)

refers to the strain rate sensitivity (SRS) when used inindentation-creep tests and tensile tests [31]. The SRS calculatedfrom the first 60 s, i.e., the primary stage, was listed in Table 2. Thevalues are in accordance with the results for 304 SS [31].

Both P and n suggest that nanoindentation along <111 > leads tothe strongest tendency to nanoindentation creep, whereas the<101 > grains show the highest resistance to continuous defor-mation during the nanoindentation-creep tests. To understand thisanisotropy in nanoindentation creep behavior, the followingmicrostructure analysis and modeling efforts were carried out.

3.3. Microstructure analysis

Fig. 3aec show the indents on grains with their surface normalsclose to [111], [001], and [101], respectively. In the vicinities of theindents, slip bands along certain directions were observed. Thedirections of the slip bands were found to be [-110], [-110] and [01-1] on the (111) surface, [±110] on the (001) surface, and [-1 ± 21]and [-101] on the (101) surface. The slip bands were all along theintersections of the {111} planes and the surface planes of eachsample. The configurations of atoms in the surface planes are

Fig. 3. SEM images of the indentations close to (111) (a), (001) (b), and (101) (c).

illustrated by the insets at the top right corners in Fig. 3aec, withthe dashed lines marking the intersections of {111} planes with thesurface planes, which are parallel to the observed slip bands in allthe samples. The patterns of the slip bands suggest that {111}planes are the active slip planes, and dislocation gliding is one ofthe deformation mechanisms for nanoindentation compression atroom temperature. The three-fold, four-fold, and two-fold sym-metries of the indentation-induced slip patterns on the (111), (001),and (101) surfaces were in accordance with previous studies[4,5,13].

Twinning was observed as an active deformation mechanismexclusively in the near {001} indented samples. The formation oftwinning was revealed by the TEM characterization of cross-sectional lamellas lifted underneath the indents near {001}.Fig. 4a shows an example of the deformation twinning that tookplace within a radius of ~2 mm beneath the indent tip, where thematerial was highly stressed and strained. Fig. 4b and c show dark-field images of the twins on the left and right sides of the indent tip,respectively, with the diffraction pattern insets indicating two ofthe {111} <112> twinning systems. Intersections of twinning wereobserved underneath the near {001} indents as the twins ofdifferent systems extended to cross each other.

Although indentation experiments generally apply multi-axialstresses [16], the presence of deformation twins in only the {001}sample rather than in the {101} or {111} samples was in agreementwith the results of uniaxial experiments [27]. That is because in theregion right underneath the indent, the direction of the compres-sive stress was close to the normal direction of the indented surface[16]. As shown in Table 1, the larger Schmid factor for twinning inthe <001> compressed condition suggested that twinning is thepredominant deformation mechanism over slip during compres-sion along <001>. It is worth noting that we observed two (-111)[-11-2] and (1-11) [1-1-2] twinning systems on the left and rightsides of the indent tip, respectively, as shown in Fig. 4bec. Thedirection of the stress inclined toward the normal direction of thecontacting surface between the sample and the indent tip. As aresult, a particular twinning system, which was (-111)[-11-2] on theleft or (1-11) [1-1-2] on the right of Fig. 4a, was activated by thereceived higher shear stress.

Deformation twinning was not observed in any of the near {101}or {111} indented samples, as shown in the STEM images in Fig. 5obtained on the [011] zone-axis. High densities of dislocationswere observed in both samples, with the densities graduallydecreasing from the indented surface to the deeper regions of thesamples. As pointed out by the white arrows, planar paralleldislocation arrays were observed in both samples on the (111)planes, suggesting that partial separation of dislocations was sig-nificant in these samples [35].

The dislocation arrays, as pointed out in Fig. 5, were furthercharacterized using diffraction contrast imaging. Fig. 6a presents abright-field image of the dislocation arrays in one of the near <111>indented samples, i.e., the (-5,6,6) sample shown in Fig. 2. Similarstructures as presented in Fig. 6, referred as “dislocation tangles”,have been observed in tensile-strained SS 304 and 316LN [35,36].Fig. 6b and c shows diffraction patterns of the [01-1] and [11e2]

Fig. 4. Bright-field (a) and dark-field (bec) TEM micrographs showing the {111} <112>deformation twinning underneath a near [001] indentation.

