Ab initio study of graphene-like monolayer molybdenum disulfide as a promising anode material for rechargeable sodium ion batteries

Jingcang Sua, Yong Peia, Zhenhua Yangb and Xianyou Wang*a
aKey Laboratory of Environmentally Friendly Chemistry and Applications of Minister of Education, School of Chemistry, Xiangtan University, Xiangtan 411105, Hunan, China. E-mail: wxianyou@yahoo.com; Fax: +86 732 58292061; Tel: +86 731 58292060
bFaculty of Materials, Optoelectronics and Physics, Key Laboratory of Low Dimensional Materials & Application Technology of Ministry of Education, Xiangtan University, Xiangtan 411105, Hunan, China

Received 2nd July 2014 , Accepted 26th August 2014

First published on 26th August 2014


Abstract

Using first-principles study based on density functional theory (DFT), the adsorption sites, diffusion kinetics, theoretical capacity and average voltage of Na atoms in graphene-like monolayer MoS2 are systematically investigated in comparison with bulk MoS2. It is found that for the graphene-like monolayer MoS2, a maximum theoretical capacity of 335 mA h g−1 could be achieved by double-side Na adsorption. Upon sodiation process, the graphene-like monolayer MoS2 can maintain a low voltage platform at about 1.0 V. A Na diffusion pathway on the graphene-like monolayer MoS2 is identified as from two adjacent T-sites passing through the nearest-neighbor H site in a zigzag manner. The activation barrier of this process is only 0.11 eV, a considerable decrease compared to that of the bulk MoS2 interlayer migration (0.70 eV), which indicates that Na can diffuse faster in the graphene-like monolayer MoS2 than in bulk MoS2. The present results suggest that the graphene-like monolayer MoS2 can provide excellent battery performance as the anode material of a sodium ion battery.


1. Introduction

Rechargeable sodium ion batteries as an alternative to lithium-ion batteries for large-scale energy storage have been undergoing rapid expansion due to the widespread abundance and low cost, and very suitable redox potential of sodium1–3 (image file: c4ra06557c-t1.tif versus standard hydrogen electrode, which is only 0.3 V above that of lithium). However, most of the scientific interest in Na-ion batteries has focused on cathode materials, few anode materials have been reported as viable. Interestingly, the reactivity of Li and Na with anode materials is quite different despite their close chemical properties. For instance, graphite, the most common anode material in Li-ion batteries, cannot be satisfactorily used as an insertion electrode in Na-ion batteries because Na atoms do not intercalate well between carbon sheets4,5 Na metal is also not a good choice because of its high reactivity and low melting point of 97.7 °C compared to 180.5 °C for Li. Thus, the identification of a novel anode with a proper Na storage voltage platform, large reversible capacity and high structural stability is a critical issue for the successful development of Na-ion batteries.

Molybdenum disulfide (MoS2), structurally similar to graphite, is constructed by three atom layers: a Mo layer sandwiched between two S layers. Similar to graphite, the MoS2 trilayers are held together by weak van der Waals forces. The large surface areas and interlayer spaces of MoS2 make it an attractive host for ion intercalation/deintercalation.6,7 Experimentally, bulk MoS2 has been investigated as electrode materials for lithium-ion batteries.8,9 However, large voluminal expansion usually results in their structural instability during the ion intercalation process.10 Recently, the breakthroughs in preparing graphene-like MoS2 nanostructures by mechanical or solvent-based exfoliation method11,12 endow MoS2 with high reversible lithium storage capacity, good cyclability, and superior rate performance.13–15 It is suggested that the exfoliated nanosheets with loose stacking can accommodate volume variation and thus alleviate the structural instability confronted by bulk crystals during the charge–discharge processes; furthermore, the ion diffusion path could be significantly shortened. Recently, several groups reported the experimental studies of MoS2 as a potential host for sodium ion intercalation/deintercalation,16,17 which open a new ground for rechargeable Na ion batteries. In this work, first principles calculations were used to systematically study the comparison of adsorption and diffusion of Na atoms on bulk and graphene-like monolayer MoS2. The theoretical capacity and average voltage for bulk and graphene-like monolayer MoS2 were also computed. Based on our study, it can be found that the graphene-like monolayer MoS2 may be a promising anode material for Na-ion batteries.

