Next Article in Journal
Evaluating Superconductors through Current Induced Depairing
Next Article in Special Issue
High-Energy X-Ray Compton Scattering Imaging of 18650-Type Lithium-Ion Battery Cell
Previous Article in Journal
Scaling between Superfluid Density and Tc in Overdoped La2−xSrxCuO4 Films
Previous Article in Special Issue
Fingerprint Oxygen Redox Reactions in Batteries through High-Efficiency Mapping of Resonant Inelastic X-ray Scattering
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Understanding Phase Stability of Metallic 1T-MoS2 Anodes for Sodium-Ion Batteries

1
Department of Physics, Northeastern University, Boston, MA 02115, USA
2
Department of Mechanical and Industrial Engineering, Northeastern University, Boston, MA 02115, USA
3
Department of Physics, School of Engineering Science, LUT University, FI-53851 Lappeenranta, Finland
*
Authors to whom correspondence should be addressed.
Condens. Matter 2019, 4(2), 53; https://doi.org/10.3390/condmat4020053
Submission received: 24 April 2019 / Revised: 2 June 2019 / Accepted: 3 June 2019 / Published: 10 June 2019

Abstract

:
We discuss metallic 1T-MoS2 as an anode material for sodium-ion batteries (SIBs). In situ Raman is used to investigate the stability of metallic MoS2 during the charging and discharging processes. Parallel first-principles computations are used to gain insight into the experimental observations, including the measured conductivities and the high capacity of the anode.

1. Introduction

The discovery of the rechargeable LiCoO2 cathode for lithium-ion batteries (LIBs) at the end of the 1970s [1] ushered in a new era of portable clean energy storage. Since then, LIBs have been adapted and utilized extensively, powering everything from portable devices to vehicles to sectors of the power grid. However, the limited storage capacity of the LIBs has spurred intense interest in the cheaper sodium-ion batteries (SIBs). A key factor in this connection is the choice of the anode material because the widely used graphite anode in commercial LIBs possesses a relatively low capacity of 372 mAh∙g−1, and it also has a very limited capacity for storing Na ion [2,3]. Further improvements in performance of Li/Na-ion batteries call for a search for other advanced electrode materials for both the cathode and the anode.
In order to ameliorate the deficiencies of the graphite anode, materials beyond graphene provide a promising new pathway. In this connection, two-dimensional transition metal dichalcogenides (2D-TMDs) have gained attention as a possible replacement for graphite as the anode material due to their large surface to volume ratio, which is predicted to provide high Li and Na capacities [4,5] and low diffusion barriers to enable high cycling rates [6,7]. 2D-TMDs, with the general formula MX2, where M denotes a transition metal atom (e.g., Mo or W) and X a chalcogen atom (S, Se, or Te), are layered materials composed of planar sheets with strong in-plane bonds and layers bonded weakly by van der Waals interactions, facilitating the isolation of single or few layers similar to graphene. The coordination of the chalcogens to the metal atoms can either be trigonal-prismatic or octahedral, producing semiconducting or metallic behavior, respectively. The metallic 1T phase of MoS2 has shown promise as an anode material due to its high electrical conductivity that promotes the performance of the battery and eliminates the use of conductive additives [8].
This paper discusses experiments performed on a 1T-MoS2 anode, along with parallel first-principles calculations based on the density functional theory (DFT). We systematically studied the intercalation of Li and Na into a 1T-MoS2 matrix for various concentrations. For low Na or Li concentration, AXMoS2 (A = Li, Na) shows good conductivity with a relatively large electrode voltage. At x = 1, 1T-MoS2 develops a band gap concomitant with the appearance of a Jahn–Teller distortion driven by changes in d-electron count, which can be mitigated via disorder to restore the good conductivity of the anode. For x > 1, the anode is always found to be metallic, and here we introduce a Na multilayer model to understand this high-capacity regime.