Page 6: The effect of grain orientation on nanoindentation

Fig. 5. Bright-field STEM micrographs, imaged on [011] zone axis, of near (101) (a) and(111) (b) indented samples.

T. Chen et al. / Acta Materialia 138 (2017) 83e9188

zones, respectively. Fig. 6a was obtained by tilting the sample fromthe [11e2] zone to the [01-1] zone around the [111] axis. Comparingthe direction of the dislocation arrays shown in Fig. 6a and thecrystal orientations indicated by Fig. 6bec, it can be seen that thedislocation arrays are parallel to the intersection of the (111) planeand the foil. Fig. 6d shows a higher-magnification bright-field

Fig. 6. (a) TEM micrographs showing the parallel dislocation arrays in one of the near <111axis, respectively. (d) Bright-field image obtained at the (111) two-beam condition schematicarea shown in (d), with the operating g vectors marked by the dashed circle (111) and soli

image of the dislocation arrays, obtained at a two-beam conditionof g ¼ 111, as indicated in Fig. 6b with the dashed circle. Fig. 6eegshows the same area as Fig. 6d, imaged in weak-beam dark-fieldmode with g ¼ 111, g ¼ 1-1-1, and g ¼ 2-20, respectively. Theoperating g vectors of Fig. 6eeg are marked by the dashed and solidcircles in Fig. 6b and the solid circle in Fig. 6c, respectively. Thedominant parallel dislocation arrays are in good contrast at g ¼ 111but out of contrast at g ¼ 1-1-1 and g ¼ 2-20, indicating that thedislocations in the arrays have a Burgers vector parallel to [110].Based on the microstructure of the dislocations, which containsdissociated dislocations, as shown by the arrows in Fig. 6dee, thetangled dislocations in the arrays are composed of 1/6[110] stair-rods formed by reactions of partial dislocations [35].

The microstructural analysis indicates that slip took place in allthree different indentation orientations near {001}, {101}, and{111}, while deformation twinning was activated only by in-dentations on near {001} grains. The operations of the slip andtwinning mechanisms identified in this work are consistent withliterature reports on fcc systems for deformation at low tempera-tures [18,27,32e34]. TEM micrographs suggest the formation ofstair-rod dislocations, which can serve as obstacles to dislocationglide and thus influence the creep behavior of the material. In the

> indented samples. (bec) Electron diffraction patterns at the [01-1] and [11e2] zoneally marked by the dashed circle in (b). (eeg) Weak-beam dark-field images of the samed circle (1-1-1) in (b), and the solid circle (2-20) in (c), respectively.

Page 7: The effect of grain orientation on nanoindentation

T. Chen et al. / Acta Materialia 138 (2017) 83e91 89

following section, we will confirm the formation of stair-rod dis-locations by MD simulation, which also correlates the dislocationstructures with the resistance to nanoindentation creep. That dis-cussion is followed by a discussion regarding the formationmechanisms of the stair-rod dislocations.

3.4. Modeling

MD simulations of nanoindentation were conducted on the(001), (011), and (111) surfaces of Fe-20Cr-25Ni at 300 K. Themaximum indentation depth was set to 1.4 nm with the displace-ment rate controlled at 0.01 nm/ps. Fig. 7 shows the evolutions ofthe nanoindentation hardness and densities of three types of dis-locations as a function of indentation depth. The densities of dis-locations were calculated using the Dislocation ExtractionAlgorithm [43]. In Fig. 7a, the hardness is the contact pressure,defined as the ratio of the indentation force and the projectedcontract area. The fluctuation in hardness decreased as the inden-tation depth increased. The hardness values between the indenta-tion depths of 0.7 and 1.4 nm were 18.18 ± 0.36 GPa,16.06 ± 0.22 GPa, and 16.26 ± 0.31 GPa for (111), (011), and (001)indentations, respectively. Compared with the experimental valuesshown in Fig. 1, the hardness values from the MD simulations werealmost one order of magnitude larger. The overestimation ofhardness fromMD simulations is generally observed in different fccsystems [3,37e39], primarily because of the higher strain rate andindentation size effect [3,37,38]. Nevertheless, the calculatedhighest hardness on the (111) surface and comparable hardness onthe (011) and (001) surfaces are qualitatively in agreement with theexperimental results in this study.

