Effect of grain boundary structures on the behavior of He defects in Ni: An atomistic study
Gong H F1, 2, 3, 4, †, Yan Y1, Zhang X S1, Lv W4, Liu T1, Ren Q S1
ATF R&D, China Nuclear Power Technology Research Institute Co., Ltd, Shenzhen 518000, China
Shanghai Institute of Applied Physics, Division of Nuclear Materials and Engineering, Chinese Academy of Sciences, Shanghai 201800, China
Key Laboratory of Interfacial Physics and Technology, Chinese Academy of Sciences, Shanghai 201800, China
Department of Mechanical Engineering, University of Michigan, Ann Arbor, Michigan 48109, USA

 

† Corresponding author. E-mail: gonghengfeng@cgnpc.com.cn

Project supported by the Program of International S&T Cooperation, China (Grant No. 2014DFG60230), the National Basic Research Program of China (Grant No. 2010CB934504), Strategically Leading Program of the Chinese Academy of Sciences (Grant No. XDA02040100), the Shanghai Municipal Science and Technology Commission, China (Grant No. 13ZR1448000), the National Natural Science Foundation of China (Grant Nos. 91326105 and 21306220).

Abstract

We investigated the effect of grain boundary structures on the trapping strength of HeN (N is the number of helium atoms) defects in the grain boundaries of nickel. The results suggest that the binding energy of an interstitial helium atom to the grain boundary plane is the strongest among all sites around the plane. The HeN defect is much more stable in nickel bulk than in the grain boundary plane. Besides, the binding energy of an interstitial helium atom to a vacancy is stronger than that to a grain boundary plane. The binding strength between the grain boundary and the HeN defect increases with the defect size. Moreover, the binding strength of the HeN defect to the grain boundary becomes much weaker than that to other grain boundaries as the defect size increases.

1. Introduction

The interaction of helium (He) defects with the grain boundary (GB) is of fundamental importance for the generation-IV molten salt reactor (MSR), where the structural materials of Ni-based alloys are exposed to the harsh reactor environment with high-temperature, strong corrosion, and high flux of He by the (n, α) transmutation reactions. Due to their extremely low solubility, He defects have a strong tendency to be trapped in other defect regions containing an excess volume, such as vacancies, dislocations, and GBs. This process leads to the subsequent formation of He bubbles (or voids) within those defects.[14] The formation of He bubbles in a material can lead to void swelling, high temperature intergranular embrittlement, surface roughening, blistering, etc., which will significantly degrade the mechanical properties of the material and negatively impact the lifetime and safety of a reactor.[58]

Previous experimental studies have attributed He embrittlement to the clustering and migration of He bubbles toward microscale features, such as GBs and dislocations.[911] He bubbles have been proposed to cause, for example, the initiation of GB cracks and premature brittle failure of materials together with a drastic reduction in ductility under static and cyclic loads.[12] Although microstructural changes caused by grain boundary bubble formation can be observed with several experimental techniques such as transmission electron microscopy (TEM) and small-angle neutron scattering (SANS), these techniques cannot directly observe He segregation at GBs without bubble formation. Studies have used TEM and He thermal desorption spectrometry (TDS) to investigate the effect of alloying elements on the mechanism of bubble growth and migration.[13] To test the hypothesis that the surface of nanoclusters is a preferential nucleation site for He bubbles, Edmonson et al.[14] studied a nanostructured ferritic alloy irradiated with He+ ions. The results showed He bubble nucleation on the surface of nanoclusters and Ti (N, C) precipitates along GBs and dislocations. However, how the He bubble interacts with defects, especially at GBs, remains unclear. In addition, advances in experiment technology and theoretical method may motivate new thinking. For instance, the interstitial He diffusion along GBs was considered to be slower than He diffusion in grains,[15] while recent thermal He desorption experiments indicated that GBs may act as easy paths for He diffusion toward surfaces.[16]

