Article pubs.acs.org/JPCC
Cite This: J. Phys. Chem. C 2019, 123, 15777−15786
Monolayer, Bilayer, and Heterostructure Arsenene as Potential Anode Materials for Magnesium-Ion Batteries: A First-Principles Study Xiao-Juan Ye,† Gui-Lin Zhu,† Jin Liu,† Chun-Sheng Liu,*,† and Xiao-Hong Yan†,‡ †
Downloaded via UNIV OF SOUTHERN INDIANA on July 22, 2019 at 20:12:30 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Key Laboratory of Radio Frequency and Micro-Nano Electronics of Jiangsu Province, College of Electronic and Optical Engineering, Nanjing University of Posts and Telecommunications, Nanjing 210023, China ‡ School of Material Science and Engineering, Jiangsu University, Zhenjiang 212013, China S Supporting Information *
ABSTRACT: Magnesium-ion batteries (MIBs) have emerged as an attractive candidate for high-performance energy storage devices because of the low-cost and dendrite-free Mg metal anodes. However, the passivation layers formed on Mg anodes result in the sluggish kinetics of Mg2+ ion diffusion. Herein, we report Mg insertion materials based on arsenene as alternative anodes to Mg metal. Our first-principles calculations reveal the following findings: (1) Mg can be adsorbed on monolayer (bilayer) arsenene and arsenene/graphene heterostructure with adsorption energies in the range of 0.82−2.48 eV, suggesting Mgadsorbed arsenene systems with good energetic stability. (2) Monolayer arsenene has a ∼3 times higher specific capacity (1429.41 mA h g−1) than the arsenene bilayer and arsenene/graphene heterostructure. Among them, the arsenene monolayer possesses the lowest average open-circuit voltage. (3) In comparison with bilayer arsenene, the arsenene monolayer and heterostructure exhibit low barriers (0.08−0.33 eV) for Mg diffusion, corresponding to a fast charge/discharge capability. (4) During the magnesiation process, the small volume changes ( hexagonal ring > pentagonal ring. However, the interaction of Mg−DV2 is weaker than that of Mg−SV. With large Mg adsorption energies, the vacancies would act as potential traps in the defective region. We next consider the Mg diffusion from the vacancy to the further site along the stable adsorption sites. For the SV defect, two pathways are studied as shown in Figure 3a,b, that is, path A: HS0 → VS1 → HS1 → VS2 → HS2 and path B: HS0 → VS1 → VS3 → HS3. During Mg migration from HS0 to VS1 site, the highest diffusion barrier (1.55 eV) is found, suggesting that Mg could be trapped in the As vacancy. From the VS1 to HS1 (VS3) site, the energy barrier decreases to 0.12 (0.93) eV for the migration of Mg atom. This means that the effect of the SV defect reduces rapidly for Mg to diffuse away from the vacancy. For the DV2 defect, we also investigate the process of Mg moving out from the vacancy site. As shown in Figure 3c,d, the energy barriers are 1.09 and 1.32 eV of HD0 → HD1 and HD0 → VD1, respectively, which are much smaller than that of HS0 → 15780
DOI: 10.1021/acs.jpcc.9b02399 J. Phys. Chem. C 2019, 123, 15777−15786
Article
The Journal of Physical Chemistry C
Figure 5. (a) Top view of the considered migration paths for Mg. (b) Energy profile of Mg diffusion on the outside surface (H1 → H1 and H1 → V1 → H1) and interlayer (V2 → V2).
Figure 6. (a) Top and side views of the representative adsorption sites on the A/G heterostructure. (b) Band structures of the A/G.
