Microscopic origin of MXenes derived from layered MAX phases

Zhonglu Guoab, Linggang Zhua, Jian Zhoua and Zhimei Sun*a
aSchool of Materials Science and Engineering, and Center for Integrated Computational Materials Engineering, International Research Institute for Multidisciplinary Science, Beihang University, Beijing 100191, China. E-mail: zmsun@buaa.edu.cn
bCollege of Materials, Xiamen University, Xiamen 361005, China

Received 31st December 2014 , Accepted 4th March 2015

First published on 4th March 2015


Abstract

Two-dimensional transition metal carbides/nitrides Mn+1Xns labeled as MXenes derived from layered transition metal carbides/nitrides referred to as MAX phases attract increasing interest due to their promising applications as Li-ion battery anodes, hybrid electro-chemical capacitors and electronic devices. To predict the possibility of forming various MXenes, it is necessary to have a full understanding of the chemical bonding and mechanical properties of MAX phases. In this work, we investigated the chemical bonding changes of MAX phases in response to tensile and shear stresses by ab initio calculations using M2AlC (M = Ti, Zr, Hf, V, Nb, Ta, Cr, Mo and W) as examples. Our results show that the M2C layer is likely to separate from the Al layer during the tensile deformation, where the failure of M2AlC is characterized by an abrupt stretch of the M–Al bonds. While under shear deformation, the M2C and Al layers slip significantly relative to each other on the (0001) basal planes. It is found that the ideal strengths of M2AlC are determined by the weak coupling of the M2C and Al layers, closely related to the valence-electron concentration. Our results unravel the possibility as well as the microscopic mechanism of the fabrication of MXenes through mechanical exfoliation from MAX phases.


1. Introduction

Recently, a new family of two-dimensional (2-D) transition metal carbides/nitrides labeled as MXenes has attracted growing interest due to their great potential applications as Li-ion battery (LIB) anodes, hybrid electro-chemical capacitors and electronic devices.1–9 More information about MXenes is referred to in review articles published very recently.10,11 The MXenes are produced by selective chemical etching of the “A” metal from MAX phases, which are a family of layered hexagonal transition metal carbides/nitrides. On the other hand, an attempt to mechanically exfoliate MXenes from MAX phases has also been tried, which however, led to multilayer structures.12 The MAX phases having a formula of Mn+1AXn, where M is an early transition metal, A is a group IIIA or IVA element, X is either C and/or N, and n = 1–3, have the combined properties of both metals and ceramics,13,14 such as unusual high stiffness, plastic deformability, good electrical and thermal conductivity. So far, more than 70 MAX phases are known to exist, among which there are roughly 50 M2AX phases (space group P63/mmc, prototype Cr2AlC).13 The structure of M2AX can be described as M2X layers interleaving with A layers.15–19 Therefore, it is possible to form many types of MXenes derived from the M2AX phases through an elegant exfoliation approach. However, up to now, only four M2C-stoichiometry MXenes (Ti2C, Nb2C, V2C and (Ti0.5Nb0.5)2C) have been synthesized.10 Furthermore, the microscopic origin of exfoliating MXenes from the MAX phases is not clear, which is important for predicting the possibility of forming various MXenes. In this work, to provide an insight into the microscopic mechanism of forming MXenes, we have extensively investigated the mechanical property and the chemical bonding changes of MAX phases in response to tensile and shear strains by ab initio calculations using M2AlC (M = Ti, Zr, Hf, V, Nb, Ta, Cr, Mo and W) as examples.