To date, there have been several studies using first principles and molecular dynamics (MD) simulations to investigate how He bubbles segregate at GBs in metals.[1721] There have been relatively few studies focusing on He interaction with GBs.[2229] MD simulations have been used to understand the migration of interstitial He inside different GBs[30] and to understand how the GB strength is affected by He or He bubbles.[3133] First principles calculations have been used to understand the diffusion and stability of He defects at the Σ5(310) symmetric tilt GB (STGB).[34] Kurtz and Heinisch applied the Finnis–Sinclair potential in MD simulations, showing that substitutional He is more weakly bound to the GB plane than the interstitial He.[22] Gao et al.[23] investigated the relationship between the maximum binding energy and the GB. Hammond et al.[35] showed that the GB does not facilitate He transport in tungsten and several other metals. Kashinath et al.[36] showed that nanoscale platelets at semi-coherent interfaces can store He bubbles, and these bubbles can remain stable under irradiation. However, there are a number of unresolved issues related to how the He defect interacts with GBs, especially in Ni-based alloys. For example, existing studies using atomistic simulations have not considered a wide range of GB structures to understand the effects of macroscopic variables on the binding properties of the N-th He atom to the remaining HeN−1VM defect (N is the number of He atoms and M is the number of vacancies), especially the behavior of He defects in the GB of Ni, or the interaction between the HeNVM defect and the GB. Establishing a trend of these binding properties for Ni-based alloys will be very useful for mesoscale simulations of He bubble nucleation and growth.

This paper focuses on understanding how the GB of Ni interacts with the interstitial He atom, its dependence on the distance of the He defect from the center of the GB plane, and how the local atomistic environment in GBs affects the trapping strength of the N-th He atom to the remaining defect, as well as the trapping strength of the HeNVM defect to the GB in Ni-based alloys. The outline of the paper is as follows. The computational techniques are presented in Section 2. In Section 3, we calculate the interaction between five STGBs and HeN defects. Section 4 presents the conclusion.

2. Methodology

All calculations were performed using MD simulations with the LAMMPS software package.[37] In the simulation cell, the atomic interactions of Ni–Ni, Ni–He, and HeHe–He atoms were described by the modified analytic embedded atom method (MAEAM),[3847] the Morse potential,[48] and the Lennard–Jones potential,[49] respectively. More information about the formulas and parameters can be found in Ref. [29]. In our MD simulations, the conjugate gradient (CG) algorithm and the Nosé constant-temperature technique[50] with a time step of 0.5 fs were used. Due to different GB configurations, the total simulation time varied between 100.0 ps and 150.0 ps. Firstly, the initial configurations of GB were constructed by the GB studio software package,[51] as shown in Fig. 1. The output of the GB studio calculations served as the input of the MD simulation. However, before being written into MD simulation, the GB needed to go through a relax and quench phase in order to obtain the truly stable GB structure. Secondly, an interstitial He atom was placed initially into the GB, and energy minimization was performed for the simulation cell by iteratively adjusting the atomic coordinates with a conjugate gradient algorithm so that high overlapping atoms could adjust their positions. Thirdly, Ni atoms were equilibrated at 300 K for 25 ps using NVT ensemble for the system to relax to an equilibrium state. Finally, the simulation cell was quenched to 0 K to obtain the stable GB configuration with an interstitial He atom. The same quench process was employed to obtain other stable GB configurations with each additional He atom. When a He atom was added to around the GB plane, it was inserted into a spherical region centered at the GB plane with a radius of 1.0 Å. Starting from the GB plane, the He atom was inserted every 1 Å perpendicular to the GB plane at each time in order to study the relationship between the distance and the binding energy. Besides, HeN defects in the GB planes were formed by inserting He atoms into the sphere of 1.0 Å radius around the GB plane accordingly each time. The same procedures of energy minimization, relax, and quench were carried out as previously to obtain the final stable GB structures containing He defects. In the simulation procedure, He interstitials were frozen, avoiding the He atom diffusion. An interstitial He atom was inserted into the GB to calculate the trapping strength of the GB to the He defect.