one possible stable adsorption site (V2). That is to say, there is only one possible direction for Mg diffusion within the intercalated region, namely, V2 → V2 (the zigzag direction). The computed energy barrier in the interlayer is about 0.56 eV, which is much higher than that on the monolayer arsenene. However, for Mg diffusion on the outside surface of bilayer arsenene, there are two possible diffusion directions, that is, H1 → V1 → H1 (the armchair direction) and H1 → H1 (the zigzag direction). The diffusion barrier (0.44 eV) along the armchair direction is about 2.7 times larger than that along the zigzag direction (0.16 eV). The Mg diffusivity along the zigzag direction is about 5 × 104 times faster than that along the armchair direction at room temperature, suggesting that the presence of the second arsenene layer enhances the anisotropic diffusion feature. 3.4. Adatom Adsorption and Diffusion on the Heterostructure. The pristine monolayer and bilayer arsenene possess band gaps with poor electrical conductivity. In addition, the Mg diffusion on arsenene has relatively high diffusion barriers, which can be attributed to the lone-pair electrons forming the electron accumulation region. Thus, we wonder whether arsenene could be integrated with other 2D materials to transfer its nonbonding electrons. The graphene is widely introduced to construct hybrid electrodes with excellent performances owing to its good electrical conductivity and mechanical resilience.38,39 Figure 6a presents the A/G van der Waals heterostructure composed of 6 × 6 graphene and 4 × 4 arsenene supercells. The lattice mismatch of about 3.4% is within the allowable range. Because carbon has a higher electronegativity than arsenic, partial lonepair electrons could be transferred from arsenene to
For the case of Mg adsorption on the outside surface, the stable adsorption sites are H1 and V1, corresponding to the adsorption energies of 1.91 and 1.80 eV, respectively. Obviously, the presence of the second arsenene significantly enhances the Mg binding strength. Moreover, the Mg atom embedded in the interlayer of bilayer arsenene is prone to stay at the top of the valley site (V2). Especially, its adsorption energy is as high as 2.48 eV. In a word, these results suggest that Mg atoms are more inclined to embed in the interlayer rather than adsorption on the outside surface during the magnesiation process. When the Mg atom is adsorbed on the surface or embedded in the interlayer of bilayer arsenene, it first transfers its 3s electrons to the substrate. Meanwhile, the substrate backdonates some electrons to the empty Mg 3p orbitals. As illustrated in Figure 4b, an overlapping within [−4, 0] eV between Mg 3s3p orbitals and As 4s4p orbitals suggests that the strong chemical bonds can be formed between the substrate and Mg. In particularly, the hybridizations between Mg 3s3p and As 4s4p states above the V2 site are more obvious than those on the H1 and V1 sites. Thus, the V2 site is the most stable adsorption position. In addition, because of the obvious charge transfer from Mg to the substrate, the Mgbilayer−arsenene system is changed from semiconductor to metal (Figure 4b), providing a good electrical conduction during the charge/discharge process. We investigate the diffusion barriers of Mg on the bilayer arsenene. There are two cases for Mg diffusion (Figure 5). First is the diffusion within the interlayer, and the second one is the diffusion on the outside surface. When the Mg atom is embedded in the interlayer of bilayer arsenene, there is only 15781
DOI: 10.1021/acs.jpcc.9b02399 J. Phys. Chem. C 2019, 123, 15777−15786
Article
The Journal of Physical Chemistry C
Figure 7. Partial DOS of (a) Mg/A/G and (b) A/Mg/G. The Fermi levels have been shifted to zero.
Figure 8. (a) Top view of the diffusion paths and (b) energy profiles of Mg diffusions on the A/G heterostructure.