It is worth noting that the highest hardness was observed on(111) surface of the Fe-20Cr-25Ni alloy in this study, whereas MDsimulations of pure aluminum and nickel showed the highesthardness on the (001) surface and the lowest hardness on the (111)surface [3,37]. Ju et al. showed that when a material is deformedthrough the formation of dislocation defects, the ease of formationof which depends on the number of slip angles of the {111} planes[3]. In this study, the stacking fault energy of the alloy was muchlower than those of pure aluminum and nickel [40]; therefore,

Fig. 7. MD simulation calculated evolutions of nanoindentation hardness (a) and densitieindentation depth exerted on the (001), (101), and (111) surfaces.

dislocations tended to glide in the form of Shockley partials, asevidenced by the observations of dissociated dislocations in Fig. 6and the high-density Shockley partials in the MD results inFig. 7b. Because of the high mobility of Shockley partials, defor-mation, even on the scale of MD simulations, was shown to occurby means of the slip. Consequently, the nanoindentation hardnessis related to the ease of activation of slip systems through theresolved shear stresses on the slip systems.

Fig. 7c and d plot the densities of Lomer-Cottrell (L-C, 1/6<110>)and Hirth 1/3<100> dislocations as functions of indentation depth,respectively. There is an obvious difference in the populations of L-C dislocations among the three indentation orientations, while theevolutions of Hirth dislocations in the three cases are similar. Onthe experimental scales, the formation of L-C dislocations wasmoresignificant, given that the formation of L-C dislocations was moreenergetically favorable than that of the Hirth dislocations [41]. Infact, 1/6<110> type dislocations are much frequently observed inexperiments than 1/3<100> type dislocations, as shown in this andprevious studies [42].

The observation of L-C dislocations from the MD simulationagreed with the TEM observations in Figs. 5 and 6. In addition, MDsimulation also showed that the population of stair-rod disloca-tions was higher in (011) indented samples than in (111) indentedsamples. In fact, this difference was qualitatively observed in theSTEM images of the indented samples, as the interspacing of thedislocation arrays is much smaller in Fig. 5a than in Fig. 5b. Thenetworks formed with L-C locks were obstacles for the glidingdislocations. As shown earlier, dislocation gliding was the operatingdeformationmechanisms for (011) and (111) indented samples. It isbelieved that the difference in the population of the L-C disloca-tions was the primary reason for the different nanoindentation-creep resistances observed in the (011) versus (111) indentationsshown in Fig. 2.

To understand the formation of the L-C dislocations, we recon-sidered the activation of slip systems based on Schmid's law undercompression. The four active slip systems under indented (011) are(1-1-1)[110], (1-1-1)[101], (111)[-101], and (111)[-110]. They canreact through the following L-C reactions to form 1/6[011] stair-rods:

s of 1/6<112> (b), 1/6<110> (c), and 1/3<100> (d) dislocations as a function of the

Page 8: The effect of grain orientation on nanoindentation

T. Chen et al. / Acta Materialia 138 (2017) 83e9190

12½110� þ 1

2

h101

i¼ 1

6

h121

iþ 16½211� þ 1

6

h112

iþ 16

h211

i

¼ 16½211� þ 1

6½011� þ 1

6

h211

i(5)

and

12½101� þ 1

2

h110

i¼ 1

6

h112

iþ 16½211� þ 1

6

h121

iþ 16

h211

i

¼ 16½211� þ 1

6½011� þ 1

6

h211

i(6)

On the other hand, the six active slip systems under indented(111) are (1-1-1)[-1-10], (1-1-1)[-10-1], (-11-1)[-1-10], (-11-1)[0-1-1], (-1-11)[0-1-1], and (-1-11)[-10-1]. The slip directions of the sixsystems are either perpendicular or at angles of 60� from eachother. As a result, no L-C reactions could happen among the sixactive slip systems. The active formation reactions of L-C locks dueto the stress along the indentation direction in (110) samples ex-plains the higher population of L-C locks in (110) indented samples.