Fig. 1. (color online) Atomic blocks showing the initial and final stable STGBs analyzed using the common neighbor analysis (CNA). The blue balls represent the FCC crystal structure. The red balls represent the irregular crystal structures. The green balls represent the HCP crystal structure. (a) Σ3 GB; (b) Σ5 GB; (c) GB; (d) Σ11 GB; (e) Σ13 GB. Panels (a1)–(e1) represent the GB structures before energy minimization and panels (a2)–(e2) represent the GB structures after relaxation and quenching.

In this work, five STGBs are considered and the parameters are listed in Table 1. These GBs are high angle GBs based on the 15° Brandon criterion. The STGBs were observed with a concentration higher than that of random GBs according to the experiments.[52,53] In our calculations, the GB energy ranges from to , as shown in Table 1, which is consistent with other works.[55] For example, Vasily et al.[55] studied the GB energy for FCC metals recently. Their results pointed out that the random GB energy is about in Ni bulk. The experiments showed that the concentration of STGB is multiples of a random distribution (MRD) of GBs.[52,54] In our calculations, the GB has the lowest GB energy (. The initial and final stable STGB structures are both shown in Fig. 1, in which the blue balls represent the FCC crystal structure, the red balls represent the irregular crystal structure, and the green balls represent the HCP crystal structure. The partial sections of the GB planes are inserted in the right of the full GB structures, which present the final stable GB structures after relaxation and quenching. Due to the periodic boundary conditions applied in all directions, the GB plane slightly slides when the energy minimization using the CG algorithm is carried out. However, the final GB structure is still much more stable, which can be used to finish the study of He doping behavior.

Table 1.

Simulation parameters, including Σ (the reciprocal density of coincidence sites of two interpenetrating lattices misoriented by a relative rotation), the grain boundary (GB) plane (normal to the z-direction), 0° GB plane (XY), rotation axis [X], the disorientation angle θ about the corresponding tilt axis (x-direction), the GB energy, the simulation cell size, and the number of atoms.

.

Generally, the GB configurations will significantly influence the thermal stability of the GB, the preferential GB segregation site for defects, and the segregation energy of solutes. The most stable GB configurations can be obtained by the process of energy minimization. In this work, the GB energy, , is calculated by

where is the total energy of the simulation cell that contains GBs, is the total energy of a Ni crystal composed of the same number of atoms, and S is the area of the GB plane in the simulation cell.

The formation energy of the HeN defect within the GB is determined by comparing the total energy of a GB containing the defect to the energy of the GB without the defect. The formation energy of the HeN defect within the GB, , is given by

where is the total energy of the GB that contains the HeN defect after relaxation, and is the total energy of the grain without the defect. The cohesive energy of a perfect He crystal is negative and very small, so it is not included in Eq. (2).

The interaction strength of the HeN defect to a GB is evaluated by the binding energy. Two types are considered here: (I) the binding energy of the N-th He atom to the remaining defect in a GB, e.g., in Eq. (3), (II) the binding energy of the HeN defect to the GB, e.g., in Eq. (4).

In this work, the binding energy of the N-th He atom to the remaining defect in a GB, , is defined as

where is the corresponding formation energy.

The total binding energy of the HeN defect interacting with the GB can be obtained by comparing the formation energy of the HeN defect in Ni bulk to that in the GB. Here, the binding energy of the HeN defect to the GB, , is calculated by

where is the formation energy of the HeN defect in Ni bulk, which is calculated by equation , where is the total energy containing HeN cluster in Ni bulk, is the cohesive energy of a He crystal, and is the cohesive energy of a Ni crystal. The binding energy of the N-th He atom to the remaining defect is obtained by equation , where the formation energy of an interstitial He atom in Ni bulk is , which is very consistent with the results of Mizuno et al.[56] and Meliu et al.,[57] and is the formation energy of the HeN defect in Ni bulk. In Eq. (4), is the formation energy of the HeN defect in Ni bulk, and is the formation energy of the defect in the GB. Generally, a positive binding energy means that it is energetically favorable for the HeN defect to segregate into the GB, while a negative binding energy means that the defect does not prefer to segregate into the GB.