graphene.40 The Hirshfeld charge population analysis suggests that arsenene donates about 0.10 e to the graphene. Importantly, the Dirac point of graphene is still well preserved in the band gap of the A/G heterostructure (Figure 6b), which can provide a good electrical conductivity. We next study the Mg adsorption properties of the A/G. Mg can be adsorbed at the arsenene side (Mg/A/G), graphene side (A/G/Mg), and interface (A/Mg/G). As shown in Figure 6a, several different adsorption sites of the Mg adatom are considered. Some of the characteristics are summarized in Table S2. First, the Mg/A/G system is energetically more favorable compared with the A/G/Mg (A/Mg/G) system. The most stable adsorption site for the Mg atom on Mg/A/G is the H3 site, which is similar to that on the pristine monolayer arsenene. Second, because of the synergistic effect, the presence of graphene significantly enhances the Mg binding strength over that of monolayer arsenene only. However, compared to Mg adsorption on the pristine graphene (0.32 eV), the heterostructure cannot evidently improve the bonding strength between graphene and Mg. Finally, the adsorption energies (1.41−1.60 eV) of Mg in the Mg/A/G and A/Mg/G systems are comparable with those in the Ti3C2 sheet-based material (∼1.40 eV) and borophene (∼2.0 eV).41,42 To clarify the adsorption mechanism, we further discuss the Hirshfeld charge analysis (Table S2) and projected DOS
(Figure 7). Although the Mg atom donates the least electrons (∼0.16 e) to the substrate in the Mg/A/G system, the heterostructure back-donates most of the electrons (∼0.20 e) to the empty Mg 3p orbitals. Indeed, the hybridization between Mg 3s3p and As 3p orbitals in the Mg/A/G system is stronger than that in the A/Mg/G. Thus, the Mg/A/G system is energetically most stable among the three adsorption cases. Because the Mg/A/G and A/Mg/G systems are energetically more stable than the A/G/Mg system, we only consider one Mg atom diffusion on the outside surface of arsenene and in the interlayer of A/G. When Mg diffusion on the outside surface of arsenene, two typical diffusion pathways (H3 → H3 and H3 → V3 → H3) between energetically favorable adsorption sites are set to calculate the corresponding energy barriers (Figure 8a). Obviously, the Mg diffusion barrier (0.15 eV) along the armchair direction (H3 → V3 → H3) is about 2.0 times larger than that along the zigzag direction (H3 → H3) (Figure 8b), indicating that Mg is relatively easy to diffuse along the zigzag direction. For the diffusion of Mg in the interlayer of A/G (H4 → H4 and H4 → V4 → H4), the energy barrier (0.17 eV) of Mg diffusion along the zigzag direction is comparable with that along the armchair direction (0.20 eV). The mobility of Mg in the interlayer (on the outside surface of arsenene) along the zigzag direction is estimated to 15782
DOI: 10.1021/acs.jpcc.9b02399 J. Phys. Chem. C 2019, 123, 15777−15786
Article
The Journal of Physical Chemistry C
Figure 9. Voltage profile for Mg adsorption on (a) monolayer arsenene, (b) bilayer arsenene, and (c) A/G. Variation of the lattice constant as a function of dopant concentration, x, in (d) monolayer, (e) bilayer, and (f) A/G. The large, medium, and small balls represent the Mg, As, and C atoms, respectively.
where EMgx−substrate and Esubstrate are the energies of the Mgx− substrate and substrate, respectively. EMg represents the energy of per atom of bulk Mg. The convex hulls of Ef are shown in Figure S4. The compounds located on the hull are thermodynamically stable, whereas those above the hull are metastable. The systems of MgxAs (x = 0.25, 0.75, 1.0, 2.0) and MgxAsC2.25 (x = 0.27, 0.78) all lie on the convex hull, indicating that the adsorption of one and two layers of Mg is energetically stable. During the magnesiation process, the formation of Mg clusters on these substrates could be nucleation centers for the dendrite growth. Therefore, the possibility of small Mg cluster formation on monolayer (bilayer) arsenene and heterostructure is explored. As shown in Figure S5, we study the stability of four Mg atoms dispersed uniformly or clustered on these substrates. Clearly, the isolated configurations are lower in energy than the clustered ones, indicating that the clustering of Mg atoms would not occur. According to these results, the specific capacity (C) is calculated by the following equation43,44
be about 3.0 (15.0) times faster than that in the armchair direction at room temperature. Owing to the transfer of lone-pair electrons from arsenene to graphene, the internal electric field could be formed in the interlayer (Figure S2), which will be beneficial to the Mg adsorption on the outside surface and within the interlayer of the A/G. Therefore, the diffusion barriers of Mg are lower than that on the pristine arsenene monolayer. Especially, the energy barrier of Mg diffusion on the outside of arsenene in the A/G is only 0.08 eV which is comparable to that of phosphorene (0.09 eV). Clearly, the synergistic effect not only enhances the Mg binding strength but also reduces the diffusion barrier. 3.5. Theoretical Specific Capacity and Voltage Profile. The ion-specific capacity is another important factor in evaluating the feasibility of electrode materials. As shown in Figure S3, in the case of Mg atoms adsorbed on monolayer (bilayer) arsenene and A/G heterostructure, the first and second layer of Mg are adsorbed above the most stable and metastable positions, respectively. Regrettably, for the third layer adsorption, the geometric configuration becomes unstable because of the substantial structural destruction. To further investigate the stability of the intermediate phases, the formation energies for Mg−substrate systems at each Mg concentration are calculated by the following equation Ef = (EMg −substrate − Esubstrate − xEMg )/(x + 1) x