However, L-C locks were observed in the (111) indented samplesby both TEM investigation and MD simulation. These were subjectto the multi-axial stresses applied by the nanoindentation. Asshown in Fig. 8a, when the material is indented, the region rightbeneath the tip of the indenter can be considered to be compressed

along the normal of the surface plane, i.e., Fu.

∕∕[111] in the case of a(111) indent. In the circumambient regions, however, the directionof the stress F is inclined toward the radial direction of the indentby an angle q.

Fig. 8b plots the Schmid factors of the 12 slip systems as afunction of q, when the material is indented along [111]. Forsimplification, q is restricted to the (110) plane. q¼ 0 represents thestress acting in the direction of [111] when the six active slip sys-tems do not react with one another. When the direction of thestress deviates from [111], for instance, with q increasing toward

Fig. 8. (a) A schema showing the deviation of stress direction from the indentationdirection, (b) Schmid factors of the fcc slip systems as a function of the deviation angleof the stress, and (c) a schema showing the distribution of L-C locks of a (111) indentedsample when looking along the indentation direction.

the negative, the Schmid factors (or resolved shear stresses) on the(-1-11)[0-1-1] and (-1-11)[-10-1] systems (dark-blue dashed line)continue to increase while the resolved shear stresses on the otherfour active slip systems at q ¼ 0 decrease. Meanwhile, the slipsystems of (-1-1-1)[-101] and (-1-1-1)[0-11] (black solid) becomemore inclined to be activated as the resolved shear stresses on themincrease. At certain points, there is higher resolved shear stressexerted on the (-1-1-1)[-101] and (-1-1-1)[0-11] systems (blacksolid) than on the (-11-1)[ -1-10] and (1-1-1)[-1-10] (green dot-dashed) and the (-11-1)[0-1-1] and (1-1-1)[-10-1] (light-bluedotted) systems. When the two slip systems on (-1-1-1) are acti-vated, they can react with the (-1-11) systems through thefollowing L-C reactions to form 1/6[-1-10] stair-rod dislocations:

12

h101

iþ 12

h011

i¼ 1

6

h211

iþ 16

h112

iþ 16

h121

iþ 16

h112

i

¼ 16

h112

iþ 16

h110

iþ 16

h112

i

(7)

and

12

h011

iþ 12

h101

i¼ 1

6

h121

iþ 16

h112

iþ 16

h211

iþ 16

h112

i

¼ 16

h112

iþ 16

h110

iþ 16

h112

i

(8)

When the direction of the stress deviates from [111] to theopposite direction, i.e., toward the positive in Fig. 8b, the light-bluedashed, black solid, and green dot-dashed lines, representing the(-11-1)[0-1-1], (1-1-1)[-10-1], (1-1-1) [10e1], (-1-1-1)[01-1], (-11-1)[-1-10] and (1-1-1)[-1-10] systems, receive higher resolved shearstresses. No L-C reaction is possible among these systems.

Taking account of the three-fold symmetry of the atomicarrangement on the (111) plane, Fig. 8c schematically shows thedistribution of L-C locks on the (111) indented sample looking alongthe indentation direction. This pattern agrees with the slip bandpatterns on indented (111), as observed in Fig. 3a and in previousstudies [6]. Moreover, when observing a cross-sectional lamellalifted out along the dashed line in Fig. 8c, as the sample shown inFig. 5b was prepared, one should see L-C locks skewed toward oneside of the indent, which is confirmed by Fig. 5b.