Table 2.

The formations energy and binding energy of HeN clusters in Ni bulk (in units of eV).

.
3. Results and discussion

Helium embrittlement in Ni-based alloys is a common issue, which can be controlled to a significant degree by controlling the diffusion and segregation of He defects in Ni bulk and in GBs. In this work, a set of model GBs are explored to provide insight into the role of structure in the He segregation behavior at GBs, while the real material has a much wider range of GB properties. To complete the binding energy calculation of an interstitial He atom in the GB, we first plot the formation energy of the He defect in various GB sites against its distance from the planes to quantify the change of the formation energy. Figure 2 is an example of such a plot for five STGBs. In this plot, the formation energy (Eq. (2)) minimizes near the GB planes, which is consistent with other works.[28,5860] For example, Tschopp et al.[28] studied the binding energy of interstitial He and di-He defects to the GB structures in α-Fe. They pointed out that in the GB planes, the formation energy of one He atom and two He atoms is the minimum. The formation energy of an interstitial He atom is 2.30 eV in the Σ3 GB plane and is the lowest among the five STGBs, which suggests that it is easier for an interstitial He atom to segregate into the Σ3 GB plane. In Σ5, Σ9, Σ11, and Σ13 GB planes, the formation energy of the He defect is 3.05 eV, 2.28 eV, 3.40 eV, and 3.28 eV, respectively, among which the formation energy is highest for Σ11 GB. By comparing the formation energy of an interstitial He defect in Ni bulk () to that (–3.40 eV) in the GB planes, we find that it is easier for the He defect to segregate into the GB planes. Our results can provide important insights to explain why He bubbles tend to segregate into the GB planes. The characteristic length of He defect segregation is studied only within 8.0 Å. In the far away region from the GB plane, the formation energy will be nearly 4.06 eV, which is equal to that in the Ni bulk environment. The main GB-affected region is shaded (light yellow) in Fig. 2. Except for Σ9, the distribution of the formation energy of the helium defect in GBs demonstrates certain symmetry around the GB planes, highlighting the symmetric nature of these GB configurations. Actually, after the relaxation and quenching process, the central position of the Σ9 GB plane exhibits a large shift relative to the initial position, which can be used to explain why the formation energy of an interstitial He atom around the Σ9 GB plane does not show the symmetry as that around other GB planes.

Fig. 2. (color online) The formation energy of an interstitial helium defect in the GB as a function of the initial distance of the He atom to the GB plane. The shaded area indicates the main interaction region.

Based on the formation energy calculations, the binding energy of an interstitial helium defect to various GBs as a function of its distance to the GB planes is explored. From Fig. 3, we can see that the binding energy reaches a maximum when an interstitial He defect locates at the GB planes. Some binding energy turns out to be negative, suggesting that the He defect does not like to stay at those positions. The Σ3 GB having a much lower GB energy possesses the highest binding energy of an interstitial He defect (Max ). For all GB structures, there is a largest binding energy in the GB plane, as shown in Fig. 3. Our results are consistent with other works.[28,54] For example, Tschopp et al.[54] studied the binding energy of HeNV1 defects to α-Fe GBs. Their results showed that the “twin” GB has significantly lower binding energy for all HeNV1 defects than other GBs, which are very consistent with our calculation.[54] Also notice that the Σ3 GB has the lowest GB energy, which explains its difference in the binding strength. For example, the study of He defects in the GB of α-Fe showed that the formation energy (and consequently, binding energy) for each particular GB is related to the GB energy.[28,54]

Fig. 3. (color online) The binding energy of an interstitial He atom to the GB plane as a function of the initial distance of the He atom to the GB plane. The shaded area indicates the main interaction region.