C=
z × xmax × F Msubstrate
(4)
where z is the valence number (z = 2 for Mg), xmax is the highest Mg concentration, F is the Faraday constant, and Msubstrate is the atomic mass of the substrate. Therefore, the monolayer (bilayer) arsenene and A/G heterostructure can reach saturation at Mg2As (Mg0.75As) and Mg0.78AsC2.25
(3) 15783
DOI: 10.1021/acs.jpcc.9b02399 J. Phys. Chem. C 2019, 123, 15777−15786
Article
The Journal of Physical Chemistry C
4. CONCLUSIONS To conclude, we have explored the possibility of arsenenebased anode materials for MIBs based on DFT calculations. Our results show that Mg atoms can form stable adsorption on monolayer arsenene with low diffusion barriers (0.21−0.33 eV). The specific capacity (1429.41 mA h g−1) of monolayer arsenene is higher than those for most existing anode materials in MIBs. However, the introduction of SV and DV in monolayer arsenene can significantly enhance the Mg binding energy and thus degrade the performance of Mg diffusion capability. Therefore, it is essential to avoid the formation of defects in the experimental fabrication of arsenene. Compared with pristine arsenene monolayer, bilayer arsenene (536.60 mA h g−1) and A/G heterostructure (409.90 mA h g−1) have smaller theoretical specific capacities. Furthermore, during magnesiation, the small volume changes of monolayer, bilayer, and heterostructure arsenene are beneficial for a long cycle of charging/discharging. We hope these results may provide a useful reference for the rational design of potential anodes based on 2D arsenene-based materials.
configurations with specific capacities of 1429.41 (536.60) and 409.90 mA h g−1, respectively. Especially, Mg exhibits a higher specific capacity on monolayer arsenene than the theoretical capacities of phosphorene (865 mA h g−1),16 MnSb2S4 (879 mA h g−1),33 and Ti2C (687 mA h g−1).41 The open-circuit voltage (OCV) is also an important parameter for rechargeable batteries. The OCVs can be obtained by calculating the average voltage (U) within the concentration range of x1 < x < x243,44 U= E(Mg x −substrate) + (x 2 − x1)EMg − E(Mg x −substrate) 1
2
2(x 2 − x1)e (5)
where E(Mgx1−substrate) and E(Mgx2−substrate) are the total energies of the Mg-adsorbed substrates at the Mg concentrations of x1 and x2, respectively. EMg represents the total energy per atom in the bulk Mg crystal. As depicted in Figure 9a−c, when the concentration increases, the values of U show decrease tendency because of the enhanced repulsion forces among neighboring Mg ions. The average electrode potentials of Mg-monolayer, Mg-bilayer, and Mg-heterostructure systems by numerically averaging the potential profile are 0.83, 1.06, and 0.86 V, respectively, which are comparable with those for Mg−phosphorene (0.833 V) and Mg−NbSe2 (1.30 V) systems.16,45 In addition, the average electrode potentials are between those of commercial anode materials for LIBs, such as 0.11 V for graphite and 1.5 V for TiO2.46,47 It is well known that the significant volume expansion upon magnesiation usually impairs the practical use of electrode materials.48,49 Thus, the lattice constant change as a function of magnesiation would play an important role in assessing the structural integrity of the electrode material. Figure 9d−f presents the change of lattice constants, (l − l0)/l0, where l and l0 are lattice constants for the magnesiated substrate and pristine substrate, respectively. The lattice constants of a and b in the monolayer (bilayer) system are slightly shrinked by 5.61% (1.12%) and 5.66% (0.99%), respectively. For the heterostructure, the expansion of out-of-plane lattice constant (c) is 15.4%. Therefore, the maximum volume contraction and expansion of the monolayer (bilayer) arsenene and A/G heterostructure are about 10.9% (2.1%) and 15.4%, respectively. In contrast to the significant volume expansion of bulk Bi (100%) during Mg insertion/deinsertion processes,50 arsenene-based materials would prolong the cycle life of the battery, suggesting that nanostructured materials can improve the performances of electrodes for MIBs. Furthermore, we perform the ab initio molecular dynamics simulation based on a canonical ensemble to further confirm the thermal stability of Mg2As, Mg0.75As, and Mg0.78AsC2.25. The final snapshots of atomistic structures after 5.0 ps at the temperature of 300 K are presented in Figure S6. Clearly, these systems can thermally withstand without structure reconstructions and bonds broken. Because of the low toxicity of gray arsenic, arsenene is a promising material for next-generation electronic and optoelectronic applications.51,52 Recent synthesis of arsenene nanosheets53 may pave the way for the development of arsenene-based anode materials in MIBs.
