Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Fast Mg2+ diffusion in Mo3(PO4)3O for Mg batteries

Ziqin Rong a, Penghao Xiao b, Miao Liu b, Wenxuan Huang a, Daniel C. Hannah b, William Scullin c, Kristin A. Persson bd and Gerbrand Ceder *bd
aDepartment of Materials Science and Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139, USA
bMaterials Science Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA. E-mail: gceder@berkeley.edu; gceder@lbl.gov
cLeadership Computing Facility, Argonne National Laboratory, Argonne, IL 60439, USA
dDepartment of Materials Science and Engineering, University of California, Berkeley, Berkeley, CA 94720, USA

Received 14th April 2017 , Accepted 22nd June 2017

First published on 26th June 2017


Abstract

In this work, we identify a new potential Mg battery cathode structure Mo3(PO4)3O, which is predicted to exhibit ultra-fast Mg2+ diffusion and relatively high voltage based on first-principles density functional theory calculations. Nudged elastic band calculations reveal that the migration barrier of the percolation channel is only ∼80 meV, which is remarkably low, and comparable to the best Li-ion conductors. This low barrier is verified by ab initio molecular dynamics and kinetic Monte Carlo simulations. The voltage and specific energy are predicted to be ∼1.98 V and ∼173 W h kg−1, respectively. If confirmed by experiments, this material would have the highest known Mg mobility among inorganic compounds.


A promising strategy to increase the energy density of rechargeable batteries is to transition from monovalent to multivalent batteries, such as Mg batteries,1–3 in which Mg2+ is reversibly inserted into/extracted from the cathode. Because dendrites are less likely to grow on Mg metal than Li metal during charging, Mg metal can be directly used as the anode, resulting in a substantial improvement of the theoretical volumetric energy density from 760 mA h cm−3 for Li–graphite to 3833 mA h cm−3 for Mg metal.4

One of the scientific challenges for the development of Mg-ion batteries is the limited mobility of Mg2+ in cathode structures. Unlike its alkali cation cousins Li+ and Na+,5–7 the higher charge of Mg2+ makes it much harder to overcome the migration barrier. In addition, the difficulty of developing Mg anodes and electrolytes that are compatible with candidate cathode materials has posed a challenge for the experimental evaluation of the electrochemical performance of interesting cathode candidates.8 As a result, the search for Mg cathode structures has only yielded a few materials that exhibit reasonably fast reversible electrochemical Mg2+ intercalation, i.e., Chevrel Mo6S8 (∼135 mA h g−1 at ∼1.0–1.3 V),3 orthorhombic V2O5 (∼150 mA h g−1 at ∼2.3–2.6 V),5,9 MoO3 (∼220 mA h g−1 at ∼1.7–2.8)5 and spinel TiS2 (∼200 mA h g−1 at ∼1–1.5 V).10,11 However, the low voltages limit the power densities of sulfides. Oxides and phosphates generally have higher voltages but at the cost of poorer Mg2+ diffusivity.

In this work, we identify a new possible phosphate compound Mo3(PO4)3O, which is shown to exhibit ultra-fast Mg2+ diffusion and relatively high voltage based on first-principles density functional theory (DFT) calculations. First-principles calculations have proven to be accurate and effective in studying the voltage and mobility of Li-ion12–19 and multivalent electrode materials.6,7,20 Our first-principles nudged elastic band (NEB) calculations21–23 predict that Mo3(PO4)3O has an unusually low Mg migration barrier of ∼80 meV, which is lower than the values previously reported for spinel TiS2 (∼550 meV)10 and Chevrel phases (∼360 meV),11 suggesting that this structure may enable very high Mg2+ diffusivity.