To understand the effect of GB characteristics on HeN defect absorption, we investigate the binding energy of an interstitial He defect to the GB plane against the disorientation angle and the GB energy, respectively. Figure 4 plots the relationship between the binding energy of an interstitial He defect to the GB plane and the GB energy in the [001], [110], and tilt orientations. The GB energy ranges from to . The GB with the lowest GB energy gives the highest binding energy. This suggests that the HeN defect is easier to segregate into the GB having lower GB energy. For all other GBs, the binding energy of an interstitial He defect to the GB planes has slight difference as a function of the GB energy. For example, Tschopp et al.[60] pointed out that most GBs within the same tilt system ([001], [110], and have very similar binding energy aside from a few “special” boundaries (e.g., the , , and STGBs) that have binding energy higher than that of the rest. Besides, for the [110] tilt orientation, as the GB energy increases, the binding energy decreases. However, for the tilt orientation, the binging energy increases with the GB energy increasing.

Fig. 4. (color online) The binding energy of an interstitial He defect to the GB plane for various GB configurations as a function of the GB energy.

We investigate the interaction between the HeN defect and the GB plane by adding interstitial He atoms into the GB plane. Here, we first select ten random numbers to generate the initial configurations of the HeN defect, and then obtain the stable configurations through the relaxation and quenching of the simulation system. From these stable HeN defect configurations, the most stable state is chosen to analyze the binding energy. It should be noted that as the number of He atoms increases, the lattice Ni atoms will be kicked out, forming the self-interstitial Ni atoms. However, we only consider in this study the trapping strength of the HeN defect in Ni bulk and at the GB: the self-interstitial generated Ni atom is ignored here. After the relaxation of the whole simulation system, the HeN defect is kept at the position of the GB plane. The binding energy of an interstitial He defect to the remaining defect in the GB plane is calculated using Eq. (3). The dependence of the binding energy on the defect size is depicted in Fig. 5. With increasing the number of He atoms in the HeN defect in the GB plane, the binding energy tends to increase. This suggests that the supplemental He atom bonds to the remaining defect and it tends to grow to decrease the total energy of the simulated system, which is very consistent with the experimental results.[61] The binding energy of the He defect to the remaining defect is positive in the GB planes, except for Σ5 and Σ11 GBs with negative values when the number of He atoms is 6. The negative binding energy indicates that the He atoms occupy the plane positions but do not bind to each other. By comparing the binding energy of an interstitial He defect to the remaining defect in Ni bulk and that in the GB plane, we find that the majority of the binding energy is generally higher in Ni bulk than that in the GB plane, which tells us that the HeN defect is much more stable in Ni bulk than in the GB plane environment. Besides, there is some maximum value as the number of He defects increases. For example, the maximum binding energy of an interstitial He defect to the remaining defects is 2.34 eV for the Σ11 GB when the number of He atoms is seven. For the Σ3, Σ5, Σ9, and Σ13 GBs, the maximum binding energy of an interstitial He defect to the remaining defects is 1.52 eV, 1.74 eV, 1.86 eV, and 0.89 eV, respectively, which suggests that at this point, the HeN defect is the most stable. The majority of the binding energy does not exceed 2.0 eV, which is consistent with other work.[18] These results suggest that the local GB environment has a significant impact on the stable structures of the HeN defect as well as on their trapping strength in the GB plane.

Fig. 5. (color online) The binding energy of the N-th He defect to the remaining defect as a function of the number of He atoms in the defect.