2. Computational details

2.1 Methodology

All the calculations were performed via the ab initio total energy and molecular-dynamics program VASP (Vienna ab initio-simulation program) developed at the Institute für Materialphysik of the Universität Wien.18,19 The interactions between valence electrons and ions were described with the projector augmented wave (PAW) pseudo-potentials.19 The GGA/PBE method was used in all the calculations of properties. The convergence tests of the total energy with respect to the k-points sampling and cut-off energy have been carefully examined, which ensure that the total energy is converged to 10−5 eV per formula unit. It is well-known that the formation of solid electrolyte interface (SEI) can occur on the surface of electrode materials and the surface is solvated. However, in this study, the adsorption and diffusion of Na atoms into idealized MoS2 systems were carried out. This method mainly calculated the behavior of metal atoms in idealized MoS2 systems to avoid the complexity of solvent and the presence of SEI. The Brillouin zones of bulk and graphene-like monolayer MoS2 were sampled with 5 × 5 × 5 and 5 × 5 × 1 k-points, respectively. Energy cut-off for the plane waves is chosen to be 400 eV. Prior to the calculation of the electronic structure, both the lattice parameters and the ionic position are fully relaxed, and the final forces on all the relaxed atoms are less than 0.01 eV Å−1. Especially, to accurately account for van der Waals interactions which might play an important role in layered structure, the PBE + D approach with the vdW correction20 was adopted for the calculation of bulk MoS2.

To study the diffusion kinetics of Na ion, energy barriers were calculated using the nudged elastic band (NEB) method.21,22 The NEB is an efficient method for determining the minimum energy path and saddle points between the given initial and final positions. It was performed with the linear interpolation of nine images between the initial and final configurations of the diffusion paths. The geometry and energy of the images were then relaxed until the largest norm of the force orthogonal to the path is less than 0.02 eV Å−1. Each image searches for its potential lowest energy configuration along the reaction path while maintaining equal distance to nearby images.

2.2 Models

In the bulk phase of MoS2, the MoS2 layers can be stacked in three patterns: 1T-MoS2 with the molybdenum atoms coordinated by the sulfur atoms and one Mo atom per unit cell (CdI2 type),23 2H-MoS2 with two-molecule layers per unit cell (P63/mmc)24 and 3R-MoS2 with trigonal prismatic coordination but with three S–Mo–S units in the c-axis direction.25 However, the most stable form of bulk phase MoS2 is 2H-MoS2, which is a stable form up to 1000 °C.26 Meanwhile, graphene-like monolayer MoS2 presents the sandwich-like structure with the Mo layer sandwiched between two S layers. Two crystal phases of graphene-like monolayer MoS2 have been reported: trigonal (1T) phase with trigonal prismatic coordination of Mo and hexagonal (1H) phase with octahedral coordination of Mo.27,28 At room temperature, 1H-phase is more stable than 1T-phase. Therefore, 2H type bulk MoS2 and 1H-phase structure graphene-like monolayer MoS2 are considered in our calculations.

To study the effects of Na concentration on the adsorption and diffusion of Na, we construct the 4 × 4 × 1 supercell for bulk MoS2, which contain 64 S atoms and 32 Mo atoms. For graphene-like monolayer MoS2, the supercell with a 4 × 4 slab and a 15 Å vacuum space in the periodic directions used to avoid interactions between two neighboring images of the supercell is constructed, containing 32 S atoms and 16 Mo atoms.