2. Materials and Methods

Computational—The ab initio calculations were performed using the pseudopotential projected augmented wave method [9] implemented in the Vienna ab initio simulation package [10,11] with an energy cutoff of 400 eV for the plane-wave basis set. Exchange-correlation effects were treated using the generalized gradient approximation (GGA) [12], and van der Waals corrections were included using the DFT-D2 method of Grimme [13], where a 14 × 14 × 7 Γ-centered k-point mesh was used to sample the Brillouin zone. A large enough vacuum of 15 Å in the z direction was used to ensure negligible interaction between the periodic images of the Na layer doubled MoS2 thin film. All the structures were relaxed using a conjugate gradient algorithm with an atomic force tolerance of 0.01 eV/Å and a total energy tolerance of 10−6 eV.
Experimental—MoO3 (18 mg, Fisher Scientific, Hampton, NH, USA), thioacetamine (21 mg, Sigma-Aldrich, St. Louis, MO, USA), and urea (0.15 g, Sigma-Aldrich, USA) were dissolved in ethanol (15 mL) and stirred for 1 h. Then the solution was transferred to a Teflon-lined stainless-steel autoclave. The autoclave was kept in a furnace for 16 h at 200 °C. After cooling to room temperature, the product was washed with ethanol a few times and dispersed in ethanol for future use.
Structural and Physical Characterization—The morphology of the as-prepared MoS2 was characterized by scanning electron microscopy (SEM) (Hitachi S4800) and transmission electron microscopy (TEM) (JEOL 1010). Raman spectroscopy was carried out on a Thermo Scientific DXR with 532 nm laser excitation. The set up for the in-situ Raman is a three-electrode cell with optical window from EL-CELL.
Electrochemical Measurement—Standard CR2025-type coin cells were assembled to measure the electrochemical performance of the as-synthesized metallic MoS2. The metallic MoS2 electrodes were directly assembled into the coin cells in an argon filled glove-box, using 1M NaPF6 dissolved in ethylene carbonate (EC) and dimethyl carbonate (DMC) (1:1 vol/vol) as an electrolyte. Cyclic voltammetry (CV) and electrochemical impedance spectra (EIS) were measured using a biologic SP-150. Galvanostatic charge–discharge tests were performed on a multichannel battery testing system (Land CT2001A).