On the basis of the discussion about the binding energy of an interstitial He to the remaining He defects, we further investigate the binding strength of the HeN defect to the GB plane. All calculations are performed with Eq. (4). The binding energy of the HeN defect to the GB plane is plotted as a function of the number of He atoms. In this plot, all the binding energies are positive, and the binding energy increases with the HeN defect size, as shown in Fig. 6. Additionally, the binding energy of an interstitial He defect to the GB plane is 1.25 eV, 1.90 eV, 1.28 eV, 1.99 eV, and 1.98 eV, respectively, for the Σ3, Σ5, Σ9, Σ11, and Σ13 GBs. By comparing the binding energy of an interstitial He defect to a vacancy () in Ni bulk with that to the GB plane (–1.99 eV), we find that the trapping strength of the GB plane to an interstitial He defect is weaker than that of the vacancy in Ni bulk, which is consistent with the results of binding strength of He and di-He defects to the GB plane in α-Fe.[34] These results can provide insights in understanding the He embrittlement or He damage in Ni-based alloys for future Generation-IV MSR.

Fig. 6. (color online) The binding energy of the HeN defect to the GB plane as a function of the number of He atoms in the defect.
4. Conclusion

The interaction between the HeN defect and the GBs of Ni bulk was studied using MD simulations. Five high angle STGBs with [100], [110], and tilt orientations were investigated. We calculated the formation energy and the binding energy of an interstitial He defect in all potential GB sites within a length scale of 8.0 Å. The results show that the binding energy of an interstitial He defect to the GB plane is the strongest among all sites around the plane. The HeN defect is much more stable in Ni bulk than in the GB plane. In addition, the binding energy of an interstitial He defect to a vacancy is stronger than that to a GB plane. The binding strength between the GB and the HeN defect increases with the HeN defect size, and the binding strength of the HeN defect to the GB is the weakest. This study is helpful to understand how a GB interacts with the HeN defect in a FCC crystal, how the GB configuration affects the He defect trapping mechanism, and ultimately how they affect the He (re-)combination and embrittlement in the GB for Ni-based alloys.