3. Results and discussion

3.1 Crystal structures of bulk and graphene-like monolayer MoS2

A full geometry optimization of both lattice parameters and atomic coordinates for 2H-type bulk MoS2 and 1H-phase graphene-like monolayer MoS2 structures was performed. The computed lattice constants are listed in Table 1 and compared with experimental data.29 It can be seen that the a-lattice constants and bond length of Mo–S for both bulk and graphene-like monolayer MoS2 calculated in GGA/PBE or PBE + vdW method are in good agreement with experimental data. For bulk MoS2 the c-lattice constant calculated in the GGA method produce an overestimated value of 13.78 Å. In contrast, the PBE + vdW approach shows an optimal value of 12.40 Å, close to the experimental value of 12.29 Å.
Table 1 Lattice parameters and bond length from DFT calculations in comparison to experimental data
  Bulk Monolayer Experiment29
PBE PBE + D PBE
a 3.18 3.19 3.18 3.16
c 13.78 12.40 12.29
c/a 4.33 3.89 3.89
dS–Mo 2.41 2.41 2.41 2.41


3.2 Adsorption of Na atoms on bulk and graphene-like monolayer MoS2

In the present study, different Na adsorption positions and amounts are given a comprehensive consideration. Binding energy (Eb) is used to represent the stability of the Na adsorbed system, which is defined as Eb = [EMoS2 + nENaENa@MoS2]/n, where ENa@MoS2 denotes the total energy of Na-adsorbed MoS2 (bulk or monolayer), EMoS2 is the total energy of the pristine MoS2 (bulk or monolayer), and ENa is the energy of an isolated Na atom. According to this definition, a more positive binding energy indicates a more favorable exothermic sodiation reaction between MoS2 and Na atoms.

First, single Na atom adsorption behavior of bulk and graphene-like monolayer MoS2 is examined. In bulk MoS2, there are two representative adsorption sites: one is the tetrahedral site (Ts) formed by one S atom from the upper triple layer and three S atoms from the lower triple layer (Fig. 1a) and the other is the octahedral site (Os), in which Na can bind to three S atoms from each of the triple layers (Fig. 1b). The total energy calculation results reveal that the adsorption site Os is energetically more stable than the Ts site in bulk MoS2. For graphene-like monolayer MoS2, there are also two stable adsorption sites: the hollow site (H) above the center of the hexagon and the top site (T) directly above one Mo atom, as shown in Fig. 1c and d. From the calculation results, the total energy of adsorption for site T is lower than site H in monolayer MoS2. Therefore, a single sodium atom adsorbs at the octahedral site (Os) in bulk MoS2 and adsorbs at the T site on graphene-like monolayer MoS2. The calculated binding energies of each site are shown in Table 2.


image file: c4ra06557c-f1.tif
Fig. 1 Side views of a Na atom adsorbed at Ts (a) and Os (b) sites in MoS2 bulk. Top (c) and side (d) views of a Na atom adsorbed on the graphene-like monolayer MoS2.
Table 2 Na bind energies and Bader charges in bulk and monolayer MoS2
  Bulk Monolayer
Ts Os H T
Eb (eV) 1.03 1.70 1.23 1.27
qNa (|e|) +0.79 +0.77 +0.86 +0.86


To quantitatively estimate chemical bonding and charge transfer between Na atoms and MoS2, Bader charge analysis30,31 is performed (Table 2). It can be seen that the bonding between the adsorbed Na and substrate MoS2 is predominantly ionic with a small part of covalence, and the valence electrons from the adsorbed Na atoms are transferred to the MoS2. The calculation results suggest Na atoms transfer 0.77|e| and 0.86|e| charge to bulk and graphene-like monolayer MoS2, respectively. The adsorbed Na forms six Na–S bonds in bulk MoS2, while in graphene-like monolayer MoS2 there are only three Na–S bonds, which will slow Na diffusion in bulk MoS2.

