research papers\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoJOURNAL OF
SYNCHROTRON
RADIATION
ISSN: 1600-5775

Full strain tensor measurements with X-ray diffraction and strain field mapping: a simulation study

CROSSMARK_Color_square_no_text.svg

aThe Peac Institute of Multiscale Sciences, Chengdu, Sichuan 610031, People's Republic of China, bKey Laboratory of Advanced Technologies of Materials, Ministry of Education, and Institute of Material Dynamics, Southwest Jiaotong University, Chengdu, Sichuan 610031, People's Republic of China, and cEuropean XFEL GmbH, 22869 Schenefeld, Germany
*Correspondence e-mail: yyzhang@pims.ac.cn, sluo@pims.ac.cn

Edited by M. Yamamoto, RIKEN SPring-8 Center, Japan (Received 26 September 2019; accepted 18 March 2020; online 15 April 2020)

Strain tensor measurements are important for understanding elastic and plastic deformation, but full bulk strain tensor measurement techniques are still lacking, in particular for dynamic loading. Here, such a methodology is reported, combining imaging-based strain field mapping and simultaneous X-ray diffraction for four typical loading modes: one-dimensional strain/stress compression/tension. Strain field mapping resolves two in-plane principal strains, and X-ray diffraction analysis yields volumetric strain, and thus the out-of-plane principal strain. This methodology is validated against direct molecular dynamics simulations on nanocrystalline tantalum. This methodology can be implemented with simultaneous X-ray diffraction and digital image correlation in synchrotron radiation or free-electron laser experiments.

1. Introduction

The strain tensor is vital for characterizing mechanical properties of materials, understanding elastic and plastic deformation mechanisms, and developing physics-based constitutive models (Milathianaki et al., 2013[Milathianaki, D., Boutet, S., Williams, G. J., Higginbotham, A., Ratner, D., Gleason, A. E., Messerschmidt, M., Seibert, M. M., Swift, D. C., Hering, P., Robinson, J., White, W. E. & Wark, J. S. (2013). Science, 342, 220-223.]; Allais et al., 1994[Allais, L., Bornert, M., Bretheau, T. & Caldemaison, D. (1994). Acta Metall. Mater. 42, 3865-3880.]; Poulsen et al., 2005[Poulsen, H. F., Wert, J. A., Neuefeind, J., Honkimäki, V. & Daymond, M. (2005). Nat. Mater. 4, 33-36.]; Zhu et al., 2017[Zhu, Y., Kang, G. & Yu, C. (2017). Int. J. Plast. 95, 191-215.]). However, it is still a challenge to measure the full bulk strain tensor, especially for fast or ultrafast dynamic experiments. The development of advanced X-ray sources such as synchrotron radiation sources and X-ray free-electron lasers (XFELs) offers the opportunity to capture in situ dynamic events (Luo et al., 2012[Luo, S. N., Jensen, B. J., Hooks, D. E., Fezzaa, K., Ramos, K. J., Yeager, J. D., Kwiatkowski, K. & Shimada, T. (2012). Rev. Sci. Instrum. 83, 073903.]; Milathianaki et al., 2013[Milathianaki, D., Boutet, S., Williams, G. J., Higginbotham, A., Ratner, D., Gleason, A. E., Messerschmidt, M., Seibert, M. M., Swift, D. C., Hering, P., Robinson, J., White, W. E. & Wark, J. S. (2013). Science, 342, 220-223.]; Wehrenberg et al., 2017[Wehrenberg, C. E., McGonegle, D., Bolme, C., Higginbotham, A., Lazicki, A., Lee, H. J., Nagler, B., Park, H. S., Remington, B. A., Rudd, R. E., Sliwa, M., Suggit, M., Swift, D., Tavella, F., Zepeda-Ruiz, L. & Wark, J. S. (2017). Nature, 550, 496-499.]; Seiboth et al., 2018[Seiboth, F., Fletcher, L. B., McGonegle, D., Anzellini, S., Dresselhaus-Cooper, L. E., Frost, M., Galtier, E., Goede, S., Harmand, M., Lee, H. J., Levitan, A. L., Miyanishi, K., Nagler, B., Nam, I., Ozaki, N., Rödel, M., Schropp, A., Spindloe, C., Sun, P., Wark, J. S., Hastings, J., Glenzer, S. H. & McBride, E. E. (2018). Appl. Phys. Lett. 112, 221907.]; Coleman et al., 2019[Coleman, A. L., Gorman, M. G., Briggs, R., McWilliams, R. S., McGonegle, D., Bolme, C. A., Gleason, A. E., Fratanduono, D. E., Smith, R. F., Galtier, E., Lee, H. J., Nagler, B., Granados, E., Collins, G. W., Eggert, J. H., Wark, J. S. & McMahon, M. I. (2019). Phys. Rev. Lett. 122, 255704.]; Brown et al., 2019[Brown, S. B., Gleason, A. E., Galtier, E., Higginbotham, A., Arnold, B., Fry, A., Granados, E., Hashim, A., Schroer, C. G., Schropp, A., Seiboth, F., Tavella, F., Xing, Z., Mao, W., Lee, H. J. & Nagler, B. (2019). Sci. Adv. 5, eaau8044.]; Huang et al., 2016a[Huang, J. W., E, J. C., Huang, J. Y., Sun, T., Fezzaa, K. & Luo, S. N. (2016a). J. Synchrotron Rad. 23, 712-717.],b[Huang, J. Y., Huang, J. W., Sun, T., Fezzaa, K., Xu, S. L. & Luo, S. N. (2016b). Acta Mater. 114, 136-145.]; Fan et al., 2016[Fan, D., Huang, J. W., Zeng, X. L., Li, Y. E. J. C., Huang, J. Y., Sun, T., Fezzaa, K., Wang, Z. & Luo, S. N. (2016). Rev. Sci. Instrum. 87, 053903.]; Turneaure et al., 2017[Turneaure, S. J., Sharma, S. M., Volz, T. J., Winey, J. M. & Gupta, Y. M. (2017). Sci. Adv. 3, eaao3561.]; Briggs et al., 2019[Briggs, R., Torchio, R., Sollier, A., Occelli, F., Videau, L., Kretzschmar, N. & Wulff, M. (2019). J. Synchrotron Rad. 26, 96-101.]), including strain and X-ray diffraction measurements.