3. Results and Discussion

Figure 1a shows a schematic of the sodiation–desodiation process of a 1T-MoS2 electrode prepared following the method described in Ref. [8]. The morphology of this electrode has been characterized by scanning electron microscopy (SEM) and transmission electron microscopy (TEM). Figure 1b,c is the SEM images of 1T-MoS2, which show the presence of lamellar structures with a vertically arrayed structure. The formation of an open structure of 1T-MoS2 nanosheets favors the accessibility of the Na ions. Moreover, the TEM image (Figure 1d) shows that 1T-MoS2 nanosheets are uniformly distributed in agreement with the SEM results (Figure 1d). Finally, Figure 1e provides a high resolution TEM image of the atomic lattice.
In-situ Raman spectroscopy in a specially designed cell with an optical window provides further insight into intercalation mechanism of Na ion. Figure 2a shows the in-situ Raman spectra collected at various cut-off potentials during the first discharge process. The two major peaks at 150 cm−1 (J1) and 220 cm−1 (J2) can be seen, which originate from the phononic modes of the metallic 1T-MoS2 [14,15,16]. When discharging the cell from 2.5 V to 0.05 V, the J1 and J2 Raman peaks remain without the appearance of any other significant features, suggesting that the 1T-MoS2 structural phase does not change. Figure 2b presents the Raman spectra during charging from 0.05 V to 3.0 V. Again, the J1 and J2 peaks are clearly seen, confirming the stability of the metallic 1T-MoS2 anode during charging. Beyond the robust presence of the J1 and J2 peaks, other weak peaks can be seen in the Raman spectra during charge/discharge in Figure 2. We believe that these spectral features are due to the electrolyte or background, but further study is needed to determine their origin. The composition of the electrolyte is 1M NaPF6 dissolved in ethylene carbonate (EC) and dimethyl carbonate (DMC) (1:1 vol/vol).
Figure 3 displays the electrochemical performance of the 1T-MoS2 for Na ion storage. Due to its good electronic conductivity, 1T-MoS2 can be directly coated on the stainless-steel current collector to work as anode without conductive additive and binder. Figure 3a shows the cyclic voltammetry (CV) curves of 1T-MoS2 anode in initial three cycles at a scan rate of 0.1 mV∙s−1 between 0.01 and 3.0 V (vs. Na+/Na). The first cycle is clearly different from the second and third cycles due to the formation of an irreversible solid-electrolyte-interface (SEI) film. Figure 3b shows the rate performance of the battery. The reversible capacity is 597, 384, 326, 297, 228, 159, 106, and 56 mAh∙g−1 at current densities of 0.08, 0.2, 0.5, 1.0, 2.0, 5.0, 10, and 20 A∙g−1, respectively. The 1T-MoS2 electrode can reach a reversible capacity of 330 mAh∙g−1 when the current density goes back to 80 mA∙g−1 even after cycling at high current densities. Figure 3c gives galvanostatic charge–discharge profiles of the 1T-MoS2 anode in different cycles at the current density of 1 A∙g−1. The initial coulombic efficiency is 57%, which is associated with SEI formation in the first cycle. Figure 3d shows the Nyquist plot of the battery. The cycling performance of 1T-MoS2 under the high current density of 1 A∙g−1 maintain 394 mAh∙g−1 after 350 cycles as illustrated in Figure 3d.
One of the main theoretical questions surrounding the use of TMDs as anode material in Na ion batteries is the phase stability of the 2D framework under intercalation. For this purpose, we examine the consequences of using 1T-MoS2 over the conventional 2H phase for various concentrations of Na ion via ground state, first-principles calculations to address the structural and electronic stabilities of these phases.
For specific capacities over the range 0–146 mAh∙g−1, we considered a 4 × 4 supercell of A-A stacked pristine 1T- and 2H-MoS2 structures, as shown in Figure 4b’,b respectively, where the Mo atom (purple sphere) in the 2H structure is trigonal-prismatically coordinated with S (yellow sphere), while in the 1T phase it is octahedrally coordinated. Under intercalation, the structures are relaxed upon Na (Li) ion insertion by minimizing the total energy (Figure 4c,d,c’,d’), as in [17].
For specific capacities greater than 146 mAh∙g−1, the MoS2 unit cell must contain more than one Na atom. By considering the transition from the usual stage 1 arrangement (Figure 5b) to a configuration with multiple layers of Na between the MoS2 sheets [18] (Figure 5c), capacities greater than 146 mAh∙g−1 can be achieved. Furthermore, in the presence of extra Na layers, the MoS2 sheets only weakly interact through van der Waals inter-sheet forces. Thus, this regime can be modeled by considering just a single MoS2 sheet with successive layers of Na, as shown in Figure 4e,f,e’,f’. The enhancement in capacity has previously been explained by a redox process wherein the NaxMoS2 compound is decomposed into metallic Mo and a hexagonal NaxS complex [19]. However, due to the persistence of the J1 and J2 peaks in the Raman spectra (Figure 2) and the continued high capacity cycling as seen in Figure 3c, this redox process can be ruled out in favor of an intercalation mechanism.
In order to compare the structural stability of the 2H and 1T phases, we calculated the stabilization energy defined as the difference in total energies per molybdenum atom, Es = E2H − E1T as a function of ion intercalation. Figure 4a compares the stabilization energies so obtained for both Na (red) and Li (blue) intercalation. The stability of the 2H phase is seen to decrease with ion concentration up to stage 1. In contrast, the stabilization energy for the 1T phase is found to increase with Na concentration, becoming the ground state for specific capacities above 51.98 mAh∙g−1. This result suggests that 1T-MoS2 is an attractive anode material since it avoids complications derived from structural phase transitions under cycling.
Physically, the stabilization of the 1T phase is mainly driven by a buckling of the Mo layer. The ground state coordination of the transition metals in TMDs varies systematically with d electron count. For instance, Group 4 metals are all octahedrally coordinated and Group 5 metals mostly have octahedral structures, although some are trigonal-prismatically coordinated. For Group 6 metals, however, the trigonal coordination is dominant with few octahedral exceptions. Group 7 metals form distorted octahedra, where the distortion is characterized by the transition metal slipping off the center of the regular octahedra in such a manner as to form infinite one-dimensional diamond chains. This phase is designated as 1T’ [20]. Therefore, as Na ions are introduced into the 1T-MoS2 matrix, electronic carriers accumulate in the Mo d-t2g states, effectively changing the d-electron count. This intercalation thus transforms Mo from a Group 6 to an effective Group 7 transition metal, thereby inducing Mo-Mo buckling characteristic of the 1T’ phase. As elucidated by Kertesz and Hoffmann [21], the 1T’ phase of ReS2 is described by a Jahn–Teller distortion in analogy to the degenerate orbitals in C4H4. Here, the experimental stabilization of the 1T’ phase is confirmed by the presence of the J1 and J2 peaks in the Raman spectra (Figure 2), which originate from the phononic modes of the one-dimensional diamond chains [14,15,16]. These modes are continually observed up to high capacity, confirming the robustness of the 1T’ structure.
Figure 5a shows the calculated voltage measured with respect to bulk Na (Li) as a function of specific capacity for each structural phase. Both phases generally exhibit trends in voltage-capacity curves similar to those seen in experiments, although the 1T’ phase exhibits a larger initial voltage in agreement with experiment. In the high capacity regime, the voltage is found to level out and become marginal, indicating negligible formation energy gained per additional Na atom as in the experiment, validating the Na multi-layer model. We also found a sensitive dependence of the voltage on the in-plane Mo-Mo buckling.
Figure 6 shows the site-projected density of states (DOS) for various Na concentrations in the 2H and 1T phases. As expected, the 2H phase is initially semiconducting with a band gap of 1.0 eV. Doped electrons introduced by Na filling in the conduction band of the pristine material follow a simple rigid band model. The metallic character of the states is maintained up through high Na concentrations. The DOS in the 1T’ phase retains a predominant metallic character except at a few particular fillings such as x = 1 due to a reorganization of the t2g states induced by the Jahn–Teller distortion. However, this band gap is not robust against disorder effects present in the real material, leaving the macroscopic system mostly metallic over the full range of Na concentrations. The observed good electron transport in the anode is thus explained by the predominantly metallic character of the material.