To further study the Na adsorption amount, the adsorption behavior of two Na atoms on the graphene-like monolayer MoS2 is examined. Five typical cases are illustrated: (a) unifacial adsorption on the top of two Mo atoms, (b) bifacial adsorption up and downward to one hollow site (H-site), (c) bifacial adsorption up and downward to one Mo atom (T-site), (d) bifacial adsorption up and downward to the two nearest neighboring Mo atoms, (e) bifacial adsorption up and downward to two next nearest neighboring Mo atoms, the geometries with the binding energy of each case are displayed in Fig. 2. The calculation results indicate that model (c) has the strongest binding energy, and the binding energy of model (d) is 0.005 eV lower than model (e). It suggests that two sodium atoms energetically prefer to coordinate upward and downward to the same Mo atom on graphene-like monolayer MoS2, and repulsive electrostatic interactions between Na+ cations lead to Na atoms coordinate decentralized. In this case, we construct the structures of more Na atoms adsorbed on the graphene-like monolayer MoS2 in a manner similar to that discussed above. Meanwhile, the adsorption behavior of two Na atoms in bulk MoS2 is studied based on the similar details and conditions as on graphene-like monolayer MoS2. The calculation results indicate that Na atoms prefer to adsorb upward and downward to the same Os site of bulk MoS2. Further investigation on binding energies as a function of the number of Na atoms for both bulk and graphene-like monolayer MoS2 is performed, as shown in Fig. 3. It can be seen that the Na binding energy of bulk MoS2 increases gradually with the increase in the number of Na atoms n from one to eight, and then displays a flat trend and maintains about 2.2 eV from 8 to 32 Na adatoms. However, the binding energies of graphene-like monolayer MoS2 are independent of sodium concentration. It displays a flat trend and maintains about 1.25 eV. The Na binding energy of graphene-like monolayer MoS2 remains 1.21 eV even at n = 32, indicating that Na atoms can still be stably adsorbed at graphene-like monolayer MoS2 and the phase separation problem can be safely avoided at such a high concentration.


image file: c4ra06557c-f2.tif
Fig. 2 Different potential positions investigated of two sodium atoms adsorbed on the graphene-like MoS2 monolayer. Binging energy (Eb) of each case is listed right below.

image file: c4ra06557c-f3.tif
Fig. 3 Binding energies with increasing number of Na atoms for bulk (image file: c4ra06557c-u1.tif) and the graphene-like monolayer MoS2 (image file: c4ra06557c-u2.tif).

3.3 Electrochemical performance

The Na storage capacity and average voltage of bulk and graphene-like monolayer MoS2 are further explored because it directly determines the potential applications of these kinds of materials. Because Na atoms can absorb on both the sides of the graphene-like monolayer MoS2, Na2MoS2 represent the highest Na storage capacity of monolayer. It can be deduced that the graphene-like monolayer MoS2 has a theoretical capacity of 335 mA h g−1. As a comparison, bulk MoS2 can only store Na atoms up to NaMoS2, the theoretical capacity is only half of the graphene-like monolayer MoS2. Therefore, the graphene-like monolayer MoS2, with fully exposed additional active sites for the accommodation of Na atoms can provide a higher power density than bulk MoS2.

Next, the average voltage for Na intercalation on graphene-like monolayer and bulk MoS2 can be calculated. The charge–discharge processes of MoS2 follow the common half-cell reaction vs. Na/Na+:

MoS2 + xNa+ + xe ↔ NaxMoS2

The theoretical procedure to calculate the average voltage through first principles can be established from the energy difference based on the equation below:32