The crystal structure of MgMo3(PO4)3O is shown in Fig. 1. MgMo3(PO4)3O was derived from known compounds such as CaFe3(PO4)3O,24 SrFe3(PO4)3O25 and Bi0.4Fe3(PO4)3O26 by substituting Fe by Mo and the other metal ion (Ca, Sr, Bi) by Mg. The structure is relaxed in both lattice parameters and atomic positions after direct substitutions. Apart from placing Mg in the Ca site in CaFe3(PO4)3O (site B in Fig. 1), we also investigated other possible sites for Mg to reside in the empty host structure Mo3(PO4)3O. One site (site A in Fig. 1) is 9.8 meV lower in energy than site B. As shown in Fig. 1, there are one A site and two B sites in the unit cell, and site A and site B are too close to accommodate Mg cations simultaneously. Hence our calculations predicted that Mg resides in site A at Mg0.5Mo3(PO4)3O composition and in site B in MgMo3(PO4)3O. Chains composed of edge-sharing MoO6 octahedra along the [010] direction form the backbone of the structure and are interconnected by predominantly corner-sharing MoO5 trigonal bipyramids, MoO4 tetrahedra, and PO4 tetrahedra. Except for the [Mo6O28] chains, there are only two edge-sharing links in the unit cell, which are between PO4 tetrahedra and MoO6 octahedra. All the other links between polyhedra are corner-sharing links, which enables the polyhedra to rotate slightly, creating an adaptive tunnel to facilitate Mg2+ diffusion. Similar to LiFePO4,27 the MgMo3(PO4)3O structure has a 1D diffusion channel along the b-axis, along which Mg2+ diffusion is expected to occur.


image file: c7cc02903a-f1.tif
Fig. 1 Crystal structure of MgMo3(PO4)3O in (a) bc and (b) ac planes. The structure is built on [Mo6O28] chains interconnected by predominantly corner-sharing polyhedra. Similar to LiFePO4, a 1D diffusion channel exists along the b-axis for Mg2+.

Fig. 2 shows the minimum energy paths for Mg2+ migration in Mo3(PO4)3O, as calculated using the NEB method. Mg2+ can follow two paths for migration from one stable site to the nearest equivalent site, an inner- or inter-channel path. The inner-channel path involves migration along the b-axis direction and across the unit cell boundary (the unit cell is shown in Fig. 1), with a very low activation barrier of ∼80 meV (Fig. 2(a1 and a2)). The inter-channel path involves migration along the c-axis direction with a much higher activation barrier of ∼1200 meV (Fig. 2(b1 and b2)). Because diffusivity scales as the inverse exponential of the activation barrier, migration along the inter-channel path is unlikely as at room temperature hopping along the c-direction should be ∼1018 times less frequent than along the inner-channel path (an increase of 60 meV in the migration barrier corresponds to a decrease of one order of magnitude in the diffusivity at room temperature).7 The percolation channel for Mg2+ intercalation is therefore along the inner-channel path. The inner-channel path is divided into two segments, as shown in Fig. 2(a1) and (a2), with migration lengths and activation energies of 5.32 and 7.00 Å and ∼80 and ∼70 meV, respectively. Multiple minima exist along both paths because the PO4 groups can easily rotate to accommodate the Mg at different positions. The flexibility of the PO4 groups is a result of the corner-sharing connection between polyhedrons in the structure.


image file: c7cc02903a-f2.tif
Fig. 2 The minimum energy paths for Mg2+ migration in Mo3(PO4)3O. (a1) Migration path and (a2) minimum energy path for inner-channel diffusion. (b1) Migration path and (b2) minimum energy path for inter-channel diffusion. A–D in (a2) are markers for the migration processes. A and B corresponds to site A and B in Fig. 1. The inner-channel diffusion in (a1) and (a2) is the percolation channel for Mg2+ intercalation.

Tables 1 and 2 display the calculated thermodynamic electrochemical properties of MgMo3(PO4)3O. The redox couple during charging and discharging is Mo3+/Mo4+. The average voltage calculated with the GGA(PBE)+U functional is 1.98 V, resulting in a specific energy of 173 W h kg−1. The voltage data and specific energy data were verified using SCAN31 and Heyd–Scuseria–Ernzerhof (HSE)32 functionals. An average voltage of 1.69 V was obtained using the HSE functional, which is generally recognized as the most reliable voltage assessment method.33 The voltage from the SCAN functional was close to that of HSE.