4. Conclusions

In summary, our study demonstrates that DFT is successful in probing details of the electronic structure of the NaxMoS2 anode material for SIBs. In particular, our results provide an explanation for the observed electronic conductivity and high capacity of the 1T-MoS2 anode. Moreover, in order to explain the high capacity of this anode, we have considered a model with multiple layers of Na placed between the 1T-MoS2 sheets. Our analysis shows that the 1T-MoS2 layer exhibits stable adhesion to the surface of the Na ion multilayer to enable uniform flow of Na ions into and out of the Na ion multilayer, providing a pathway for optimizing rechargeable SIBs based on 1T-MoS2.

Author Contributions

H.L., Y.J., D.C., and H.Z. designed and performed experiments, and analyzed data. C.L., B.B., and A.B. were responsible for the theoretical calculations and the interpretation of the results. All authors contributed to the writing of the manuscript.

Funding

H.Z. acknowledges the financial startup support and Tier 1 support from Northeastern University. The theoretical work was supported by the US Department of Energy (DOE), Office of Science, Basic Energy Sciences grant number DE-FG02-07ER46352 (core research), and benefited from Northeastern University’s Advanced Scientific Computation Center (ASCC), the NERSC supercomputing center through DOE grant number DE-AC02-05CH11231, and support (testing the efficacy of advanced functionals in complex materials) from the DOE EFRC: Center for Complex Materials from First Principles (CCM), under DOE grant number DE-SC0012575.