Many theoretical works have been done focusing on the mechanical properties of M2AlC compounds in the past decades.15,16,20–29 In our previous studies, we have investigated the elastic properties of nanolayered M2AlC/M2AlN,15,16,25,26 with M = Ti, Zr, Hf, V, Nb, Ta, Cr, Mo and W, by ab initio calculations. We have also classified the M2AlC phases into two groups by whether the bulk modulus of binary MC is conserved or changed after the incorporation of Al,16 which can be understood in terms of the coupling between the MC and Al layers. On the other hand, the M2AlC phases (M = Sc, Y, La, Ti, Zr, Hf, V, Nb, Ta, Cr, Mo, W) could also be classified into two groups based on the valence-electron concentration induced changes in C44.22 On the other hand, the elastic properties of MXenes have also been calculated recently.30 However, these works are mainly focused on elastic properties, which are the secondary derivative of the total energy with respect to a small strain. To gain a deeper understanding of the micro mechanism for the exfoliation of M2C from M2AlC that is achieved by the cleavage of M2C and Al layers, it is essential to investigate the mechanical properties under deformation, i.e. complete stress–strain relations, from which the first maximum stress is ideal strength.31–33 Ideal strength is the first derivative of total energy with respect to the critical point where instability occurs under deformation, which can be considered as the lowest stress required for cleavage or fracture of perfect crystal.34,35 Furthermore, the macroscopic mechanical properties can be interpreted by studying the atomic displacement and the chemical bonding changes at the microscopic scale when a large strain is applied. In this paper, we have calculated the stress–strain curves of M2AlC with M = Ti, Zr, Hf, V, Nb, Ta, Cr, Mo and W at a large strain and investigated the deformation mechanism by ab initio calculations. We have also discussed the relationship between the valence-electron concentration of M and the ideal strength of M2AlC.

2. Theoretical methods

Our calculations are based on Density Functional Theory (DFT) in conjunction with projector augmented wave (PAW) potentials, as implemented in the Vienna ab initio simulation package.36 For the exchange-correlation functional the generalized gradient approximations (GGA)37 of Perdew–Burke–Ernzerhof (PBE)38 was employed. The PAW potentials with the semi-core electrons treated as valence states were used: 3p3d4s for Ti, V and Cr; 4p4d5s for Zr, Nb and Mo; 5p5d6s for Hf, Ta and W; 3s3p for Al; 2s2p for C. The coulomb effect (U) for “d” electrons of the transition metals has been tested to be negligible, and hence it is not considered in the present work. The tetrahedron method with Blöchl corrections39 was used for cohesive energy calculations. The convergence with respect to self-consistent iterations was achieved when the total energy difference between cycles was less than 10−5 eV. The k-points of 9 × 9 × 9 and 11 × 11 × 11 automatically generated with Monkhorst–Pack scheme40 were used for structure optimization and static self-consistent calculations for the M2AlC unit cell, respectively. While for the 3 × 3 × 1 supercell with 72 atoms, the k-point mesh of 2 × 2 × 2 and 4 × 4 × 4 were employed for structure relaxation and static self-consistent calculation, respectively. The cut-off energy was set as 500 eV. The chosen of these computational parameters are based on the convergence tests for the total energy of the systems. The electron localization function (ELF)41 was analyzed by using the VESTA code.42

For the layered hexagonal structure of M2AlC phases, the [0001] direction has the lowest strength,13,16 therefore tension along the [0001] direction of the unit cell was used to investigate the interlayer interactions and the overall strength of the compounds. For layered hexagonal ternary compounds, the activated slip systems were quite limited on the (0001) basal-plane. Furthermore, perfect dislocations on the (0001) basal planes were detected by transmission electron microscopy after a deformation at room temperature.43,44 To understand the micro-scale shear deformation mechanism of M2AlC, shearing along the (0001)[1[2 with combining macron]10] direction was studied. Supercell with 72 atoms was employed in the shear deformation, which has been tested large enough to avoid the periodic limit. To ensure that the material is under uniaxial tension or pure shear, the lattice vectors and internal atomic positions were simultaneously relaxed using a conjugate-gradient scheme until the Hellmann–Feynman stress tensor components orthogonal to the pre-set strain are less than 0.2 GPa.33,45 Then, the fully relaxed structures under deformation were used to calculate the variations of bonds length, lattice parameter, cohesive energy, DOS and ELF versus strains in the present work.

Finally, it is worth to address the van der Waals effect on the interlayers interactions during tension along the [0001] direction. For this purpose, we calculated the stress–strain curves for Ti2AlC and Nb2AlC by the DFT-D2 method which includes the van der Waals corrections.46 The results show that the van der Waals effect only slightly increases the absolute value of the ideal tensile stress for both Ti2AlC and Nb2AlC, but it does not change the relative trend of the ideal strengths and hence should not have any influence on the conclusion. Therefore, we do not include the van der Waals correction in the present results.