Table 1 Average voltages of MgMo3(PO4)3O calculated using different levels of theory
Functionals Average voltage (V) Specific energy (W h kg−1)
GGA+U (Mo U = 4.38 eV)28–30 1.98 173
SCAN31 1.52 133
HSE0632 1.69 148


Table 2 Thermodynamic electrochemical properties of MgMo3(PO4)3O
Mo3(PO4)3O energy above hull MgMo3(PO4)3O energy above hull Volume change in charging/discharging
42 meV per atom 36 meV per atom 2%
Volumetric capacity Energy density Gravimetric capacity
330 A h L−1 651 W h L−1 87 mA h g−1


Since MgMo3(PO4)3O is a hypothetical compound derived by substituting ions in other compounds, we also evaluate its relative phase stability, by constructing the Mg–Mo–P–O phase diagram using available compounds in the Materials Project Database.34 Both the charged and discharged structures are metastable, with a moderate energy above the energy hull of 42 and 36 meV per atom, respectively. These energies measure the driving force to decompose into other phases. A broad analysis of the known compounds in ICSD (Inorganic Crystal Structure Database) has indicated that this energy range for metastability is quite common, indicating the possibility that this compound may be synthesizable.35 The charged material is unstable against decomposition to MoO2 and Mo2P3O11, and the discharged material is metastable with respect to MoO2, Mo2P3O11, and Mg3(PO4)2. Based on the energy above the hull for experimentally synthesized Chevrel phases, i.e., Mo6S8 (67 meV per atom) and Mg2Mo6S8 (49 meV per atom), the stability of MgMo3(PO4)3O is considered reasonable for synthesis. Since MgMo3(PO4)3O is a compound that is derived by a double substitution from CaFe3(PO4)3O, and is metastable with the composition given, a carefully constructed route to synthesize it will be necessary. We suggest that one starts with a more stable [X]Mo3(PO4)3O compound, where X is another working ion such as Ca2+, Zn2+, Li+, etc. and then exchange X with Mg2+ at elevated temperature. Finally, the volume change during charging and discharging is very small (∼2%), which is highly favorable for the reliability of electrodes.

To understand the unusually high mobility of Mg2+ in this compound, and large difference in activation barriers between the inner- and inter-channel paths, we analyzed the coordination number for Mg2+ in every image of the two paths. Mg2+ almost always maintains 4-fold coordination along the inner-channel path, which differs greatly from its behavior along the inter-channel path, where Mg2+ experiences a coordination number change of 4 → 2 → 4. A larger coordination number change has been previously shown to lead to larger site energy differences along the migration path and ultimately to a larger activation barrier.7

Moreover, the corner-sharing connections of most of the polyhedra in the structure facilitate Mg2+ migration by enabling rotation to accommodate the presence of a local Mg ion. From this perspective, the inner-channel along the b-axis is more advantageous than the inter-channel along the c-axis because rotation of the MoO6 octahedra along the edge-sharing [010] [Mo6O28] chains is much easier around the b-axis than the c-axis. In addition, the void in the middle of the inter-channel path is too open, and moderate polyhedral rotations cannot mediate the coordination number decrease.

To verify the migration barriers obtained from zero-K NEB calculations, we performed ab initio molecular dynamics (AIMD) simulations.36,37 The mean square displacement (MSD) at 650 K is plotted in Fig. 3. The main contribution of the displacement is along the b-axis, which further confirms the 1D diffusion channel, as indicated by the inner-channel path. The diffusivity calculated using AIMD is 2.82 × 10−5 cm2 s−1.


image file: c7cc02903a-f3.tif
Fig. 3 MSD of Mg2+ at 650 K from AIMD simulations. The main contribution of the displacement is along b-axis, which shows a one-dimensional diffusion channel, as indicated by the inner-channel path in Fig. 2(a1) and (a2).

For comparison, a kinetic Monte Carlo (kMC)38,39 simulation was conducted based on the inner-channel path in Fig. 2(a2). Hopping rates between local minima were calculated using harmonic transition state theory:40

 
image file: c7cc02903a-t1.tif(1)
where N is the number of atoms; υi and υi* are the positive normal mode frequencies at the local minimum and transition state, respectively; and ΔE is the energy barrier, as determined from NEB calculations. The pre-factors image file: c7cc02903a-t2.tif and barriers between minima A, B, C, and D in Fig. 2(a2) are listed in Table 3. The other half of the path is symmetrically equivalent to ABCD.

Table 3 Summary of hopping processes
Process Calculated prefactor (THz) Barrier (meV)
A → B 1.59 70.4
B → A 0.58 60.6
B → C 0.46 66.5
C → B 2.75 16.5
C → D 8.18 29.5
D → C 2.24 46.3


The diffusivity at 650 K determined using kMC is 13.70 × 10−5 cm2 s−1. The fact that the diffusivities from AIMD and kMC are within one order of magnitude confirms that the effective barrier of diffusion is low. The room temperature (300 K) Mg diffusivity estimated using kMC is 4.68 × 10−5 cm2 s−1.