■
ASSOCIATED CONTENT
S Supporting Information *
The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpcc.9b02399. Relaxed atomic structures of monolayer arsenene with vacancy defects; calculated formation energies for defective arsenene; calculated adsorption energies of Mg on A/G, electrons transfer from Mg to the substrate, back-donation of electrons from the substrate to the Mg 3p orbitals; the electrostatic potentials for the A/G heterostructure; the geometric structures of Mg adsorption on the monolayer, bilayer, and heterostructure; calculated convex hulls of Mg-adsorbed systems; four Mg atoms isolated or clustered on the substrates; and the molecular dynamics simulations of Mg adsorption (PDF)
■
AUTHOR INFORMATION
Corresponding Author
*E-mail:
[email protected]. ORCID
Xiao-Juan Ye: 0000-0002-1690-3293 Chun-Sheng Liu: 0000-0001-8856-9581 Notes
The authors declare no competing financial interest.
■
ACKNOWLEDGMENTS This work is supported by the Jiangsu Specially Appointed Professor Plan, the National Natural Science Foundation of China (grant no. 11704198), the Natural Science Foundation of Jiangsu Province (grant no. BK20150826), and the Scientific Research Foundation of Nanjing University of Posts and Telecommunications (grant nos. NY215035 and NY217038).
■
REFERENCES
(1) Armand, M.; Tarascon, J.-M. Building Better Batteries. Nature 2008, 451, 652−657. (2) Tang, Y.; Zhang, Y.; Li, W.; Ma, B.; Chen, X. Rational Material Design for Ultrafast Rechargeable Lithium-Ion Batteries. Chem. Soc. Rev. 2015, 44, 5926−5940. 15784
DOI: 10.1021/acs.jpcc.9b02399 J. Phys. Chem. C 2019, 123, 15777−15786
Article
The Journal of Physical Chemistry C
(23) Tian, W.; Zhang, S.; Huo, C.; Zhu, D.; Li, Q.; Wang, L.; Ren, X.; Xie, L.; Guo, S.; Chu, P. K.; Zeng, H.; Huo, K. Few-Layer Antimonene: Anisotropic Expansion and Reversible Crystalline-Phase Evolution Enable Large-Capacity and Long-Life Na-Ion Batteries. ACS Nano 2018, 12, 1887−1893. (24) Blöchl, P. E. Projector Augmented-Wave Method. Phys. Rev. B: Condens. Matter Mater. Phys. 1994, 50, 17953. (25) Heyd, J.; Scuseria, G. E.; Ernzerhof, M. Hybrid Functionals Based on a Screened Coulomb Potential. J. Chem. Phys. 2003, 118, 8207−8215. (26) Tkatchenko, A.; Scheffler, M. Accurate Molecular van der Waals Interactions from Ground-State Electron Density and FreeAtom Reference Data. Phys. Rev. Lett. 2009, 102, 073005. (27) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132, 154104. (28) Hamann, D. R.; Schlüter, M.; Chiang, C. Norm-Conserving Pseudopotentials. Phys. Rev. Lett. 1979, 43, 1494. (29) Govind, N.; Petersen, M.; Fitzgerald, G.; King-Smith, D.; Andzelm, J. A Generalized Synchronous