3. Results and discussion

To investigate interlayer interactions of the M2AlC compounds, a tensile strain along the [0001] direction and a shear deformation along the (0001)[1[2 with combining macron]10] direction were performed as schematically illustrated in Fig. 1.
image file: c4ra17304j-f1.tif
Fig. 1 The schematic view of M2AlC under (a) the [0001] tension strain, (b) no-strain and (c) the (0001)[1[2 with combining macron]10] shear strain.

Fig. 2a shows the stress–strain relations for M2AlC under uniaxial tension along the [0001] direction. For all the considered compounds, the stress increases monotonically with increasing the applied strain until reaching a critical point, after which the stress starts to fall. The critical point at the strain–stress curve of M2AlC corresponds to a strain of larger than 20%, indicating good resistance to the tensile deformation, which is in good agreement with the experimental results.13 Since all these nine M2AlC compounds considered here show similar stress–strain curves, we use Nb2AlC as the model system to investigate the deformation mechanism under tensile strains. Fig. 2b displays the lattice parameters (c and a) and bond lengths versus the applied strain for Nb2AlC. As the tensile strain was applied along the [0001] direction, c increases monotonically with strain, while the lattice parameter a decreases slightly. On the other hand, from the initial state to the critical tensile strain, the change in the Nb–Al length with strain is relatively slow and the inter-layer Nb–Al bonds are stretched without breaking, resulting in a monotonic increase of the stress. While after the critical tensile strain, the slope of the curve is larger compared to that before the critical point, indicating that the change in the Nb–Al length with strain is faster and the Nb–Al bonds are finally stretched to be unstable, resulting in a monotonic decrease of the tensile stress. Finally, it is obviously seen that the uniaxial tension along the [0001] direction predominantly elongates the inter-layer Nb–Al bonds, while the stronger Nb–C bonds remain almost unchanged. Fig. 2b reveals variation of the cohesive energy of the relaxed unit cell under various tensile strains. The monotonically increase in cohesive energy with the applied strain is consistent with the gradual stretching of the Nb–Al bonds. It also indicates that extra energy is required to elongate or compress a chemical bond from its initial stable state.


image file: c4ra17304j-f2.tif
Fig. 2 (a) The calculated stress–strain curves of M2AlC, (b) the variations of lattice parameter (a, c) and the bond lengths (Nb–Al, Nb–C) of Nb2AlC, (c) the variations of cohesive energy of Nb2AlC under tensile strain along the [0001] direction. The lattice parameter and bond lengths are normalized by the equation of Vstrain/V0, where V0 and Vstrain represent the values in structures under zero and various strains, respectively. The positions of the labeled atoms (Nb2) can be found in Fig. 1.

Further analysis on the electron localization functions (ELF)41 visualizes the chemical bonding changes as a series of tensile strains were applied. The topological analysis of ELF is a very useful tool for the determination of chemical bonding strength. To gain a directly observation of the bond variation and atom distortion during the tensile deformation, Fig. 3 displays the ELF contour plots projected on the (11[2 with combining macron]0) planes for Nb2AlC at various tensile strains (0, 10%, 20%, 28%, 35% and 49%). It is clearly seen in Fig. 3 that with increasing the engineering tensile strain, the Nb–C bond strength remains unchanged, while the Nb–Al bonds are weakened significantly, resulting in the electrons being localized at around Al atoms. All the other eight M2AlC compounds display similar characters during applying the tensile strain. Given the analysis above, the deformation mechanism of M2AlC under the uniaxial tensile strain along the [0001] direction can be concluded as the stretch and breaking of M–Al bonds, and finally the compounds are separated into Al and M2C layers. Meanwhile, though the anisotropy of chemical bonding of MAX phase has been commonly accepted, the present results directly verify the assertion and simulate the bonds variations under deformation with detailed calculations. Therefore, the ideal tensile strengths of M2AlC are determined by the coupling strength of M2C and Al layers. In other words, two dimensional M2Cs could be synthesized through mechanical exfoliation, if the external applied tensile stress could exceed the ideal tensile strength of M2AlC.


image file: c4ra17304j-f3.tif
Fig. 3 The ELF contour plots projected on the (11[2 with combining macron]0) planes of Nb2AlC under the [0001] tensile strains of 0, 10%, 20%, 28%, 35% and 49%.