Acknowledgments

We thank the Kostas Research Institute at Northeastern University and the Center for Nanoscale System (CNS) at Harvard University for allowing us the use of their facilities. The theoretical work was also supported by Research Computing at Northeastern University by providing high-performance computing and storage through the Discovery Cluster.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Goodenough, J.B. How we made the Li-ion rechargeable battery. Nat. Electron. 2018, 1, 204. [Google Scholar] [CrossRef]
  2. Winter, M.; Besenhard, J.O.; Spahr, M.E.; Novák, P. Insertion electrode materials for rechargeable lithium batteries. Adv. Mater. 1998, 10, 725–763. [Google Scholar] [CrossRef]
  3. Fukuda, K.; Kikuya, K.; Isono, K.; Yoshio, M. Foliated natural graphite as the anode material for rechargeable lithium-ion cells. J. Power Sources 1997, 69, 165–168. [Google Scholar] [CrossRef]
  4. Tan, X.; Cabrera, C.R.; Chen, Z. Metallic BSi3 silicene: A promising high capacity anode material for lithium-ion batteries. J. Phys. Chem. C 2014, 118, 25836–25843. [Google Scholar] [CrossRef]
  5. Kulish, V.V.; Malyi, O.I.; Persson, C.; Wu, P. Phosphorene as an anode material for Na-ion batteries: A first-principles study. Phys. Chem. Chem. Phys. 2015, 17, 13921–13928. [Google Scholar] [CrossRef] [PubMed]
  6. Shu, H.; Li, F.; Hu, C.; Liang, P.; Cao, D.; Chen, X. The capacity fading mechanism and improvement of cycling stability in MoS2-based anode materials for lithium-ion batteries. Nanoscale 2016, 8, 2918–2926. [Google Scholar] [CrossRef] [PubMed]
  7. Ahmed, B.; Anjum, D.H.; Hedhili, M.N.; Alshareef, H.N. Mechanistic insight into the stability of HfO2-coated MoS2 nanosheet anodes for sodium ion batteries. Small 2015, 11, 4341–4350. [Google Scholar] [CrossRef]
  8. Jiao, Y.; Mukhopadhyay, A.; Ma, Y.; Yang, L.; Hafez, A.M.; Zhu, H. Ion transport nanotube assembled with vertically aligned metallic MoS2 for high rate lithium-ion batteries. Adv. Energy Mater. 2018, 8, 1702779. [Google Scholar] [CrossRef]
  9. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  10. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab-initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  11. Kresse, G.; Hafner, J. Ab initio molecular dynamics for open-shell transition metals. Phys. Rev. B 1993, 48, 13115–13118. [Google Scholar] [CrossRef]
  12. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  13. Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef] [PubMed]
  14. Tan, S.J.R.; Sarkar, S.; Zhao, X.; Luo, X.; Luo, Y.Z.; Poh, S.M.; Abdelwahab, I.; Zhou, W.; Venkatesan, T.; Chen, W.; et al. Temperature- and phase-dependent phonon renormalization in 1T’-MoS2. ACS Nano 2018, 12, 5051–5058. [Google Scholar] [CrossRef] [PubMed]
  15. Yin, X.; Wang, Q.; Cao, L.; Tang, C.S.; Luo, X.; Zheng, Y.; Wong, L.M.; Wang, S.J.; Quek, S.Y.; Zhang, W.; et al. Tunable inverted gap in monolayer quasi-metallic MoS2 induced by strong charge-lattice coupling. Nat. Commun. 2017, 8, 486. [Google Scholar] [CrossRef]
  16. Voiry, D.; Mohite, A.; Chhowalla, M. Phase engineering of transition metal dichalcogenides. Chem. Soc. Rev. 2015, 44, 2702–2712. [Google Scholar] [CrossRef] [PubMed]
  17. Mortazavi, M.; Wang, C.; Deng, J.; Shenoy, V.B.; Medhekar, N.V. Ab initio characterization of layered MoS2 as anode for sodium-ion batteries. J. Power Sources 2014, 268, 279–286. [Google Scholar] [CrossRef]
  18. Kühne, M.; Börrnert, F.; Fecher, S.; Ghorbani-Asl, M.; Biskupek, J.; Samuelis, D.; Krasheninnikov, A.V.; Kaiser, U.; Smet, J.H. Reversible superdense ordering of lithium between two graphene sheets. Nature 2018, 564, 234. [Google Scholar] [CrossRef] [PubMed]
  19. Wang, X.; Shen, X.; Wang, Z.; Yu, R.; Chen, L. Atomic-scale clarification of structural transition of MoS2 upon sodium intercalation. ACS Nano 2014, 8, 11394–11400. [Google Scholar] [CrossRef]
  20. Wilson, J.A.; Yoffe, A.D. The transition metal dichalcogenides discussion and interpretation of the observed optical, electrical and structural properties. Adv. Phys. 1969, 18, 193–335. [Google Scholar] [CrossRef]
  21. Kertesz, M.; Hoffmann, R. Octahedral vs. trigonal-prismatic coordination and clustering in transition-metal dichalcogenides. J. Am. Chem. Soc. 1984, 106, 3453–3460. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic diagram of sodiation /desodiation of 1T-MoS2 electrode; (b,c) SEM images of 1T-MoS2; (d) TEM image of 1T-MoS2; (e) HRTEM image of 1T-MoS2.