High-speed X-ray phase contrast imaging (Lu et al., 2017[Lu, L., Sun, T. K. K. F., Fezzaa, K., Gong, X. L. & Luo, S. N. (2017). Mater. Sci. Eng. A, 701, 143-148.]; Huang et al., 2016b[Huang, J. Y., Huang, J. W., Sun, T., Fezzaa, K., Xu, S. L. & Luo, S. N. (2016b). Acta Mater. 114, 136-145.]) is utilized to obtain two-dimensional (2D) strain fields via digital image correlation (DIC) (Lu et al., 2014[Lu, L., Fan, D., Bie, B. X., Ran, X. X., Qi, M. L., Parab, N., Sun, J. Z., Liao, H. J., Hudspeth, M. C., Claus, B., Fezzaa, K., Sun, T., Chen, W., Gong, X. L. & Luo, S. N. (2014). Rev. Sci. Instrum. 85, 076101.], 2016[Lu, L., Huang, J. W., Fan, D., Bie, B. X., Sun, T., Fezzaa, K., Gong, X. L. & Luo, S. N. (2016). Acta Mater. 120, 86-94.]). However, the 2D strain field is only mapped on the plane perpendicular to the incident X-ray beam (in-plane strain components), and strains along the incident direction cannot be resolved. On the other hand, high-speed X-ray diffraction usually provides lattice strain measurements (Ma et al., 2017[Ma, L., Wang, L., Nie, Z., Wang, F., Xue, Y., Zhou, J., Cao, T., Wang, Y. & Ren, Y. (2017). Acta Mater. 128, 12-21.]; Zhang et al., 2017[Zhang, M., Li, L., Ding, J., Wu, Q., Wang, Y.-D., Almer, J., Guo, F. & Ren, Y. (2017). Acta Mater. 141, 294-303.]; Collins et al., 2015[Collins, D. M., Mostafavi, M., Todd, R. I., Connolley, T. & Wilkinson, A. J. (2015). Acta Mater. 90, 46-58.]).

Given the 2D nature of in-plane strain field mapping, the missing bulk strain components have to be obtained via other means such as X-ray diffraction. X-ray Laue microdiffraction or Laue DIC (Petit et al., 2015[Petit, J., Castelnau, O., Bornert, M., Zhang, F. G., Hofmann, F., Korsunsky, A. M., Faurie, D., Le Bourlot, C., Micha, J. S., Robach, O. & Ulrich, O. (2015). J. Synchrotron Rad. 22, 980-994.]), suitable for up to hundreds of grains, was applied to measure the local strain tensor, but these quasi-static techniques can hardly be used for highly transient events. A recent study (Zhang et al., 2019[Zhang, Y. Y., Tang, M. X., Cai, Y., E, J. C. & Luo, S. N. (2019). J. Synchrotron Rad. 26, 413-421.]) demonstrated that equivalent volumetric strain can be obtained under 1D strain loading conditions or planar shock compression via analysis of 2D diffraction patterns; however, all the principal strain components cannot be resolved. It is highly desirable to develop techniques to obtain the full bulk strain tensor in dynamic experiments.

For bulk strain tensor measurements (in particular under short time scale dynamic loading), we propose a methodology combining 2D strain mapping and X-ray analysis of 2D diffraction patterns: the former resolves two in-plane principal strains via imaging, and the latter volumetric strain (thus the out-of-plane principal strain). We validate this methodology against direct molecular dynamics (MD) simulations on nanocrystalline tantalum under four typical loading modes. This methodology can be implemented with simultaneous X-ray diffraction and DIC in synchrotron radiation or free-electron laser experiments.

This article is organized as follows. Section 2[link] addresses the methodology of simulated experimental geometry, MD simulations, strain field mapping and the Singh analysis of 2D diffraction patterns; Section 3[link] presents results and discussion, validating the strain measurements against direct MD simulations, and Section 4[link] presents conclusion.

2. Methodology

2.1. Simulated experimental geometry

The simulation geometry for simultaneous diffraction and imaging measurements is presented in Fig. 1[link] as in experiments (Lu et al., 2016[Lu, L., Huang, J. W., Fan, D., Bie, B. X., Sun, T., Fezzaa, K., Gong, X. L. & Luo, S. N. (2016). Acta Mater. 120, 86-94.]). The loading direction is along the x-axis, while the incident X-ray beam is along the y-axis. We use the transmission geometry for diffraction simulations. Here a 2D detector is set perpendicular to the incident X-ray direction (the normal detector position); for an arbitrary detector position, a geometrical correction can be applied.

[Figure 1]
Figure 1
Schematic illustrating simultaneous X-ray imaging (for strain mapping) and diffraction in the transmission mode. DIC: digital image correlation. 2ϑ: diffraction angle; χ: azimuthal angle.

In experiments, the strain field is mapped onto the plane perpendicular to the incident X-ray beam (the y-axis), normally done via X-ray DIC with phase contrast imaging (Lu et al., 2014[Lu, L., Fan, D., Bie, B. X., Ran, X. X., Qi, M. L., Parab, N., Sun, J. Z., Liao, H. J., Hudspeth, M. C., Claus, B., Fezzaa, K., Sun, T., Chen, W., Gong, X. L. & Luo, S. N. (2014). Rev. Sci. Instrum. 85, 076101.]). Here we obtain strain fields on the xz-plane from atomic positions in MD trajectories (see below), to mimic X-ray DIC measurement.