Fig. 4 displays the total and projected densities of states (DOS) for the strain-free Nb2AlC and that under tensile strains of 10%, 28% and 49%. As seen in Fig. 4, for unstrained Nb2AlC, the Al 3p and Nb 4d states, as well as the C 2p and Nb 4d states are hybridized at the energies of ∼−1.5 eV and ∼−4 eV, respectively. With increasing the tensile strain, it is found that the Al 3p and Nb 4d states are significantly shifted towards the Fermi level and the quantitative number of total DOS at Fermi level correspondingly increases in Table 1, indicating the instability of the stretched Nb–Al bonds. Meanwhile, the hybridization of the Al 3p and Nb 4d states are much weaker compared with the initial states, suggesting that the Nb–Al bonds are softened and can finally be broken by tensile strain. However, the Nb–C bonds do not show any reduction, thus the Nb–C bonds remain stable under all the investigated tensile strains. These conclusions on bonding picture are in good agreement with the above analysis on the basis of ELF results.


image file: c4ra17304j-f4.tif
Fig. 4 The calculated total and partial density of states for Nb2AlC under the [0001] tensile strains of (a) 0, (b) 10%, (c) 28% and (d) 49%.
Table 1 The quantitative numbers of total density of states at the Fermi level for Nb2AlC under the [0001] tensile strains of 0, 10%, 28% and 49%
Tensile strain 0 10% 28% 49%
Numbers of TDOS at EF 3.64 4.41 4.78 5.68


In case of shear deformation of M2AlC, the stress–strain relations under pure shear strain along the (0001)[1[2 with combining macron]10] direction is illustrated in Fig. 5a. For all the 9 investigated compounds, the stresses reach the peak position of the curve at a shear strain of around 12%, and then decrease to zero when the strain is about 23%. Afterwards, the stresses go through a negative minimum and reach to about zero again at a strain of around 49%. To gain a full understanding of these interesting stress–strain curves of M2AlC along the (0001)[1[2 with combining macron]10] direction, we have plotted the bond length change of Nb2AlC under various shear strains in Fig. 5b. It is seen that when Nb2AlC is under shear deformation, Nb2–Al bond is stretched and softened gradually, while Nb1–Al is being compressed up to the critical strain of 12%, where the stress reaches the maximum. Above the critical strain, Nb1–Al bond is recovered to the initial bond length and Nb2–Al bond is completely broken at a strain of 23%, which leads to a large stress decrease to be about zero. The positive shear stresses during this shear stage (from 0 to 23%) are coincided with the increase of cohesive energy versus strains in Fig. 5c. With further increasing the shear strain (from 23% to 49%), the distance between Nb2 and Al atoms increases monotonously and the stresses undergo a negative minimum and reach to another zero again. Meanwhile, the cohesive energy decreases significantly, which could be used to interpret the negative values of stress, indicating that this shear deformation stage (from 23% to 49%) is energetically favorable. Finally, it is interesting to see in Fig. 5b that the Nb–C bond lengths remain nearly unchanged during the whole shear process, indicating that the Nb2C is stable under the investigated shear deformation.


image file: c4ra17304j-f5.tif
Fig. 5 (a) The shear stress–strain curves of M2AlC, (b) the variations of bond length (Nb1–Al, Nb2–Al and Nb–C) of Nb2AlC, (c) the variations of cohesive energy of Nb2AlC under shear strain along the (0001)[1[2 with combining macron]10] direction. The bond lengths are normalized by the equation of Vstrain/V0, where V0 and Vstrain represent the values in structures under zero and various strains, respectively.