It is thought that most of the dislocations and twins wereinduced by the plastic deformation during the nanoindentationloading period. The continuous deformation under the load-holding creep period relied on the continuous operation of theactive slip or twinning systems. Comparing the cases of <111> and<011> indentations, the ease of L-C lock formation under <011>compression led to the stronger resistance to creep of {011} sam-ples that is observed in Fig. 2. In the case of <001> indentation, theintersections of the twins could prevent the growth of twins underthe held load. Together with twin-slip interactions, they suppressedthe deformation of the material during the nanoindentation creeptests [33].

4. Conclusion

Systematic nanoindentation-based tests, together with micro-structural characterization using SEM and TEM, were conducted ona model 709 alloy (Fee20Cre25Ni) at room temperature. Anisot-ropies in elastic modulus, hardness, and nanoindentation creepwere investigated. A distinct grain-orientation dependency of theelastic modulus was observed for {001}, {101} and {111} grains inascending order in the range of 197e234 GPa. Similarly, {111} grains

Page 9: The effect of grain orientation on nanoindentation

T. Chen et al. / Acta Materialia 138 (2017) 83e91 91

exhibited the highest hardness, up to 2.21 GPa, but the hardnessdiscrepancy between {101} and {001} grains was not significant.

During the indentation creep tests, the time-dependentdisplacement showed a similarity to general creep curves con-taining primary and secondary creep stages. Following the power-law equation, the indentation stress exponent (n) in the secondarystage was fitted to be from ~18 to 34 for the near {111} grains. Thestress exponent values for the grains near {111} were comparable tothe stress exponent extrapolated from the literature creep data foralloy 800. Microstructural characterization revealed slip andtwinning as the two deformation mechanisms of the alloy. Unlikeslip, which was present underneath surfaces indented near {001},{101}, and {111}, twinning was observed only underneath the sur-faces indented near {001}. This finding is explainable by theorientation-dependent Schmid factors for slip and twinning underuniaxial compression. Stair-rod dislocations were observed as aresult of dislocation reactions and were confirmed by MD simula-tions. The different deformation mechanisms with varied disloca-tion substructures resulted in a descending nanoindentation creepresistance of the {101}, {001}, and {111} grains.

Acknowledgments

This material is based upon work supported by the U.S.Department of Energy, Office of Nuclear Energy, the Nuclear EnergyUniversity Program (project no. 14-6346) under contract numberDE-AC05-00OR22725. The authors would like to thank Dr. H Bei(ORNL) for reviewing this manuscript.

References

[1] A.C. Fischer-Cripps, Nanoindentation, Springer, 2011.[2] Y. Gao, H. Bei, Strength statistics of single crystals and metallic glasses under

small stressed volumes, Prog. Mater. Sci. 82 (2016) 118e150.[3] S.-P. Ju, C.-T. Wang, C.-H. Chien, J. Huang, S.-R. Jian, The nanoindentation re-

sponses of nickel surfaces with different crystal orientations, Mol. Simul. 33(2007) 905e917.

[4] Y. Wang, D. Raabe, C. Klüber, F. Roters, Orientation dependence of nano-indentation pile-up patterns and of nanoindentation microtextures in coppersingle crystals, Acta Mater. 52 (2004) 2229e2238.

[5] E. Renner, Y. Gaillard, F. Richard, F. Amiot, P. Delobelle, Sensitivity of the re-sidual topography to single crystal plasticity parameters in Berkovich nano-indentation on FCC nickel, Int. J. Plasticity 77 (2016) 118e140.

[6] Y. Liu, S. Varghese, J. Ma, M. Yoshino, H. Lu, R. Komanduri, Orientation effectsin nanoindentation of single crystal copper, Int. J. Plasticity 24 (2008)1990e2015.

[7] V.M.F. Marques, B. Wunderle, C. Johnston, P.S. Grant, Nanomechanical char-acterization of Sn-Ag-Cu/Cu joints-Part 2: nanoindentation creep and itsrelationship with uniaxial creep as a function of temperature, Acta Mater. 61(2013) 2471e2480.