Reference
[1] Zinkle S J 2005 Phys. Plasmas 12 058101
[2] Ishiyama Y Kodama M Yokota N Asano K Kato T Fukuya K 1996 J. Nucl. Mater. 239 90
[3] Stoller R E Odette G R 1988 J. Nucl. Mater. 155 1328
[4] Lewis M B Farrell K 1986 Nucl. Instrum. Methods Phys. Res. B 16 163
[5] Bloom E E Busby J T Duty C E Maziasz P J McGreevy T E Nelson B E Pint B A Tortorelli P F Zinkle S J 2007 J. Nucl. Mater. 367 1
[6] Zinkle S J Busby J T 2009 Mater. Today 12 12
[7] Yamamoto T Odette G R Kishimoto H 2006 J. Nucl. Mater 356 27
[8] Trinkaus H Singh B N 2003 J. Nucl. Mater 323 229
[9] Johnson P B Mazey D J 1978 Nature 276 595
[10] Trinkaus H Singh B N 2003 J. Nucl. Mater 318 234
[11] Zinkle S J Busby J T 2009 J. Nucl. Mater. 386 8
[12] Trinkaus H 1983 Radiation Effects 78 189
[13] Kalashnikov A N Chernov I I Kalin B A Binyukova S Y 2002 J. Nucl. Mater. 307 362
[14] Edmondson P D Parish C M Zhang Y Hallén A 2011 Scripta Mater. 65 731
[15] Lane P L Goodhew P J 1983 Philos. Mag. 48 965
[16] Lefaix-jeuland H Moll S Jourdan T Legendre F 2013 J. Nucl. Mater. 434 152
[17] Tschopp M A McDowell D L 2007 Philos. Mag. 87 3147
[18] Baskes M I Vitek V 1985 Metall. Trans. 16 1625
[19] Yamaguchi M Nishiyama Y Kaburaki H 2007 Phys. Rev. B. 76 0355418
[20] Wachowicz E Kiejna A 2011 Modell. Simul. Mater. Sci. Eng. 9 025001
[21] Rhodes N R Tschopp M A Solanki K N 2013 Modell. Simul. Mater. Sci. Eng. 21 035009
[22] Kurtz R J Heinisch H L 2004 J. Nucl. Mater. 329 1199
[23] Gao F Heinisch H L Kurtz R J 2006 J. Nucl. Mater. 351 133
[24] Kurtz R J Heinisch H L Gao F 2008 J. Nucl. Mater. 382 134
[25] Gao F Heinisch H L Kurtz R J 2009 J. Nucl. Mater. 386 390
[26] Zhang L Shu X L Jin S Zhang Y Lu G H 2010 J. Phys.: Condens. Matter 22 375401
[27] Zhang L Fu C C Lu G H 2013 Phys. Rev. 87 134107
[28] Tschopp M A Gao F Yang L Solanki K N 2014 J. Appl. Phys. 115 1
[29] Xia J X Hu W Y Yang J Y Ao B Y 2006 Phys. Stat. Soli. B 243 1
[30] Terentyev D He X 2010 Comput. Mater. Sci. 49 858
[31] Hafez H S Z Lucas G Schäublin R 2009 Europhys. Lett. 85 6008
[32] Demkowicz M J Bhattacharyya D Usov I Wang Y Q Nastasi M Misra A 2010 Appl. Phys. Lett. 97 161903
[33] Zhang Y F Millett P C Tonks M Zhang L Z 2012 J. Phys.: Condens. Matter 24 305005
[34] Zhang L Zhang Y Lu G H 2013 J. Phys.: Condens. Matter 25 095001
[35] Hammond K D Hu L Maroudas D Wirth B D 2015 Europhys. Lett. 110 52002
[36] Kashinath A Misra A Demkowicz M J 2013 Phys. Rev. Lett. 110 086101
[37] http://lammps.sandia.gov/
[38] Daw M S Baskes M I 1984 Phys. Rev. 29 6443
[39] Baskes M I 1992 Phys. Rev. 46 2727
[40] Ouyang Y Zhang B Liao S Jin Z 1996 Phys. B 101 161
[41] Deng H Hu W Shu X Zhang B 2003 Surf. Sci. 543 97
[42] Yang J Hu W Deng H Zhao D 2004 Surf. Sci. 572 2074
[43] Hu W Zhang B Huang B Gao F Bacon D J 2001 J. Phys.: Conden. Matter 13 1193
[44] Hu W Deng H Yuan X Fukumoto M 2003 Euro. Phys. J. 34 429
[45] Hu W Shu X Zhang B 2002 Comput. Mater. Sci. 23 175
[46] Hu W Fukumoto M 2002 Modell. Simula. Mater. Sci. 10 707
[47] Johnson R A 1990 Phys. Rev. 41 9717
[48] Baskes M I Melius C F 1979 Phys. Rev. 20 3197
[49] Johnson R A 1973 J. Phys. F: Metal Phys. 3 295
[50] Nosé S 1991 Prog. Theor. Phys. Suppl. 103 1
[51] https://staff.aist.go.jp/h.ogawa/GBstudio/indexE.html
[52] Beladi H Rohrer G S 2013 Acta Mater. 61 1404
[53] Beladi H Tphrer G S 2013 Metall. Mater. Trans. 44 115
[54] Tschopp M A Gao F Solanki K N 2014 J. Appl. Phys. 115 1
[55] Bulatov V V Reed B W Kumar M 2014 Acta Mater. 65 161
[56] Mizuno T Asato M Hoshino T Kawakami K 2001 J. Magn. Magn. Mater. 226 386
[57] Baskes M I Melius C F 1981 Phys. Rev. 20 3197
[58] Demkowicz M J Anderoglu O Zhang X Misra A 2011 J. Mater. Res. 26 1666
[59] Bai X M Vernon L J Hoagland R G Voter A F Nastasi M Uberuaga B P 2012 Phys. Rev. 85 214103
[60] Tschopp M A Solanki K N Gao F Sun X Khaleel M A Horstemeyer M F 2012 Phy. Rev. 85 064108
[61] Ryazanov A Voskoboinikov R E Trinkaus H 1996 J. Nucl. Mater. 1085 233