Transit Method for Transition State Location. Comput. Mater. Sci. 2003, 28, 250−258. (30) Henkelman, G.; Jónsson, H. Improved Tangent Estimate in the Nudged Elastic Band Method for Finding Minimum Energy Paths and Saddle Points. J. Chem. Phys. 2000, 113, 9978−9985. (31) Sibari, A.; Marjaoui, A.; Lakhal, M.; Kerrami, Z.; Kara, A.; Benaissa, M.; Ennaoui, A.; Hamedoun, M.; Benyoussef, A.; Mounkachi, O. Phosphorene as a Promising Anode Material for (Li/Na/Mg)-Ion Batteries: A First-Principle Study. Sol. Energy Mater. Sol. Cells 2018, 180, 253−257. (32) Laidler, K. J. The Development of the Arrhenius Equation. J. Chem. Educ. 1984, 61, 494−498. (33) Zhang, Z.; Zhang, Y.; Li, Y.; Lin, J.; Truhlar, D. G.; Huang, S. MnSb2S4 Monolayer as an Anode Material for Metal-Ion Batteries. Chem. Mater. 2018, 30, 3208−3214. (34) Arsentev, M.; Missyul, A.; Petrov, A. V.; Hammouri, M. TiS3 Magnesium Battery Material: Atomic-Scale Study of Maximum Capacity and Structural Behavior. J. Phys. Chem. C 2017, 121, 15509−15515. (35) Sun, X.; Liu, Y.; Song, Z.; Li, Y.; Wang, W.; Lin, H.; Wang, L.; Li, Y. Structures, Mobility and Electronic Properties of Point Defects in Arsenene, Antimonene and An Antimony Arsenide Alloy. J. Mater. Chem. C 2017, 5, 4159−4166. (36) Nayak, P. K.; Yang, L.; Brehm, W.; Adelhelm, P. From LithiumIon to Sodium-Ion Batteries: Advantages, Challenges, and Surprises. Angew. Chem., Int. Ed. 2018, 57, 102−120. (37) Kundu, D.; Talaie, E.; Duffort, V.; Nazar, L. F. The Emerging Chemistry of Sodium Ion Batteries for Electrochemical Energy Storage. Angew. Chem., Int. Ed. 2015, 54, 3431−3448. (38) Yu, J.; Wu, J.; Wang, H.; Zhou, A.; Huang, C.; Bai, H.; Li, L. Metallic Fabrics as the Current Collector for High-Performance Graphene-Based Flexible Solid-State Supercapacitor. ACS Appl. Mater. Interfaces 2016, 8, 4724−4729. (39) Xie, X.; Ao, Z.; Su, D.; Zhang, J.; Wang, G. MoS2/Graphene Composite Anodes with Enhanced Performance for Sodium-Ion Batteries: The Role of the Two-Dimensional Heterointerface. Adv. Funct. Mater. 2015, 25, 1393−1403. (40) Xia, C.; Xue, B.; Wang, T.; Peng, Y.; Jia, Y. Interlayer Coupling Effects on Schottky Barrier in the Arsenene-Graphene van der Waals Heterostructures. Appl. Phys. Lett. 2015, 107, 193107. (41) Xie, Y.; Dall’Agnese, Y.; Naguib, M.; Gogotsi, Y.; Barsoum, M. W.; Zhuang, H. L.; Kent, P. R. C. Prediction and Characterization of MXene Nanosheet Anodes for Non-Lithium-Ion Batteries. ACS Nano 2014, 8, 9606−9615. (42) Mortazavi, B.; Dianat, A.; Rahaman, O.; Cuniberti, G.; Rabczuk, T. Borophene as An Anode Material for Ca, Mg, Na or Li Ion Storage: A First-Principle Study. J. Power Sources 2016, 329, 456− 461.