image file: c4ra06557c-t2.tif
where image file: c4ra06557c-t3.tif, image file: c4ra06557c-t4.tif are the total energies of Nax1MoS2 and Nax2MoS2, ENa is the total energy of metallic sodium in a body-centered-cubic (bcc) phase, x2x1 refers to the number of sodium ions intercalated. According to this equation, the computed average voltage of the graphene-like monolayer MoS2 display a flat platform at about 1.0 V, which provides significant feasibility to be applied as Na ion battery anodes. In contrast, the calculated average voltage of bulk MoS2 is between 1.7–2.0 V, which is inapplicable for anode materials. Experimental data reported recently in literature on Na intercalation in MoS2 are qualitatively consistent with the theoretical predictions mentioned above. Park et al. reported that bulk MoS2 delivered a first discharge capacity of 190 mA h g−1 at 0.4–2.6 V by intercalation/deintercalation reaction and maintained a capacity of 84 mA h g−1 even after 100 cyles.8 In addition, an exfoliated MoS2/C composite has been reported as an excellent anode material for sodium ion batteries. The graphene-like exfoliated MoS2/C composite achieved a high capacity of ∼400 mA h g−1 at an average working voltage of 1.0 V for 100 cycles.33 Accordingly, it can be suggested that the graphene-like monolayer MoS2 is more suitable as an anode for Na ion battery electrode material because of its lower electrode potential and higher capacity than bulk MoS2.

3.4 Na diffusion

It is known that the rate performance of sodium ion battery electrode material is mainly determined by the electrical conductivity and sodium diffusion rate. Fast Na diffusion determines high charge–discharge rates of batteries. Thus, it is significant to study the diffusion of sodium atoms on MoS2 in terms of its actual application as the anode material of Na ion batteries. Na migration in the interlayer of bulk MoS2 is firstly considered. The diffusion path is investigated between two adjacent octahedral-sites (Os1–Os2). The diffusion pathway and energy barrier are shown in Fig. 4a–c. The computed diffusion barrier for this path is 0.70 eV. For comparison, Na migration in the graphene-like monolayer MoS2 is also examined. In this case, Na migration between two adjacent T-sites (T1–T2) is investigated because of the energetic favorability to the H sites. It can be found that the specific path of sodium diffusion is from a T1 site to an adjacent T2 site passing through the nearest-neighbor H site in a zigzag manner, as is presented in Fig. 4d–f. The energy barrier of this process is only 0.11 eV, which is considerably lower than that of bulk MoS2. According to the Arrhenius equation, the diffusion constant (D) is proportional to exp(−Ebarrier/kBT), where Ebarrier and kB are the activation energy and Boltzmann constant, respectively. Therefore, it can be concluded that at room temperature Na diffusion on the graphene-like monolayer MoS2 is about 109 faster than on bulk MoS2. The significantly enhanced mobility of the graphene-like monolayer MoS2 makes it more attractive for promising anode materials with fast charge–discharge rates.
image file: c4ra06557c-f4.tif
Fig. 4 Na migration path in the interlayer of bulk MoS2 and the corresponding energy profiles along the Na diffusion path: (a) side view (b) top view (c) diffusion path; Na diffusion path on the surface of the graphene-like monolayer MoS2 and corresponding energy profiles along the Na diffusion path: (d) side view (e) top view (f) diffusion path.

4. Conclusion

In conclusion, based on density functional theory (DFT), the adsorption, diffusion properties, as well as the maximum capacity and average voltage of the graphene-like monolayer MoS2 in comparison with bulk MoS2 were systematically investigated by first principles calculations. The results demonstrated that the monolayer can stably adsorb Na up to Na2MoS2, which converts to a higher theoretical capacity (335 mA h g−1). Upon sodiation processes, the graphene-like monolayer MoS2 can maintain a lower and applicable voltage platform (1.0 V) compared to bulk MoS2 (1.7–2.0 V). Although the bulk MoS2 have higher Na binding energies, the lower Na mobilities do not make them ideal candidates for anode materials. Na mobility can be considerably enhanced by the dimensionality reduction, the diffusion barrier decreases from 0.70 eV to 0.11 eV because the MoS2 changes from bulk to a monolayer structure. Therefore, because of high stability and low Na diffusion energy, high theoretical capacity, low average voltage, the graphene-like monolayer MoS2 can be used as a promising anode material for high performance rechargeable Na-ion batteries.