2.2. Molecular dynamics simulations

For MD simulations, we use the large-scale Atomic/Molecular Massively Parallel Simulator (Plimpton, 1995[Plimpton, S. (1995). J. Comput. Phys. 117, 1-19.]) and a widely used embedded-atom method (EAM) potential (Ravelo et al., 2013[Ravelo, R., Germann, T. C., Guerrero, O., An, Q. & Holian, B. L. (2013). Phys. Rev. B, 88, 134101.]) for tantalum. This EAM potential is accurate for describing the properties of tantalum as shown by numerous shock (Wang et al., 2014[Wang, L., Zhao, F., Zhao, F. P., Cai, Y., An, Q. & Luo, S. N. (2014). J. Appl. Phys. 115, 053528.]; Tang et al., 2018[Tang, M. X., Zhang, Y. Y., E, J. C. & Luo, S. N. (2018). J. Synchrotron Rad. 25, 748-756.]; Hahn et al., 2017[Hahn, E. N., Germann, T. C., Ravelo, R., Hammerberg, J. E. & Meyers, M. A. (2017). Acta Mater. 126, 313-328.]) and nonshock simulations (Remington et al., 2014[Remington, T. P., Ruestes, C. J., Bringa, E. M., Remington, B. A., Lu, C. H., Kad, B. & Meyers, M. A. (2014). Acta Mater. 78, 378-393.]; Huang et al., 2017[Huang, C., Peng, X., Fu, T., Chen, X., Xiang, H., Li, Q. & Hu, N. (2017). Mater. Sci. Eng. A, 700, 609-616.]).

We construct a nanocrystalline tantalum configuration with dimensions of about 200 nm × 40 nm × 40 nm, corresponding to 17.5 million atoms. This configuration contains about 2500 grains, and the average grain size is about 5 nm. Four loading modes commonly encountered in experiments are simulated: compression and tension under 1D strain conditions as in planar shock compression, and compression and tension under 1D stress conditions as in materials testing systems, referred to as 1D-strain compression, 1D-strain tension, 1D-stress compression, and 1D-stress tension, respectively. Prior to loading, the configuration is relaxed at 0 K, and then annealed with the constant-pressure–temperature ensemble under 3D periodic boundary conditions at 300 K and zero pressure (Wang et al., 2016[Wang, L., Cai, Y., He, A. M. & Luo, S. N. (2016). Phys. Rev. B, 93, 174106.]).

For 1D-strain compression/tension loading, the relaxed nanocrystalline configuration is compressed/stretched along the x-axis, and the microcanonical ensemble is used to mimic adiabatic loading as in shock compression. The dimension along the x-axis is reduced/increased at a fixed decrement/increment, corresponding to a strain rate of 109 s−1 (Spearot et al., 2007[Spearot, D. E., Tschopp, M. A., Jacob, K. I. & Mcdowell, D. L. (2007). Acta Mater. 55, 705-714.]; Tschopp & McDowell, 2008[Tschopp, M. A. & McDowell, D. L. (2008). J. Mech. Phys. Solids, 56, 1806-1830.]). For 1D-stress compression/tension loading, the configuration dimension along the x-axis is changed under a constant strain rate of 109 s−1, while the stress-free boundary condition is applied along the y- and z-axes (Spearot et al., 2007[Spearot, D. E., Tschopp, M. A., Jacob, K. I. & Mcdowell, D. L. (2007). Acta Mater. 55, 705-714.]). The temperature is set at 300 K.

All the loading simulations are conducted under 3D periodic boundary conditions. The time step for integration of the equation of motion is 1 fs. Physical quantities such as particle velocity, density, stress, and temperature can be obtained from trajectories via binning analysis (Luo et al., 2009[Luo, S. N., An, Q., Germann, T. C. & Han, L. B. (2009). J. Appl. Phys. 106, 013502.]).

Stress tensor can be decomposed into hydrostatic and deviatoric stress components as

[\left(\matrix{\sigma _{{xx}}&0&0\cr 0&\sigma _{{yy}}&0\cr 0&0&\sigma _{{zz}}}\right) = \sigma _{{\rm p}}{\bf I}+\left(\matrix{{{2} \over {3}}t&0&0\cr 0&-{{1} \over {3}}t&0\cr 0&0&-{{1} \over {3}}t}\right),]

where I is the identity matrix, σxxσyy = σzz for homogeneous isotropic solids, [{\sigma}_{{\rm p}}] = [({{1}/{3}}){(\sigma_{{xx}}+\sigma_{{yy}}+\sigma_{{zz}})}] represents the mean normal stress or so-called equivalent hydrostatic stress, and t = [\sigma_{{xx}}-({{1}/{2}}){(\sigma_{{yy}}+\sigma_{{zz}})}] denotes the differential stress.

The bulk strain tensor, ij (i, j = x, y, z), can be obtained from the dimension (Li) and shape changes of a whole simulation cell (Method I). The normal engineering strains are determined from the dimension changes as

[\varepsilon _{{ii}} = {{\Delta L_{i}} \over {L_{{i}}}}\qquad {\rm with}\quad i = x,y,z. \eqno(1)]

Shear strain components are the change of the angles between edges of the whole simulation cell, and equal zero for the four common loading modes investigated in this work.

2.3. Strain field mapping and analysis

For strain field mapping, we take two atomic configurations from MD simulations, the reference (0 ps; undeformed) and current configurations (e.g. 80 ps; deformed). For each atom m, we calculate the atomic strain tensor (Stukowski, 2010[Stukowski, A. (2010). Modell. Simul. Mater. Sci. Eng. 18, 015012.]) as [\varepsilon^{m}_{{ij}}] = (1/2)(umi,j+umj,i) (i, j = x, y, z). um is the atomic displacement of the mth atom and [\varepsilon^{m}_{{ij}}] essentially represents the 3D strain field around this atom.