To unravel the mechanism of shear deformation of M2AlC, the variation of crystal structure is also analyzed for the case of Nb2AlC. Fig. 6 illustrates the side view of crystal structure for Nb2AlC under the shear strains of 0, 23% and 49% to investigate the atom displacement during the shear deformation. For Nb2AlC at the zero-strain state (Fig. 6a), the labeled Al atom locates on the top left of Nb3 atom, where the Nb2–Al bond has the normal bond length. At a shear strain of 23%, the labeled Al atom slips to the position right above Nb3 atom (Fig. 6b), resulting in the distance between Nb2 and Al atoms increases significantly. Since Nb–C bonds remain stable during the shear deformation, this displacement mostly origins from the slipping between Al and the Nb2C layers. Meanwhile, due to the highest cohesive energy, the structure under the shear strain of 23% is unstable and could be considered as the middle state during the slip displacement of Al atoms. Hence with further increasing the shear strain to be around 49%, the labeled Al atom slips to the top right of Nb3 atom in Fig. 6c and the structure reaches to a stable state, corresponding to a accomplishment of a one-atom-step relatively displacement between Al and Nb2C layer. The present simulated slipping between Al and Nb2C layers is consistent with the experimentally observed perfect dislocations on the (0001) basal planes by transmission electron microscopy after deformation at room-temperature.43,44 In a word, the shear deformation process of M2AlC can be described as the relative slipping between the Al and M2C layers on the (0001) basal planes.


image file: c4ra17304j-f6.tif
Fig. 6 The side view of crystal structures for Nb2AlC under the (0001)[1[2 with combining macron]10] shear strains of (a) 0, (b) 23% and (c) 49%, where the dashed lines surrounded Nb2C layer are used for guiding the eyes.

The first maximum in a tensile (shear) stress–strain curve is defined as the ideal tensile (shear) strength. As seen in Fig. 7, for the 9 individual compounds, a general rule can be concluded that the ideal tensile strength increases with the valence electron concentration, i.e., elements M in group IVB leads to the lowest strength while those in group VIB are the strongest when they form the compound M2AlC. However, the variance trend of the ideal shear strength is relatively complex, and the strength of the M2AlC containing elements in group IVB is quite small considering the downward curve from group VB to group VIB. Apparently, both the tensile and shear strength of the compound is very sensitive to the valance electron concentration.


image file: c4ra17304j-f7.tif
Fig. 7 The calculated ideal strengths and the ratio of tensile to shear strength for M2AlC (M = Ti, Zr, Hf, V, Nb, Ta, Cr, Mo and W).

To all the considered M2AlC phases, their ideal tensile strengths are much larger than the ideal shear strength as can be seen in Fig. 7. Meanwhile, the corresponding critical points in tensile deformation are all much larger compared to those in shear deformation. The crack formation needs a corresponding local tensile stress, which is in the direction perpendicular to the cleavage plane and is larger than the ideal tensile strength. Therefore, for M2AlC, the dislocation nucleation after loading a force larger than the ideal shear stress will be activated before crack formation which needs a force larger than the ideal tensile stress. This means that 2-D M2Cs are much easier to obtain through shear rather than tensile deformation. According to the black solid line in Fig. 7, it is expected that producing 2-D M2Cs is much more easily when M is an element from group IVB and VIB than that from group VB. Finally, among the nine investigated compounds, it can be predicted that two dimensional W2C and Zr2C are the relative easy to obtain because of their low shear strength and high ratio of tensile/shear strength.

4. Conclusions

In summary, we have systematically investigated the tensile and shear properties of layered M2AlC with M = Ti, Zr, Hf, V, Nb, Ta, Cr, Mo and W using the ab initio calculations in order to investigate the possibility of fabricating MXenes via mechanical exfoliation. Using Nb2AlC as an example, we unraveled the deformation mechanism of M2AlC. During both the tensile and shear deformation, the M–C bonds remain stable, while the bond strength of M–Al varies with the engineering strain, which could be used to understand the experimental exfoliation of 2-D M2Cs from M2AlC. In the case of tensile deformation, all the M–Al bonds will be stretched gradually, resulting in the stress increasing to the ideal tensile strength. With further increasing the tensile strain, all the M–Al bonds are broken and the stress is relaxed, leading to M2AlC separated into M2C and Al layers. While for the shear deformation, half of the M–Al bonds are compressed while the other half ones are stretched up to be broken, leading to the slip displacement of Al layer with respect to M2C layers in the (0001) basal planes. Finally, we have also shown that the variation of ideal strength of the M2AlC compounds can be understood in terms of the valence-electron concentrations. It is worth to mention that the present conclusions based on the ground state calculations at 0 K are applicable to describe the nature of exfoliating 2-D M2C from M2AlC phases at room temperature or above before the phase separation of the MAX phases. More importantly, the quantitative and systematic calculations in the present study demonstrate the possibility and microscopic process of mechanically exfoliating 2-D M2C from M2AlC phases, and hence benefit to the synthesis of novel MXenes.