In summary, the key property for developing Mg battery cathode materials is the Mg2+ cation mobility in the host structure. In this work, we show that Mo3(PO4)3O exhibits extraordinarily fast Mg2+ cation mobility based on NEB (activation barrier ∼80 meV), AIMD (diffusivity ∼2.82 × 10−5 cm2 s−1 at 650 K), and kMC (diffusivity ∼13.70 × 10−5 cm2 s−1 at 650 K, ∼4.68 × 10−5 cm2 s−1 at 300 K) simulations. This is to our knowledge the lowest migration barrier ever predicted for Mg2+ in an oxide. Its voltage is slightly higher than previously reported sulfides based on GGA+U (1.98 V), SCAN (1.52 V), and HSE06 (1.69 V) calculations, but the capacity of 91 mA h g−1 is relatively low. Our systematic first-principles studies indicate that Mo3(PO4)3O may be a promising 1D cathode material for Mg batteries and is worthy of possible experimental investigation. In addition, the unusually high predicted mobility indicates that while Mg2+ diffusion generally is slow in inorganic compounds, there may be notable exceptions.

This work was supported by the Joint Center for Energy Storage Research (JCESR), an Energy Innovation Hub funded by the U.S. Department of Energy, Office of Science and Basic Energy Sciences (Subcontract 3F-31144). This research used resources of the Argonne Leadership Computing Facility, which is a DOE Office of Science User Facility supported under Contract DE-AC02-06CH11357.