[8] H. Ljungcrantz, M. Od�en, L. Hultman, J.E. Greene, J.-E. Sundgren, Nano-indentation studies of single-crystal (001)-, (011)-, and (111)-oriented TiNlayers on MgO, J. Appl. Phys. 80 (12) (1996) 6725e6733, http://dx.doi.org/10.1063/1.363799.

[9] P. Hosemann, Studying radiation damage in structural materials by using ionaccelerators, Rev. Accel. Sci. Technol. (2012) 161e182, http://dx.doi.org/10.1142/9789814383998_0008.

[10] D. Kiener, A.M. Minor, O. Anderoglu, Y. Wang, S.A. Maloy, P. Hosemann,Application of small-scale testing for investigation of ion-beam-irradiatedmaterials, J. Mater. Res. 27 (21) (2012) 2724e2736, http://dx.doi.org/10.1557/jmr.2012.303.

[11] A. Lupinacci, K. Chen, Y. Li, M. Kunz, Z. Jiao, G.S. Was, M.D. Abad, A.M. Minor,P. Hosemann, Characterization of ion beam irradiated 304 stainless steelutilizing nanoindentation and Laue microdiffraction, J. Nucl. Mater 458 (2015)70e76, http://dx.doi.org/10.1016/j.jnucmat.2014.11.050.

[12] C. Tromas, J.C. Stinville, C. Templier, P. Villechaise, Hardness and elasticmodulus gradients in plasma-nitrided 316L polycrystalline stainless steelinvestigated by nanoindentation tomography, Acta Mater. 60 (2012)1965e1973.

[13] Y.Z. Xia, H. Bei, Y.F. Gao, D. Catoor, E.P. George, Synthesis, characterization, and

nanoindentation response of single crystal FeeCreNi alloys with FCC and BCCstructures, Mater. Sci. Eng. A 611 (2014) 177e187.

[14] M.J. Mayo, W.D. Nix, A micro-indentation study of superplasticity in Pb, Snand Sn-38 wt Pb, Acta Metall. 36 (8) (1988) 2183e2192.

[15] S.A. Syed Asif, J.B. Pethica, Nano-scale indentation creep testing at non-ambient temperature, J. Adhes. 67 (1998) 153.

[16] C.L. Wang, Y.H. Lai, J.C. Huang, T.G. Nieh, Creep of nanocrystalline nickel: adirect comparison between uniaxial and nanoindentation creep, Scr. Mater 62(2010) 175e178.

[17] N. Ogbonna, N.A. Fleck, A.C.F. Cocks, Transient creep analysis of ball inden-tation, Int. J. Mech. Sci. 37 (11) (1995) 1179e1202.

[18] R. Goodall, T.W. Clyne, A critical appraisal of the extraction of creep param-eters from nanoindentation data obtained at room temperature, Acta Mater.54 (2006) 5489e5499.

[19] M.P. Brady, Y. Yamamoto, M.L. Santella, P.J. Maziasz, B.A. Pint, C.T. Liu, Z.P. Lu,H. Bei, The development of alumina-forming austenitic stainless steels forhigh-temperature structural use, JOM 60 (2008) 12.

[20] D.J. Powell, R. Pilkington, D.A. Miller, The precipitation characteristics of 20Cr/25Ni-Nb stabilised stainless steel, Acta Metall. 36 (3) (1988) 713e724.

[21] T. Sourmail, H.K.D.H. Bhadeshia, Microstructural evolution in two variants ofNF709 at 1023 and 1073 K, Metall. Mater. Trans. A 36 (2005) 23.

[22] T.A. Allen, G.S. Was, Modeling radiation-induced segregation in austeniticFeeCreNi alloys, Acta Mater. 46 (1998) 3679e3691.

[23] W.C. Oliver, G.M. Pharr, An improved technique for determining hardness andelastic-modulus using load and displacement sensing indentation experi-ments, J. Mater. Res. 7 (1992) 1564e1583.