The average atomic strain tensor for N atoms or bulk strain tensor follows as

[\varepsilon _{{ij}} = {{1} \over {2N}} \sum\limits_{{m\,=\,1}}^{N} \left({u_{{i,j}}^{m}+u_{{j,i}}^{m}}\right). \eqno(2)]

This method for calculating bulk engineering strain is referred to as Method II. Then we calculate the volumetric strain (p) as follows,

[\varepsilon _{{\rm p}} = \left(1+\varepsilon _{{xx}}\right)\left(1+\varepsilon _{{yy}}\right)\left(1+\varepsilon _{{zz}}\right)-1. \eqno(3)]

In experiments, only the 2D engineering strain tensor (Method III) can be obtained via the projection along an imaging direction. In addition, the strain tensor is often expressed in terms of the Green–Lagrange strain tensor [Lu et al., 2014[Lu, L., Fan, D., Bie, B. X., Ran, X. X., Qi, M. L., Parab, N., Sun, J. Z., Liao, H. J., Hudspeth, M. C., Claus, B., Fezzaa, K., Sun, T., Chen, W., Gong, X. L. & Luo, S. N. (2014). Rev. Sci. Instrum. 85, 076101.]; Stukowski, 2010[Stukowski, A. (2010). Modell. Simul. Mater. Sci. Eng. 18, 015012.]]. We thus also calculate the Green–Lagrange strain for each atom m,

[E_{{ij}}^{\,m} = {{1} \over {2}} \left(u^{m}_{{i,j}}+u^{m}_{{j,i}}+u^{m}_{{k,i}}u^{m}_{{k,j}}\right), \eqno(4)]

with i, j, k = x, y, z. Eij m is then projected along the y-axis to obtain the 2D Green–Lagrange strain field. Note that Eij m and [\varepsilon^{m}_{{ij}}] are equivalent for describing strain field, and can be converted from one to the other (Lu et al., 2014[Lu, L., Fan, D., Bie, B. X., Ran, X. X., Qi, M. L., Parab, N., Sun, J. Z., Liao, H. J., Hudspeth, M. C., Claus, B., Fezzaa, K., Sun, T., Chen, W., Gong, X. L. & Luo, S. N. (2014). Rev. Sci. Instrum. 85, 076101.]; Stukowski, 2010[Stukowski, A. (2010). Modell. Simul. Mater. Sci. Eng. 18, 015012.]).

2.4. X-ray diffraction simulations and analysis

Atomic configurations from MD simulations are used as input to diffraction calculations with GAPD, a GPU-accelerated Atom-based Polychromatic Diffraction simulation code (E et al., 2018[E, J. C., Wang, L., Chen, S., Zhang, Y. Y. & Luo, S. N. (2018). J. Synchrotron Rad. 25, 604-611.]). The scattered X-ray intensity is calculated in 3D reciprocal space, and then projected onto a 2D detector to obtain diffraction patterns as in experiments. Diffraction angle (2ϑ) and azimuthal angle (χ) are defined in Fig. 1[link], and the loading direction (the x-axis) forms an angle of θ with the normal of a diffracting plane under consideration. θ is referred to as the loading–diffraction geometry angle (Zhang et al., 2019[Zhang, Y. Y., Tang, M. X., Cai, Y., E, J. C. & Luo, S. N. (2019). J. Synchrotron Rad. 26, 413-421.]), and can be varied as desired. There exists the following relation among θ, 2ϑ, and χ under simulated geometry,

[\cos{{\theta}} = \cos{{\vartheta}}\cos{{\chi}}. \eqno(5)]

The X-ray wavelength in our calculations is 0.512 Å (24.22 keV), the peak intensity wavelength of the undulator source with a period of 18 mm and a gap of 12 mm at 32ID-B of the Advanced Photon Source. Simulated 2ϑ ranges from 10° to 30°, and the corresponding range of θ is 5°–90°.

From 2D X-ray diffraction patterns, we can deduce the bulk density under the equivalent hydrostatic pressure (σp) and residual strength (Singh, 1993[Singh, A. K. (1993). J. Appl. Phys. 73, 4278-4286.]; Zhang et al., 2019[Zhang, Y. Y., Tang, M. X., Cai, Y., E, J. C. & Luo, S. N. (2019). J. Synchrotron Rad. 26, 413-421.]). The methodology has been demonstrated to be appropriate for 1D-strain compression including shock compression (Zhang et al., 2019[Zhang, Y. Y., Tang, M. X., Cai, Y., E, J. C. & Luo, S. N. (2019). J. Synchrotron Rad. 26, 413-421.]). The measured lattice spacing, dm(hkl), is related to θ via

[d_{{\rm m}}(hkl) = d_{{\rm p}}(hkl) \left[1+\left(1-3\cos^{2}{\theta}\right) Q(hkl)\right]. \eqno(6)]

Here, dp(hkl) denotes d-spacing due to σp, and Q(hkl) is a factor which depends on t and single-crystal elastic compliances. Since dm = [\lambda/(2\sin{\vartheta}]), we obtain a scatter plot of dm versus [(1-3\cos^{2}{\theta})], which can then be fitted with equation (6)[link]. More analysis details have been presented elsewhere (Zhang et al., 2019[Zhang, Y. Y., Tang, M. X., Cai, Y., E, J. C. & Luo, S. N. (2019). J. Synchrotron Rad. 26, 413-421.]).

Given dp(hkl) determined from 2D diffraction patterns, we have p = [dp(hkl)/d0(hkl)]3 − 1. Since xx and zz are evaluated from strain field mapping, the unknown tensor component yy can be solved with equation (3)[link], and the full strain tensor, determined by virtue of a combination of 2D strain field mapping (imaging) and diffraction patterns (Method IV). Method IV is experimentally feasible for full strain tensor measurements.

3. Results and discussion

3.1. MD simulations

We explore the four most common loading modes in high-pressure experiments and mechanical testing: 1D-strain compression, 1D-strain tension, 1D-stress compression, and 1D-stress tension. The initial configurations are the same for these loading modes. The stress–strain curves directly obtained from MD simulations are shown in Fig. 2[link], including normal stress (σxx), equivalent hydrostatic stress (σp), and differential stress (t). As an example, the full strain tensor ij and volumetric strain p at xx = 8.00%, calculated directly from the simulation cell dimensions/shapes (Method I), are presented in Table 1[link] for the four loading modes.