(3) Lin, D.; Liu, Y.; Cui, Y. Reviving the Lithium Metal Anode for High-Energy Batteries. Nat. Nanotechnol. 2017, 12, 194−206. (4) Vaalma, C.; Buchholz, D.; Weil, M.; Passerini, S. A Cost and Resource Analysis of Sodium-Ion Batteries. Nat. Rev. Mater. 2018, 3, 18013. (5) Kim, H.; Kim, H.; Ding, Z.; Lee, M. H.; Lim, K.; Yoon, G.; Kang, K. Recent Progress in Electrode Materials for Sodium-Ion Batteries. Adv. Energy Mater. 2016, 6, 1600943. (6) Hwang, J.-Y.; Myung, S.-T.; Sun, Y.-K. Sodium-Ion Batteries: Present and Future. Chem. Soc. Rev. 2017, 46, 3529−3614. (7) Kim, D.-M.; Jung, S. C.; Ha, S.; Kim, Y.; Park, Y.; Ryu, J. H.; Han, Y.-K.; Lee, K. T. Cointercalation of Mg2+ Ions into Graphite for Magnesium-Ion Batteries. Chem. Mater. 2018, 30, 3199−3203. (8) Saha, P.; Datta, M. K.; Velikokhatnyi, O. I.; Manivannan, A.; Alman, D.; Kumta, P. N. Rechargeable Magnesium Battery: Current Status and Key Challenges for the Future. Prog. Mater. Sci. 2014, 66, 1−86. (9) Peng, L.; Zhu, Y.; Chen, D.; Ruoff, R. S.; Yu, G. TwoDimensional Materials for Beyond-Lithium-Ion Batteries. Adv. Energy Mater. 2016, 6, 1600025. (10) Son, S.-B.; Gao, T.; Harvey, S. P.; Steirer, K. X.; Stokes, A.; Norman, A.; Wang, C.; Cresce, A.; Xu, K.; Ban, C. An Artificial Interphase Enables Reversible Magnesium Chemistry in Carbonate Electrolytes. Nat. Chem. 2018, 10, 532−539. (11) Song, J.; Sahadeo, E.; Noked, M.; Lee, S. B. Mapping the Challenges of Magnesium Battery. J. Phys. Chem. Lett. 2016, 7, 1736− 1749. (12) Asl, H. Y.; Fu, J.; Kumar, H.; Welborn, S. S.; Shenoy, V. B.; Detsi, E. In Situ Dealloying of Bulk Mg2Sn in Mg-Ion Half Cell as an Effective Route to Nanostructured Sn for High Performance Mg-Ion Battery Anodes. Chem. Mater. 2018, 30, 1815−1824. (13) Liu, Z.; Lee, J.; Xiang, G.; Glass, H. F. J.; Keyzer, E. N.; Dutton, S. E.; Grey, C. P. Insights into the electrochemical performances of Bi anodes for Mg ion batteries using 25Mg solid state NMR spectroscopy. Chem. Commun. 2017, 53, 743−746. (14) Liang, Y.; Jing, Y.; Gheytani, S.; Lee, K.-Y.; Liu, P.; Facchetti, A.; Yao, Y. Universal Quinone Electrodes for Long Cycle Life Aqueous Rechargeable Batteries. Nat. Mater. 2017, 16, 841−848. (15) Zhang, P.; Wang, F.; Yu, M.; Zhuang, X.; Feng, X. TwoDimensional Materials for Miniaturized Energy Storage Devices: From Individual Devices to Smart Integrated Systems. Chem. Soc. Rev. 2018, 47, 7426−7451. (16) Han, X.; Liu, C.; Sun, J.; Sendek, A. D.; Yang, W. Density Functional Theory Calculations for Evaluation of Phosphorene as a Potential Anode Material for Magnesium Batteries. RSC Adv. 2018, 8, 7196−7204. (17) Boukhvalov, D. W.; Rudenko, A. N.; Prishchenko, D. A.; Mazurenko, V. G.; Katsnelson, M. I. Chemical Modifications and Stability of Phosphorene with Impurities: A First Principles Study. Phys. Chem. Chem. Phys. 2015, 17, 15209−15217. (18) Zhang, S.; Guo, S.; Chen, Z.; Wang, Y.; Gao, H.; GómezHerrero, J.; Ares, P.; Zamora, F.; Zhu, Z.; Zeng, H. Recent Progress in 2D Group-VA Semiconductors: From Theory to Experiment. Chem. Soc. Rev. 2018, 47, 982−1021. (19) Zhang, S.; Yan, Z.; Li, Y.; Chen, Z.; Zeng, H. Atomically