Acknowledgements

This work is funded by the National Natural Science Foundation of China under project no. 51472211, Scientific and Technical Achievement Transformation Fund of Hunan Province under project no. 2012CK1006, Key Project of Strategic New Industry of Hunan Province under project no. 2013GK4018, and Science and Technology plan Foundation of Hunan Province under project no. 2013FJ4062. YP is supported by Natural Science Foundation of China (Grant no. 21103144) and Hunan Provincial Natural Science Foundation of China (12JJ7002, 12JJ1003).

References

  1. S. Kim, D. Seo and X. Ma, et al. Electrode Materials for rechargeable sodium-ion batteries potential alternatives to current lithium-ion batteries[J], Adv. Energy Mater., 2012, 2(7), 710–721 CrossRef PubMed.
  2. V. Palomares, P. Serras and I. Villaluenga, et al. Na-ion batteries, recent advances and present challenges to become low cost energy storage systems[J], Energy Environ. Sci., 2012, 5(3), 5884–5901 CAS.
  3. M. D. Slater, D. Kim and E. Lee, et al. Sodium-Ion Batteries[J], Adv. Funct. Mater., 2012, 123(8), 947–958 Search PubMed.
  4. J. Sangster, C–Na, (Carbon–Sodium) System[J], J. Phase Equilib. Diffus., 2007, 28(6), 571–579 CrossRef CAS.
  5. D. A. Stevens and J. R. Dahn, The mechanisms of lithium and sodium insertion in carbon materials[J], J. Electrochem. Soc., 2001, 148(8), A803–A811 CrossRef CAS PubMed.
  6. H. Wang, H. Feng and J. Li, Graphene and Graphene-like Layered Transition Metal Dichalcogenides in Energy Conversion and Storage[J], Small, 2014, 10(11), 2165–2181 CrossRef CAS PubMed.
  7. M. Chhowalla, H. S. Shin and G. Eda, et al. The chemistry of two-dimensional layered transition metal dichalcogenide nanosheets[J], Nat. Chem., 2013, 5(4), 263–275 CrossRef PubMed.
  8. J. Park, J. Kim and J. Park, et al. Discharge mechanism of MoS2 for sodium ion battery: Electrochemical measurements and characterization[J], Electrochim. Acta, 2013, 92, 427–432 CrossRef CAS PubMed.
  9. T. Stephenson, Z. Li and B. Olsen, et al. Lithium ion battery applications of molybdenum disulfide (MoS2) nanocomposites[J], Energy Environ. Sci., 2014, 7(1), 209–231 CAS.
  10. K. Chang and W. Chen, L-Cysteine-assisted synthesis of layered MoS2/graphene composites with excellent electrochemical performances for lithium ion batteries[J], ACS Nano, 2011, 5(6), 4720–4728 CrossRef CAS PubMed.
  11. B. Radisavljevic, A. Radenovic and J. Brivio, et al. Single-layer MoS2 transistors[J], Nat. Nanotechnol., 2011, 6(3), 147–150 CrossRef CAS PubMed.
  12. C. Lee, Q. Li and W. Kalb, et al. Frictional characteristics of atomically thin sheets[J], Science, 2010, 328(5974), 76–80 CrossRef CAS PubMed.
  13. K. Chang, W. Chen and L. Ma, et al. Graphene-like MoS2/amorphous carbon composites with high capacity and excellent stability as anode materials for lithium ion batteries[J], J. Mater. Chem., 2011, 21(17), 6251–6257 RSC.
  14. J. Xiao, D. Choi and L. Cosimbescu, et al. Exfoliated MoS2 nanocomposite as an anode material for lithium ion batteries[J], Chem. Mater., 2010, 22(16), 4522–4524 CrossRef CAS.
  15. G. Du, Z. Guo and S. Wang, et al. Superior stability and high capacity of restacked molybdenum disulfide as anode material for lithium ion batteries[J], Chem. Commun., 2010, 46(7), 1106–1108 RSC.
  16. L. David, R. Bhandavat and G. Singh, MoS2/graphene Composite Paper For Sodium-Ion Battery Electrodes[J], ACS Nano, 2014, 8(2), 1759–1770 CrossRef CAS PubMed.