Table 1
Bulk strain tensor (∊ij) and volumetric strain (∊p) for the four loading modes with ∊xx = 8.00%

The off-diagonal values of ij are zero and thus not shown. Method I: based on dimension and shape changes of a simulation cell; Method II: based on 3D strain field mapping; Method III: based on 2D strain mapping; Method IV: based on 2D strain field mapping combined with analysis of 2D X-ray diffraction patterns.

  Strain (%)
  Method I Method II Method III Method IV
1D strain compression
xx −8.00 −8.08 −8.08 −8.08
yy 0.00 0.16 0.16
zz 0.00 0.16 0.16 0.16
p −8.00 −7.79 −7.79
 
1D strain tension
xx 8.00 7.90 7.90 7.90
yy 0.00 −0.10 0.10
zz 0.00 −0.10 −0.10 −0.10
p 8.00 7.68 7.90
 
1D stress compression
xx −8.00 −8.06 −8.06 −8.06
yy 3.80 3.62 4.19
zz 3.82 3.65 3.65 3.65
p −0.86 −1.25 −0.71
 
1D stress tension
xx 8.00 7.80 7.80 7.80
yy −3.28 −3.40 −2.82
zz −3.35 −3.46 −3.46 −3.46
p 0.96 0.53 1.14
[Figure 2]
Figure 2
Bulk stress–strain curves for four representative loading modes: (a) 1D-strain compression, (b) 1D-strain tension, (c) 1D-stress compression and (d) 1D-stress tension. For better visualization, the absolute values of tensile stresses are used for plotting.

For 1D-strain compression [Fig. 2(a)[link]], σxx increases to about 71.0 GPa, and σp increases to about 68.0 GPa as bulk loading strain xx increases to 20.00%, while t increases to 5.2 GPa at xx = 9.00% and then decreases to 4.2 GPa due to yield. σp is slightly less than σxx. For 1D-strain tension [Fig. 2(b)[link]], as strain xx reaches 9.00%, voids nucleate in the sample and tensile failure occurs. σxx and σp reach their peak values (15.2 GPa and 12.5 GPa, respectively) at about xx = 9.50%, while t peaks at xx = 8.50% (4.4 GPa).

Different from the 1D-strain compression, σxx in 1D-stress compression [Fig. 2(c)[link]] reaches a much lower peak value of 5.1 GPa at xx = 6.00%, and t is the same as σxx due to the lack of lateral confinement (stress-free along the y- and z-axes), and [\sigma _{{\rm p}}] = [({{1}/{3}})\sigma_{{xx}}]. For 1D-stress tension [Fig. 2(d)[link]], σxx and t reach a peak stress of 5.0 GPa at xx = 6.00%, and [\sigma_{{\rm p}}] = [({{1}/{3}})\sigma_{{xx}}].

3.2. Strain field mapping

3D strain field maps are obtained from two atomic configurations (Method II). Since the Green–Lagrange strain field is measured with X-ray phase contrast imaging in experiments, we also calculate 2D Green–Lagrange strain field maps via 2D binning along the x- and z-axes of the 3D strain maps. We select the central regions of about 300 Å × 300 Å for visualization (Fig. 3[link]; bulk strain xx = 8.00%).

[Figure 3]
Figure 3
Green–Lagrange strain fields of Exx for four deformation modes with bulk strain component xx = 8.00%: (a) 1D-strain compression, (b) 1D-strain tension, (c) 1D-stress compression, and (d) 1D-stress tension. Color coding is based on Exx.

Strain is more localized at grain boundaries than in grain interiors, and under 1D-stress loading than under 1D-strain loading (Fig. 3[link]). Strain inhomogeneity and strain at grain boundaries are higher under tension than under compression [Figs. 3(b) and 3(d)[link]].

From 3D and 2D strain maps, we can calculate bulk strain tensor components ij, and volumetric strain (p). Methods II and III yield accurate results compared with Method I (Table 1[link]). However, yy and p cannot be obtained with Method III, which on the other hand can be obtained from analysis of 2D X-ray diffraction patterns.

3.3. X-ray diffraction

1D X-ray diffraction profiles are most commonly used to obtain d-spacing and lattice strain of corresponding diffracting planes. The {110} lattice strains are calculated from the 1D diffraction curves for the four loading modes (Fig. 4[link]), and listed in Table 2[link].

Table 2
Diffraction angles (2ϑ), interplanar spacings (d) and lattice strains (∊l) for the {110} diffraction peak calculated from 1D diffraction profiles for bulk strain ∊xx = 8.00%

Subscript 0 refers to undeformed.

Loading modes 2ϑ0 2ϑ d0 (Å) d (Å) l
1D-strain compression 12.57° 12.96° 2.338 2.268 −3.00%
1D-strain tension 12.57° 12.18° 2.338 2.413 3.21%
1D-stress compression 12.57° 12.65° 2.338 2.324 −0.60%
1D-stress tension 12.57° 12.42° 2.338 2.367 1.24%
[Figure 4]
Figure 4
{110} diffraction profiles for (a) 1D-strain compression and tension, and (b) 1D-stress compression and tension.

However, lattice strain is not necessarily equal to bulk strain; their relation depends on sample microstructure, diffraction geometry, and stress conditions. For example, for the same bulk strain xx = 8.00%, the shift of 2ϑ is much bigger under 1D-strain loading than that under 1D-stress loading (Table 2[link]). If we assume hydrostatic deformation, the bulk volumetric strains [p = (d/d0)3 − 1] would be −8.70%, 9.94%, −1.79% and 3.77%, for 1D-strain compression and tension, and 1D-stress compression and tension, respectively. The corresponding bulk volumetric strains directly obtained from MD simulations are considerably different, being −8.00%, 8.00%, −0.86% and 0.96%. Thus, such simplified analyses of 1D diffraction curves are inappropriate.