[24] S. Plimpton, Fast parallel algorithms for short-range molecular dynamics,J. Comp. Phys. 117 (1995) 1e19.

[25] G. Bonny, D. Terentyev, R. Pasianot, S. Ponc�e, A. Bakaev, Interatomic potentialto study plasticity in stainless steels: the FeNiCr model alloy, Model. Simul.Mater. Sci. Eng. 19(085008).

[26] A. Stukowski, Visualization and analysis of atomistic simulation data withOVITOethe open visualization tool, Model. Simul. Mater. Sci. Eng. 18 (2009)015012.

[27] I. Karaman, H. Sehitoglu, K. Gall, Y.I. Chumlyakov, H.J. Maier, Deformation ofsingle crystal hadfield steel by twinning and slip, Acta Mater. 48 (2000)1345e1359.

[28] C.A. Brookes, J.B. O'Neill, B.A.W. Redfern, Anisotropy in the hardness of singlecrystals, Proc. Roy. Soc. Lond. A 322 (1971) 73e88.

[29] L. Tan, G. Shaw, K. Sridharan, W.C. Crone, Effects of oxygen ion implantationon wear behavior of NiTi shape memory alloy, Mech. Mater. 37 (2005)1059e1068.

[30] No. 50A NIMS Creep Data Sheet, National Institute for Materials Science,Japan, 2015.

[31] G. Huang, D.K. Matlock, G. Krauss, Martensite formation, strain rate sensi-tivity, and deformation behavior of type 304 stainless steel sheet, Metall.Trans. A 20 (1989) 1239.

[32] U.F. Kocks, Laws for work-hardening and low-temperature creep, J. Eng.Mater. Technol. 98 (1976) 76e85.

[33] L. Remy, The interaction between slip and twinning systems and the influenceof twinning on the mechanical behavior of fcc metals and alloys, Metall. Trans.A 12 (1981) 387.

[34] G.H. Xiao, N.R. Tao, K. Lu, Effects of strain, strain rate and temperature ondeformation twinning in a Cu-Zn alloy, Scr. Mater. 59 (2008) 975e978, http://dx.doi.org/10.1016/j.scriptamat.2008.06.060.

[35] H.-J. Kestenbach, The effect of applied stress on partial dislocation separationand dislocation substructure in austenitic stainless steel, Philos. Mag. 36 (6)(1977) 1509e1515.

[36] T.S. Byun, E.H. Lee, J.D. Hunn, Plastic deformation in 316LN stainless steel echaracterization of deformation microstructures, J. Nucl. Mater. 321 (1) (2003)29e39, http://dx.doi.org/10.1016/s0022-3115(03)00195-8.

[37] R. Komanduri, N. Chandrasekaran, L.M. Raff, MD simulation of indentation andscratching of single crystal aluminum, Wear 240 (2000) 113e143.

[38] T.-H. Fang, S.-R. Jian, D.-S. Chuu, Molecular dynamics analysis of effects ofvelocity and loading on the nanoindentation, Jpn. J. Appl. Phys. 41 (2002) L1328eL 1331.

[39] G. Ziegenhain, A. Hartmaier, H.M. Urbassek, Pair vs many-body potentials:influence on elastic and plastic behavior in nanoindentation of fcc metals,J. Mech. Phys. Solids 57 (9) (2009) 1514e1526, http://dx.doi.org/10.1016/j.jmps.2009.05.011.

[40] R.E. Schramm, R.P. Reed, Stacking fault energies of seven commercialaustenitic stainless steels, Metall. Trans. A 6 (1975) 1345.

[41] J.P. Hirth, J. Lothe, Theory of Dislocations, McGraw Hill Book Company, NewYork, 1968.

[42] H.P. Karnthaler, E. Wintner, Weak beam TEM studies of composite disloca-tions in F.C.C. alloys, Acta Metall. 23 (1975) 1501e1509.

[43] A. Stukowski, K. Albe, Extracting dislocations and non-dislocation crystaldefects from atomistic simulation data, Model. Simul. Mater. Sci. Eng. 18(2010) 085001.