  17. S. Chou, Exfoliated MoS2/C Composite As a High-Performance Anode for Sodium Ion Batteries[C], presented at the 17th International Meeting on Lithium Batteries, June 10–14, 2014.
  18. G. Kresse and J. Furthmüller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set[J], Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54(16), 11169 CrossRef CAS.
  19. G. Kresse and D. Joubert, From ultrasoft pseudopotentials to the projector augmented-wave method[J], Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 59(3), 1758 CrossRef CAS.
  20. F. Ortmann, F. Bechstedt and W. G. Schmidt, Semiempirical van der Waals correction to the density functional description of solids and molecular structures[J], Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 73(20), 205101 CrossRef.
  21. G. Henkelman and H. Jonsson, Improved tangent estimate in the nudged elastic band method for finding minimum energy paths and saddle points[J], J. Chem. Phys., 2000, 113(22), 9978–9985 CrossRef CAS PubMed.
  22. R. A. Olsen, G. J. Kroes and G. Henkelman, et al. Comparison of methods for finding saddle points without knowledge of the final states[J], J. Chem. Phys., 2004, 121(20), 9776–9792 CrossRef CAS PubMed.
  23. F. Wypych and R. Schollhorn, 1T-MoS2, a new metallic modification of molybdenum disulfide[J], Chem. Commun., 1992,(19), 1386–1388 RSC.
  24. G. Seifert, H. Terrones and M. Terrones, et al. Structure and electronic properties of MoS2 nanotubes[J], Phys. Rev. Lett., 2000, 85(1), 146 CrossRef CAS.
  25. B. Schonfeld, J. J. Huang and S. C. Moss, Anisotropic mean-square displacements (MSD) in single-crystals of 2H-and 3R-MoS2[J], Acta Crystallogr., Sect. B: Struct. Sci., 1983, 39(4), 404–407 CrossRef.
  26. J. A. Wilson and A. D. Yoffe, The transition metal dichalcogenides discussion and interpretation of the observed optical, electrical and structural properties[J], Adv. Phys., 1969, 18(73), 193–335 CrossRef CAS.
  27. M. A. Py and R. R. Haering, Structural destabilization induced by lithium intercalation in MoS2 and related compounds[J], Can. J. Phys., 1983, 61(1), 76–84 CrossRef CAS PubMed.
  28. V. Alexiev, R. Prins and T. Weber, DFT study of MoS2 and hydrogen adsorbed on the (101 [combining macron] 0) face of MoS2[J], Phys. Chem. Chem. Phys., 2001, 3(23), 5326–5336 RSC.
  29. J. A. Wilson and A. D. Yoffe, The transition metal dichalcogenides discussion and interpretation of the observed optical, electrical and structural properties[J], Adv. Phys., 1969, 18(73), 193–335 CrossRef CAS.
  30. R. F. Bader, Atoms in molecules[M], Wiley Online Library, 1990 Search PubMed.
  31. G. Henkelman, A. Arnaldsson and H. Jonsson, A fast and robust algorithm for Bader decomposition of charge density[J], Comput. Mater. Sci., 2006, 36(3), 354–360 CrossRef PubMed.
  32. M. K. Aydinol, A. F. Kohan and G. Ceder, et al. Ab initio study of lithium intercalation in metal oxides and metal dichalcogenides[J], Phys. Rev. B: Condens. Matter Mater. Phys., 1997, 56(3), 1354 CrossRef CAS.
  33. Y. Wang, K. H. Seng and S. Chou, et al. Reversible sodium storage via conversion reaction of a MoS2–C composite[J], Chem. Commun., 2014, 50, 10730–10733 RSC.

This journal is © The Royal Society of Chemistry 2014
Click here to see how this site uses Cookies. View our privacy policy here.