2D diffraction patterns allow one to extract more information with high accuracy on the basis of the Singh theory (Singh, 1993[Singh, A. K. (1993). J. Appl. Phys. 73, 4278-4286.]), including differential stress (or residual strength) and equivalent hydrostatic strain under nonhydrostatic conditions (Zhang et al., 2019[Zhang, Y. Y., Tang, M. X., Cai, Y., E, J. C. & Luo, S. N. (2019). J. Synchrotron Rad. 26, 413-421.]). Fig. 5(a)[link] shows three major diffraction rings, {110}, {200}, and {112}, for 1D-strain compression with xx = 8.00%. The strongest {110} ring is used for the analysis with the Singh theory, and the results are shown in Fig. 5(c)[link]. We obtain dp = 2.276 Å, and thus the equivalent hydrostatic volumetric strain is −7.79%. The residual strength is obtained to be t = 5.34 GPa, with the aggregate shear modulus G calculated directly from MD simulations. The corresponding values directly from MD simulations are p = −8.00% and t = 5.19 GPa (Table 3[link]), which agree with the analysis of 2D diffraction patterns. For other loading modes, the agreement between direct MD simulations and the analysis of 2D diffraction patterns (Table 3[link]) is also reasonable. Thus, the Singh analysis can be used for obtaining bulk volumetric strain for different loading modes.

Table 3
Fitting to full {110} diffraction rings for ∊xx = 8.00%

Subscripts 1 and 2 refer to MD results and the Singh analysis of diffraction patterns, respectively.

Loading modes dp (Å) Q t1 (GPa) t2 (GPa) [\varepsilon _{{\rm p_{1}}}] [\varepsilon _{{\rm p_{2}}}]
1D-strain compression 2.276 0.0157 5.19 5.34 −8.00% −7.79%
1D-strain tension 2.399 −0.0146 −4.38 −4.48 8.00% 7.90%
1D-stress compression 2.333 0.0151 4.98 4.99 −0.86% −0.71%
1D-stress tension 2.348 −0.0153 −4.90 −4.71 0.96% 1.14%
[Figure 5]
Figure 5
Diffraction analysis with full and partial diffraction rings. For the latter, χ spans over 0–45° and 315°–360°. (a) Full diffraction rings of tantalum under uniaixal strain compression (xx = 0.08), and (b) corresponding partial diffraction rings. White curves: deformed (D); black curves: undeformed (U). (c) dm versus [(1-3\cos^{2}{\theta})] plot for the full {110} diffraction ring. Red dots: `data;' black line: linear fitting. (d) Comparison of the dm versus [(1-3\cos^{2}{\theta})] plots for fitting with full (black curve) and partial (green) {110} diffraction rings.

In certain experiments, it is unfeasible to obtain full diffraction rings due to physical constraints. We explore below whether/how partial diffraction rings can be used for accurate measurements. As an example, we select 1/4 of the {110} diffraction rings (χ spans over 0–45° and 315°–360°) for analysis [Fig. 5(b)[link]]. The Singh analysis [Fig. 5(d)[link]] yields a volumetric strain of about −7.14% and residual strength of about 6.27 GPa, as opposed to −7.79% and 5.34 GPa obtained from the full diffraction rings. The significant differences are due to the fact that the θ-range corresponding to the selected χ-range is 5°–45° [Fig. 6(a)[link]] and not representative. Similarly, for 45° ≤ χ < 135°, θ ranges from 45° to 90°; as a result, the dm versus [1-3\cos^{2}{\theta}] scatter plot only spans a partial range, so the linear fitting is not well constrained [Fig. 6(b)[link]]. On the other hand, a properly chosen χ-range, e.g. 0°–90°, can cover a much wider/whole θ-range (5°–90°). Consequently, the fitting is much better and in agreement with that using the full diffraction ring [Fig. 6(c)[link]].

[Figure 6]
Figure 6
(a) Distribution of θ on the detector plane for the simulated geometry. (b) Fitting for the {110} diffraction ring with 45° ≤ χ ≤ 135°. (c) Fitting for the {110} diffraction ring with 0° ≤ χ ≤ 90°. Black and green lines: full and partial diffraction rings, respectively; red dots: `data.'

With a properly chosen χ-range, we analyze partial {110} diffraction rings (0° ≤ χ ≤ 90°, Fig. 7[link]) for four loading modes with xx = 8.00%. Their corresponding fittings with equation (6)[link] show different features (shapes and shifts in 2ϑ). The fitting results of t and p (Table 4[link]) show that the direct simulations agree with the Singh analysis for these four loading modes. Note that, for 1D-stress loading, equivalent hydrostatic strains are much smaller compared with the applied strain xx = 8.00% due to the Poisson's effect.

Table 4
Fitting to partial {110} diffraction rings for ∊xx = 8.00%

Subscripts 1 and 2 refer to MD results and the Singh analysis of diffraction patterns, respectively.

Loading modes dp (Å) Q t1 (GPa) t2 (GPa) [\varepsilon_{{\rm p_{1}}}] [\varepsilon_{{\rm p_{2}}}]
1D-strain compression 2.279 0.0164 5.19 5.58 −8.00% −7.41%
1D-strain tension 2.396 −0.0142 −4.38 −4.35 8.00% 7.49%
1D-stress compression 2.334 0.0157 4.98 5.17 −0.86% −0.55%
1D-stress tension 2.346 −0.0164 −4.90 −5.05 0.96% 0.99%
[Figure 7]
Figure 7
{110} diffraction rings of nanocrystalline tantalum under (a) 1D-strain compression, (b) 1D-strain tension, (c) 1D-stress compression, and (d) 1D-stress tension loading. Black and white curves: fittings to simulated diffraction rings for the undeformed and deformed (xx = 8.00%) cases, respectively.