Figure 1. (a) Schematic diagram of sodiation /desodiation of 1T-MoS2 electrode; (b,c) SEM images of 1T-MoS2; (d) TEM image of 1T-MoS2; (e) HRTEM image of 1T-MoS2.
Condensedmatter 04 00053 g001
Figure 2. In situ Raman spectra of the 1T-MoS2 electrode at different voltages during (a) discharging and (b) charging processes.
Figure 2. In situ Raman spectra of the 1T-MoS2 electrode at different voltages during (a) discharging and (b) charging processes.
Condensedmatter 04 00053 g002
Figure 3. (a) CV curves of 1T-MoS2 for three initial cycles at scan rate of 0.1 mV∙s−1. (b) Rate performance of 1T-MoS2 at current densities of 0.08, 0.2, 0.5, 1.0, 2.0, 5.0, 10, and 20 A∙g−1. (c) Charge–discharge curves of 1T-MoS2 electrode at different cycles. (d) Nyquist plot of 1T-MoS2 electrode. (e) Long-term cycling performance of 1T-MoS2 electrode at current density of 1 A∙g−1.
Figure 3. (a) CV curves of 1T-MoS2 for three initial cycles at scan rate of 0.1 mV∙s−1. (b) Rate performance of 1T-MoS2 at current densities of 0.08, 0.2, 0.5, 1.0, 2.0, 5.0, 10, and 20 A∙g−1. (c) Charge–discharge curves of 1T-MoS2 electrode at different cycles. (d) Nyquist plot of 1T-MoS2 electrode. (e) Long-term cycling performance of 1T-MoS2 electrode at current density of 1 A∙g−1.
Condensedmatter 04 00053 g003
Figure 4. (a) Comparison of the stabilization energy per molybdenum atom under Na and Li ion intercalation. (bf) Relaxed crystal structures of 2H-NaxMoS2 for various Na ion concentrations. (b’f’) same as (bf) except for 1T/1T’-NaxMoS2.
Figure 4. (a) Comparison of the stabilization energy per molybdenum atom under Na and Li ion intercalation. (bf) Relaxed crystal structures of 2H-NaxMoS2 for various Na ion concentrations. (b’f’) same as (bf) except for 1T/1T’-NaxMoS2.
Condensedmatter 04 00053 g004
Figure 5. Schematic of the Na layer-doubling process used to generate specific capacities greater than 146 mAh∙g−1 of the stage 1 structure. (a) Voltage verses specific capacity for both the 2H and 1T/1T’ structural phases under Na ion intercalation. The voltage for Li is also given for comparison. (b) Stage 1 structure where the MoS2 sheets (rectangles with violet shading) are separated by one layer of Na (gold circles) giving a stoichiometric ratio of 1:1 between Mo and Na. (c) A ‘beyond’ stage 1 structure displaying 2 layers of Na between the MoS2 sheets yielding a capacity of 260 mAh∙g−1. The unit cell for each structure is highlighted with red line.
Figure 5. Schematic of the Na layer-doubling process used to generate specific capacities greater than 146 mAh∙g−1 of the stage 1 structure. (a) Voltage verses specific capacity for both the 2H and 1T/1T’ structural phases under Na ion intercalation. The voltage for Li is also given for comparison. (b) Stage 1 structure where the MoS2 sheets (rectangles with violet shading) are separated by one layer of Na (gold circles) giving a stoichiometric ratio of 1:1 between Mo and Na. (c) A ‘beyond’ stage 1 structure displaying 2 layers of Na between the MoS2 sheets yielding a capacity of 260 mAh∙g−1. The unit cell for each structure is highlighted with red line.
Condensedmatter 04 00053 g005
Figure 6. Site-projected density of states for various Na concentration for the 2H and 1T/1T’ phases.
Figure 6. Site-projected density of states for various Na concentration for the 2H and 1T/1T’ phases.
Condensedmatter 04 00053 g006

Share and Cite

MDPI and ACS Style

Lane, C.; Cao, D.; Li, H.; Jiao, Y.; Barbiellini, B.; Bansil, A.; Zhu, H. Understanding Phase Stability of Metallic 1T-MoS2 Anodes for Sodium-Ion Batteries. Condens. Matter 2019, 4, 53. https://doi.org/10.3390/condmat4020053

AMA Style

Lane C, Cao D, Li H, Jiao Y, Barbiellini B, Bansil A, Zhu H. Understanding Phase Stability of Metallic 1T-MoS2 Anodes for Sodium-Ion Batteries. Condensed Matter. 2019; 4(2):53. https://doi.org/10.3390/condmat4020053

Chicago/Turabian Style

Lane, Christopher, Daxian Cao, Hongyan Li, Yucong Jiao, Bernardo Barbiellini, Arun Bansil, and Hongli Zhu. 2019. "Understanding Phase Stability of Metallic 1T-MoS2 Anodes for Sodium-Ion Batteries" Condensed Matter 4, no. 2: 53. https://doi.org/10.3390/condmat4020053

Article Metrics

Back to TopTop