Thin Arsenene and Antimonene: Semimetal-Semiconductor and IndirectDirect Band-Gap Transitions. Angew. Chem. 2015, 127, 3155−3158. (20) Zhang, B.; Zhang, H.; Lin, J.; Cheng, X. First-Principle Study of Seven Allotropes of Arsenene and Antimonene: Thermodynamic, Electronic and Optical Properties. Phys. Chem. Chem. Phys. 2018, 20, 30257−30266. (21) Wang, Y.; Ding, Y. Electronic Structure and Carrier Mobilities of Arsenene and Antimonene Nanoribbons: A First-Principle Study. Nanoscale Res. Lett. 2015, 10, 254. (22) Li, Z.; Xu, W.; Yu, Y.; Du, H.; Zhen, K.; Wang, J.; Luo, L.; Qiu, H.; Yang, X. Monolayer Hexagonal Arsenene with Tunable Electronic Structures and Magnetic Properties via Impurity Doping. J. Mater. Chem. C 2016, 4, 362−370. 15785
DOI: 10.1021/acs.jpcc.9b02399 J. Phys. Chem. C 2019, 123, 15777−15786
Article
The Journal of Physical Chemistry C (43) Liu, J.; Liu, C.-S.; Ye, X.-J.; Yan, X.-H. Monolayer InP3 as a Reversible Anode Material for Ultrafast Charging Lithium- and Sodium-Ion Batteries: A Theoretical Study. J. Mater. Chem. A 2018, 6, 3634−3641. (44) Liu, C.-S.; Yang, X.-L.; Liu, J.; Ye, X.-J. Theoretical Prediction of Two-Dimensional SnP3 as a Promising Anode Material for Na-Ion Batteries. ACS Appl. Energy Mater. 2018, 1, 3850−3859. (45) Peng, C.; Lyu, H.; Wu, L.; Xiong, T.; Xiong, F.; Liu, Z.; An, Q.; Mai, L. Lithium-and Magnesium-Storage Mechanisms of Novel Hexagonal NbSe2. ACS Appl. Mater. Interfaces 2018, 10, 36988− 36995. (46) Fan, X.; Mao, J.; Zhu, Y.; Luo, C.; Suo, L.; Gao, T.; Han, F.; Liou, S.-C.; Wang, C. Superior Stable Self-Healing SnP3 Anode for Sodium-Ion Batteries. Adv. Energy Mater. 2015, 5, 1500174. (47) Yang, Z.; Choi, D.; Kerisit, S.; Rosso, K. M.; Wang, D.; Zhang, J.; Graff, G.; Liu, J. Nanostructures and Lithium Electrochemical Reactivity of Lithium Titanites and Titanium Oxides: A Review. J. Power Sources 2009, 192, 588−598. (48) Mao, M.; Gao, T.; Hou, S.; Wang, C. A Critical Review of Cathodes for Rechargeable Mg Batteries. Chem. Soc. Rev. 2018, 47, 8804−8841. (49) Hannah, D. C.; Sai Gautam, G.; Canepa, P.; Rong, Z.; Ceder, G. Magnesium Ion Mobility in Post-Spinels Accessible at Ambient Pressure. Chem. Commun. 2017, 53, 5171−5174. (50) Shao, Y.; Gu, M.; Li, X.; Nie, Z.; Zuo, P.; Li, G.; Liu, T.; Xiao, J.; Cheng, Y.; Wang, C.; Zhang, J.-G.; Liu, J. Highly Reversible Mg Insertion in Nanostructured Bi for Mg Ion Batteries. Nano Lett. 2014, 14, 255−260. (51) Wang, Y.; Ye, M.; Weng, M.; Li, J.; Zhang, X.; Zhang, H.; Guo, Y.; Pan, Y.; Xiao, L.; Liu, J.; Pan, F.; Lu, J. Electrical Contacts in Monolayer Arsenene Devices. ACS Appl. Mater. Interfaces 2017, 9, 29273−29284. (52) Hart, M.; Chen, J.; Michaelides, A.; Sella, A.; Shaffer, M. S. P.; Salzmann, C. G. One-Dimensional Arsenic Allotropes: Polymerization of Yellow Arsenic inside Single-Wall Carbon Nanotubes. Angew. Chem., Int. Ed. 2018, 57, 11649−11653. (53) Vishnoi, P.; Mazumder, M.; Pati, S. K.; R. Rao, C. N. Arsenene Nanosheets and Nanodots. New J. Chem. 2018, 42, 14091−14095.
15786
DOI: 10.1021/acs.jpcc.9b02399 J. Phys. Chem. C 2019, 123, 15777−15786