The Singh analysis of partial/full 2D diffraction patterns yields bulk volumetric strain (equivalent hydrostatic strain, p). In addition, two strain components xx and zz can be obtained from strain field mapping, e.g. via X-ray DIC. Then, yy can be calculated with equation (3)[link], and the full strain tensor obtained for these loading modes.

In experiments, strain fields can be obtained via DIC with X-ray phase contrast imaging or optical imaging, and along with simultaneous X-ray diffraction, full strain tensors can be obtained for the most common loading modes. Both imaging and diffraction can be carried out at the single-bunch level at synchrotrons or XFELs (Luo et al., 2012[Luo, S. N., Jensen, B. J., Hooks, D. E., Fezzaa, K., Ramos, K. J., Yeager, J. D., Kwiatkowski, K. & Shimada, T. (2012). Rev. Sci. Instrum. 83, 073903.]; Lu et al., 2016[Lu, L., Huang, J. W., Fan, D., Bie, B. X., Sun, T., Fezzaa, K., Gong, X. L. & Luo, S. N. (2016). Acta Mater. 120, 86-94.]; Seiboth et al., 2018[Seiboth, F., Fletcher, L. B., McGonegle, D., Anzellini, S., Dresselhaus-Cooper, L. E., Frost, M., Galtier, E., Goede, S., Harmand, M., Lee, H. J., Levitan, A. L., Miyanishi, K., Nagler, B., Nam, I., Ozaki, N., Rödel, M., Schropp, A., Spindloe, C., Sun, P., Wark, J. S., Hastings, J., Glenzer, S. H. & McBride, E. E. (2018). Appl. Phys. Lett. 112, 221907.]). Thus, Method IV can be implemented for in situ, real-time, dynamic synchrotron or XFEL experiments.

4. Conclusions

We have developed a methodology of bulk strain tensor measurements for four typical loading modes, 1D strain/stress compression/tension, via a combination of imaging-based strain field mapping and simultaneous X-ray diffraction. Strain field mapping measures two in-plane principal strains (xx and zz); X-ray diffraction analysis yields volumetric strain, and thus the out-of-plane principal strain (yy). This methodology is validated against direct molecular dynamics simulations on nanocrystalline tantalum, and can be implemented with simultaneous X-ray diffraction and digital image correlation in synchrotron radiation or free-electron laser experiments.

Acknowledgements

The computations were performed at the Supercomputing Center of the Peac Institute of Multiscale Sciences.

Funding information

Funding for this research was provided by: Scientific Challenge Project of China (grant No. TZ2018001); National NaturalScience Fundation of China (grant No. 11627901).