Acknowledgements

This work is supported by National Natural Science Foundation for Distinguished Young Scientists of China (51225205) and the National Natural Science Foundation of China (61274005).

References

  1. M. R. Lukatskaya, O. Mashtalir, C. E. Ren, Y. Dall'Agnese, P. Rozier, P. L. Taberna, M. Naguib, P. Simon, M. W. Barsoum and Y. Gogotsi, Science, 2013, 341, 1502–1505 CrossRef CAS PubMed.
  2. Q. Tang, Z. Zhou and P. Shen, J. Am. Chem. Soc., 2012, 134, 16909–16916 CrossRef CAS PubMed.
  3. X. Zhang, M. Xue, X. Yang, Z. Wang, G. Luo, Z. Huang, X. Sui and C. Li, RSC Adv., 2015, 5, 2762–2767 RSC.
  4. M. Naguib, O. Mashtalir, J. Carle, V. Presser, J. Lu, L. Hultman, Y. Gogotsi and M. W. Barsoum, Two-Dimensional Transition Metal Carbides, ACS Nano, 2012, 6, 1322–1331 CrossRef CAS PubMed.
  5. Y. Xie, M. Naguib, V. N. Mochalin, M. W. Barsoum, Y. Gogotsi, X. Yu, K. Nam, X. Yang, A. I. Kolesnikov and P. R. C. Kent, J. Am. Chem. Soc., 2014, 136, 6385–6394 CrossRef CAS PubMed.
  6. L. Yate, L. E. Coy, G. Wang, M. Beltrán, E. Diaz-Barriga, E. M. Saucedo, M. A. Ceniceros, K. Załeski, I. Llarena, M. Möller and R. F. Ziolo, RSC Adv., 2014, 4, 61355–61362 RSC.
  7. X. Zhang, Z. Ma, X. Zhao, Q. Tang and Z. Zhou, J. Mater. Chem. A 2015, 3, 4960–4966 Search PubMed.
  8. Y. Lee, Y. Hwang, S. B. Cho and Y. Chung, Phys. Chem. Chem. Phys., 2014, 16, 26273–26278 RSC.
  9. Z. Ma, Z. Hu, X. Zhao, Q. Tang, D. Wu, Z. Zhou and L. Zhang, J. Phys. Chem. C, 2014, 118, 5593–5599 CAS.
  10. M. Naguib, V. N. Mochalin, M. W. Barsoum and Y. Gogotsi, Adv. Mater., 2014, 26, 992–1005 CrossRef CAS PubMed.
  11. Q. Tang and Z. Zhou, Prog. Mater. Sci., 2013, 58, 1244–1315 CrossRef CAS PubMed.
  12. X. Zhang, J. Xu, H. Wang, J. Zhang, H. Yan, B. Pan, J. Zhou and Y. Xie, Angew. Chem., Int. Ed., 2013, 52, 4361–4365 CrossRef CAS PubMed.
  13. M. W. Barsoum and M. Radovic, Annu. Rev. Mater. Res., 2011, 41, 195–227 CrossRef CAS.
  14. Z. Sun, Int. Mater. Rev., 2011, 56, 143–166 CrossRef CAS PubMed.
  15. Z. Sun, R. Ahuja, S. Li and J. M. Schneider, Appl. Phys. Lett., 2003, 83, 899–901 CrossRef CAS PubMed.
  16. Z. Sun, D. Music, R. Ahuja, S. Li and J. M. Schneider, Phys. Rev. B: Condens. Matter Mater. Phys., 2004, 70, 092102–092105 CrossRef.
  17. Z. Sun, M. Denis, R. Ahuja and M. S. Jochen, J. Phys.: Condens. Matter, 2005, 17, 7169–7176 CrossRef CAS.
  18. Y. Zhou and Z. Sun, Mater. Res. Innovations, 2000, 3, 286–291 CrossRef CAS.
  19. M. Khazaei, M. Arai, T. Sasaki, M. Estili and Y. Sakka, Sci. Technol. Adv. Mater., 2014, 15, 014208–014214 CrossRef.
  20. Z. Sun, J. Zhou, D. Music, R. Ahuja and J. M. Schneider, Scr. Mater., 2006, 54, 105–107 CrossRef CAS PubMed.