References

First citationAllais, L., Bornert, M., Bretheau, T. & Caldemaison, D. (1994). Acta Metall. Mater. 42, 3865–3880.  CrossRef CAS Web of Science Google Scholar
First citationBriggs, R., Torchio, R., Sollier, A., Occelli, F., Videau, L., Kretzschmar, N. & Wulff, M. (2019). J. Synchrotron Rad. 26, 96–101.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationBrown, S. B., Gleason, A. E., Galtier, E., Higginbotham, A., Arnold, B., Fry, A., Granados, E., Hashim, A., Schroer, C. G., Schropp, A., Seiboth, F., Tavella, F., Xing, Z., Mao, W., Lee, H. J. & Nagler, B. (2019). Sci. Adv. 5, eaau8044.  Web of Science CrossRef PubMed Google Scholar
First citationColeman, A. L., Gorman, M. G., Briggs, R., McWilliams, R. S., McGonegle, D., Bolme, C. A., Gleason, A. E., Fratanduono, D. E., Smith, R. F., Galtier, E., Lee, H. J., Nagler, B., Granados, E., Collins, G. W., Eggert, J. H., Wark, J. S. & McMahon, M. I. (2019). Phys. Rev. Lett. 122, 255704.  Web of Science CrossRef PubMed Google Scholar
First citationCollins, D. M., Mostafavi, M., Todd, R. I., Connolley, T. & Wilkinson, A. J. (2015). Acta Mater. 90, 46–58.  Web of Science CrossRef CAS Google Scholar
First citationE, J. C., Wang, L., Chen, S., Zhang, Y. Y. & Luo, S. N. (2018). J. Synchrotron Rad. 25, 604–611.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationFan, D., Huang, J. W., Zeng, X. L., Li, Y. E. J. C., Huang, J. Y., Sun, T., Fezzaa, K., Wang, Z. & Luo, S. N. (2016). Rev. Sci. Instrum. 87, 053903.  Web of Science CrossRef PubMed Google Scholar
First citationHahn, E. N., Germann, T. C., Ravelo, R., Hammerberg, J. E. & Meyers, M. A. (2017). Acta Mater. 126, 313–328.  Web of Science CrossRef CAS Google Scholar
First citationHuang, C., Peng, X., Fu, T., Chen, X., Xiang, H., Li, Q. & Hu, N. (2017). Mater. Sci. Eng. A, 700, 609–616.  Web of Science CrossRef CAS Google Scholar
First citationHuang, J. W., E, J. C., Huang, J. Y., Sun, T., Fezzaa, K. & Luo, S. N. (2016a). J. Synchrotron Rad. 23, 712–717.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationHuang, J. Y., Huang, J. W., Sun, T., Fezzaa, K., Xu, S. L. & Luo, S. N. (2016b). Acta Mater. 114, 136–145.  Web of Science CrossRef CAS Google Scholar
First citationLu, L., Fan, D., Bie, B. X., Ran, X. X., Qi, M. L., Parab, N., Sun, J. Z., Liao, H. J., Hudspeth, M. C., Claus, B., Fezzaa, K., Sun, T., Chen, W., Gong, X. L. & Luo, S. N. (2014). Rev. Sci. Instrum. 85, 076101.  Web of Science CrossRef PubMed Google Scholar
First citationLu, L., Huang, J. W., Fan, D., Bie, B. X., Sun, T., Fezzaa, K., Gong, X. L. & Luo, S. N. (2016). Acta Mater. 120, 86–94.  Web of Science CrossRef CAS Google Scholar
First citationLu, L., Sun, T. K. K. F., Fezzaa, K., Gong, X. L. & Luo, S. N. (2017). Mater. Sci. Eng. A, 701, 143–148.  Web of Science CrossRef CAS Google Scholar
First citationLuo, S. N., An, Q., Germann, T. C. & Han, L. B. (2009). J. Appl. Phys. 106, 013502.  Web of Science CrossRef Google Scholar
First citationLuo, S. N., Jensen, B. J., Hooks, D. E., Fezzaa, K., Ramos, K. J., Yeager, J. D., Kwiatkowski, K. & Shimada, T. (2012). Rev. Sci. Instrum. 83, 073903.  Web of Science CrossRef PubMed Google Scholar
First citationMa, L., Wang, L., Nie, Z., Wang, F., Xue, Y., Zhou, J., Cao, T., Wang, Y. & Ren, Y. (2017). Acta Mater. 128, 12–21.  Web of Science CrossRef Google Scholar
First citationMilathianaki, D., Boutet, S., Williams, G. J., Higginbotham, A., Ratner, D., Gleason, A. E., Messerschmidt, M., Seibert, M. M., Swift, D. C., Hering, P., Robinson, J., White, W. E. & Wark, J. S. (2013). Science, 342, 220–223.  Web of Science CrossRef CAS PubMed Google Scholar
First citationPetit, J., Castelnau, O., Bornert, M., Zhang, F. G., Hofmann, F., Korsunsky, A. M., Faurie, D., Le Bourlot, C., Micha, J. S., Robach, O. & Ulrich, O. (2015). J. Synchrotron Rad. 22, 980–994.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationPlimpton, S. (1995). J. Comput. Phys. 117, 1–19.  CrossRef CAS Web of Science Google Scholar
First citationPoulsen, H. F., Wert, J. A., Neuefeind, J., Honkimäki, V. & Daymond, M. (2005). Nat. Mater. 4, 33–36.  Web of Science CrossRef CAS Google Scholar
First citationRavelo, R., Germann, T. C., Guerrero, O., An, Q. & Holian, B. L. (2013). Phys. Rev. B, 88, 134101.  Web of Science CrossRef Google Scholar
First citationRemington, T. P., Ruestes, C. J., Bringa, E. M., Remington, B. A., Lu, C. H., Kad, B. & Meyers, M. A. (2014). Acta Mater. 78, 378–393.  Web of Science CrossRef CAS Google Scholar
First citationSeiboth, F., Fletcher, L. B., McGonegle, D., Anzellini, S., Dresselhaus-Cooper, L. E., Frost, M., Galtier, E., Goede, S., Harmand, M., Lee, H. J., Levitan, A. L., Miyanishi, K., Nagler, B., Nam, I., Ozaki, N., Rödel, M., Schropp, A., Spindloe, C., Sun, P., Wark, J. S., Hastings, J., Glenzer, S. H. & McBride, E. E. (2018). Appl. Phys. Lett. 112, 221907.  Web of Science CrossRef Google Scholar
First citationSingh, A. K. (1993). J. Appl. Phys. 73, 4278–4286.  CrossRef Web of Science Google Scholar
First citationSpearot, D. E., Tschopp, M. A., Jacob, K. I. & Mcdowell, D. L. (2007). Acta Mater. 55, 705–714.  Web of Science CrossRef CAS Google Scholar
First citationStukowski, A. (2010). Modell. Simul. Mater. Sci. Eng. 18, 015012.  Web of Science CrossRef Google Scholar
First citationTang, M. X., Zhang, Y. Y., E, J. C. & Luo, S. N. (2018). J. Synchrotron Rad. 25, 748–756.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationTschopp, M. A. & McDowell, D. L. (2008). J. Mech. Phys. Solids, 56, 1806–1830.  Web of Science CrossRef CAS Google Scholar
First citationTurneaure, S. J., Sharma, S. M., Volz, T. J., Winey, J. M. & Gupta, Y. M. (2017). Sci. Adv. 3, eaao3561.  Web of Science CrossRef PubMed Google Scholar
First citationWang, L., Cai, Y., He, A. M. & Luo, S. N. (2016). Phys. Rev. B, 93, 174106.  Web of Science CrossRef Google Scholar
First citationWang, L., Zhao, F., Zhao, F. P., Cai, Y., An, Q. & Luo, S. N. (2014). J. Appl. Phys. 115, 053528.  Web of Science CrossRef Google Scholar
First citationWehrenberg, C. E., McGonegle, D., Bolme, C., Higginbotham, A., Lazicki, A., Lee, H. J., Nagler, B., Park, H. S., Remington, B. A., Rudd, R. E., Sliwa, M., Suggit, M., Swift, D., Tavella, F., Zepeda-Ruiz, L. & Wark, J. S. (2017). Nature, 550, 496–499.  Web of Science CrossRef CAS PubMed Google Scholar
First citationZhang, M., Li, L., Ding, J., Wu, Q., Wang, Y.-D., Almer, J., Guo, F. & Ren, Y. (2017). Acta Mater. 141, 294–303.  Web of Science CrossRef CAS Google Scholar
First citationZhang, Y. Y., Tang, M. X., Cai, Y., E, J. C. & Luo, S. N. (2019). J. Synchrotron Rad. 26, 413–421.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationZhu, Y., Kang, G. & Yu, C. (2017). Int. J. Plast. 95, 191–215.  Web of Science CrossRef Google Scholar

© International Union of Crystallography. Prior permission is not required to reproduce short quotations, tables and figures from this article, provided the original authors and source are cited. For more information, click here.

Journal logoJOURNAL OF
SYNCHROTRON
RADIATION
ISSN: 1600-5775
Follow J. Synchrotron Rad.
Sign up for e-alerts
Follow J. Synchrotron Rad. on Twitter
Follow us on facebook
Sign up for RSS feeds