  21. Z. Sun and R. Ahuja, Appl. Phys. Lett., 2006, 88, 161913–161915 CrossRef PubMed.
  22. D. Music, Z. Sun, A. A. Voevodin and J. M. Schneider, Solid State Commun., 2006, 139, 139–143 CrossRef CAS PubMed.
  23. Z. Sun, Y. Zhou and J. Zhou, Philos. Mag. Lett., 2000, 80, 289–293 CrossRef CAS.
  24. D. Music, Z. Sun and J. M. Schneider, Solid State Commun., 2006, 137, 306–309 CrossRef CAS PubMed.
  25. D. Music, Z. Sun, R. Ahuja and J. M. Schneider, J. Phys.: Condens. Matter, 2006, 18, 8877–8881 CrossRef CAS.
  26. Z. Sun, D. Music, R. Ahuja and J. M. Schneider, Phys. Rev. B: Condens. Matter Mater. Phys., 2005, 71, 193402–193405 CrossRef.
  27. Z. Sun, S. Li, R. Ahuja and J. M. Schneider, Solid State Commun., 2004, 129, 589–592 CrossRef CAS PubMed.
  28. Z. Sun, R. Ahuja and J. M. Schneider, Phys. Rev. B: Condens. Matter Mater. Phys., 2003, 68, 224112–224115 CrossRef.
  29. D. Music, Z. Sun and J. M. Schneider, Phys. Rev. B: Condens. Matter Mater. Phys., 2005, 71, 092102–092104 CrossRef.
  30. M. Kurtoglu, M. Naguib, Y. Gogotsi and M. W. Barsoum, MRS Comm., 2012, 2, 133–137 CrossRef CAS.
  31. L. Qi and D. C. Chrzan, Phys. Rev. Lett., 2014, 112, 115503–115507 CrossRef.
  32. X. Li, S. Schönecker, J. Zhao, B. Johansson and L. Vitos, Phys. Rev. B: Condens. Matter Mater. Phys., 2013, 87, 214203–214214 CrossRef.
  33. C. Jiang and S. G. Srinivasan, Nature, 2013, 496, 339–342 CrossRef CAS PubMed.
  34. D. M. Clatterbuck, C. R. Krenn, M. L. Cohen and J. W. Morris, Phys. Rev. Lett., 2003, 91, 135501–135504 CrossRef CAS.
  35. D. M. Clatterbuck, D. C. Chrzan and J. W. Morris Jr, Acta Mater., 2003, 51, 2271–2283 CrossRef CAS.
  36. J. Hafner, J. Comput. Chem., 2008, 29, 2044–2078 CrossRef CAS PubMed.
  37. J. P. Perdew and Y. Wang, Phys. Rev. B: Condens. Matter Mater. Phys., 1992, 45, 13244–13249 CrossRef.
  38. J. P. Perdew, K. Burke and Y. Wang, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 16533–16539 CrossRef CAS.
  39. P. E. Blöchl, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50, 17953–17979 CrossRef.
  40. H. J. Monkhorst and J. D. Pack, Phys. Rev. B: Solid State, 1976, 13, 5188–5192 CrossRef.
  41. B. Silvi and A. Savin, Nature, 1994, 371, 683–686 CrossRef CAS.
  42. K. Momma and F. Izumi, J. Appl. Crystallogr., 2011, 44, 1272–1276 CrossRef CAS.
  43. L. Farber, M. W. Barsoum, A. Zavaliangos, T. El-Raghy and I. Levin, J. Am. Ceram. Soc., 1998, 81, 1677–1681 CrossRef CAS PubMed.
  44. M. W. Barsoum, L. Farber and T. El-Raghy, Metall. Mater. Trans. A, 1999, 30, 1727–1738 CrossRef.
  45. M. Zhang, M. Lu, Y. Du, L. Gao, C. Lu and H. Liu, J. Chem. Phys., 2014, 140, 174505–174510 CrossRef PubMed.
  46. S. Grimme, J. Comput. Chem., 2006, 27, 1787–1799 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.