Notes and references

  1. P. Canepa, S. Gautam, D. C. Hannah, R. Malik, M. Liu, K. G. Gallagher, K. A. Persson and G. Ceder, Chem. Rev., 2017, 117, 4287–4341 CrossRef CAS PubMed.
  2. J. Muldoon, C. B. Bucur and T. Gregory, Chem. Rev., 2014, 114, 11683–11720 CrossRef CAS PubMed.
  3. D. Aurbach, Z. Lu, A. Schechter, Y. Gofer, H. Gizbar, R. Turgeman, Y. Cohen, M. Moshkovich and E. Levi, Nature, 2000, 407, 724–727 CrossRef CAS PubMed.
  4. H. D. Yoo, I. Shterenberg, Y. Gofer, G. Gershinsky, N. Pour and D. Aurbach, Energy Environ. Sci., 2013, 6, 2265–2279 CAS.
  5. G. Gershinsky, H. D. Yoo, Y. Gofer and D. Aurbach, Langmuir, 2013, 29, 10964–10972 CrossRef CAS PubMed.
  6. M. Liu, Z. Rong, R. Malik, P. Canepa, A. Jain, G. Ceder and K. A. Persson, Energy Environ. Sci., 2015, 8, 964–974 CAS.
  7. Z. Rong, R. Malik, P. Canepa, G. Sai Gautam, M. Liu, A. Jain, K. Persson and G. Ceder, Chem. Mater., 2015, 27, 6016–6021 CrossRef CAS.
  8. P. Canepa, G. S. Gautam, R. Malik, S. Jayaraman, Z. Rong, K. R. Zavadil, K. Persson and G. Ceder, Chem. Mater., 2015, 27, 3317–3325 CrossRef CAS.
  9. G. Amatucci, F. Badway, A. Singhal, B. Beaudoin, G. Skandan, T. Bowmer, I. Plitz, N. Pereira, T. Chapman and R. Jaworski, J. Electrochem. Soc., 2001, 148, A940–A950 CrossRef CAS.
  10. X. Sun, P. Bonnick, V. Duffort, M. Liu, Z. Rong, K. A. Persson, G. Ceder and L. F. Nazar, Energy Environ. Sci., 2016, 9, 2273–2277 CAS.
  11. M. Liu, A. Jain, Z. Rong, X. Qu, P. Canepa, R. Malik, G. Ceder and K. A. Persson, Energy Environ. Sci., 2016, 9, 3201–3209 CAS.
  12. S. P. Ong, V. L. Chevrier, G. Hautier, A. Jain, C. Moore, S. Kim, X. Ma and G. Ceder, Energy Environ. Sci., 2011, 4, 3680–3688 CAS.
  13. Y. Mo, S. P. Ong and G. Ceder, Chem. Mater., 2014, 26, 5208–5214 CrossRef CAS.
  14. G. K. P. Dathar, D. Sheppard, K. J. Stevenson and G. Henkelman, Chem. Mater., 2011, 23, 4032–4037 CrossRef CAS.
  15. T. Song, H. Cheng, H. Choi, J.-H. Lee, H. Han, D. H. Lee, D. S. Yoo, M.-S. Kwon, J.-M. Choi and S. G. Doo, ACS Nano, 2011, 6, 303–309 CrossRef PubMed.
  16. K.-S. Park, P. Xiao, S.-Y. Kim, A. Dylla, Y.-M. Choi, G. Henkelman, K. J. Stevenson and J. B. Goodenough, Chem. Mater., 2012, 24, 3212–3218 CrossRef CAS.
  17. D. A. Tompsett and M. S. Islam, Chem. Mater., 2013, 25, 2515–2526 CrossRef CAS.
  18. A. Urban, D.-H. Seo and G. Ceder, npj Comput. Mater., 2016, 2, 16002 CrossRef CAS.
  19. A. Jain, Y. Shin and K. A. Persson, Nat. Rev. Mater., 2016, 1, 15004 CrossRef CAS.
  20. G. Sai Gautam, P. Canepa, A. Abdellahi, A. Urban, R. Malik and G. Ceder, Chem. Mater., 2015, 27, 3733–3742 CrossRef CAS.
  21. G. Mills and H. Jónsson, Phys. Rev. Lett., 1994, 72, 1124 CrossRef CAS PubMed.
  22. G. Mills, H. Jónsson and G. K. Schenter, Surf. Sci., 1995, 324, 305–337 CrossRef CAS.
  23. Z. Rong, D. Kitchaev, P. Canepa, W. Huang and G. Ceder, J. Chem. Phys., 2016, 145, 074112 CrossRef PubMed.
  24. M. Hidouri and M. Ben Amara, Acta Crystallogr., Sect. E: Struct. Rep. Online, 2009, 65, i66 CAS.
  25. V. Morozov, K. Pokholok, B. Lazoryak, A. Malakho, A. Lachgar, O. Lebedev and G. Van Tendeloo, J. Solid State Chem., 2003, 170, 411–417 CrossRef CAS.
  26. A. Benabad, K. Bakhous, F. Cherkaoui and E. M. Holt, Acta Crystallogr., Sect. C: Cryst. Struct. Commun., 2000, 56, 1292–1293 Search PubMed.
  27. R. Malik, D. Burch, M. Bazant and G. Ceder, Nano Lett., 2010, 10, 4123–4127 CrossRef CAS PubMed.
  28. Z. Wu and R. E. Cohen, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 73, 235116 CrossRef.
  29. A. Jain, G. Hautier, S. P. Ong, C. J. Moore, C. C. Fischer, K. A. Persson and G. Ceder, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 84, 045115 CrossRef.
  30. L. Wang, T. Maxisch and G. Ceder, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 73, 195107 CrossRef.
  31. D. A. Kitchaev, H. Peng, Y. Liu, J. Sun, J. P. Perdew and G. Ceder, Phys. Rev. B: Condens. Matter Mater. Phys., 2016, 93, 045132 CrossRef.
  32. J. Heyd, G. E. Scuseria and M. Ernzerhof, J. Chem. Phys., 2003, 118, 8207–8215 CrossRef CAS.
  33. G. Ceder, MRS Bull., 2010, 35, 693–701 CrossRef CAS.
  34. S. Ping Ong, L. Wang, B. Kang and G. Ceder, Chem. Mater., 2008, 20, 1798–1807 CrossRef.
  35. W. Sun, S. T. Dacek, S. P. Ong, G. Hautier, A. Jain, W. D. Richards, A. C. Gamst, K. A. Persson and G. Ceder, Sci. Adv., 2016, 2, e1600225 Search PubMed.
  36. G. Kresse and J. Hafner, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 47, 558 CrossRef CAS.
  37. G. Kresse and J. Hafner, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 48, 13115 CrossRef CAS.
  38. A. Chatterjee and D. G. Vlachos, J. Comput.-Aided Mater. Des., 2007, 14, 253–308 CrossRef.
  39. G. Henkelman and H. Jónsson, J. Chem. Phys., 2001, 115, 9657–9666 CrossRef CAS.
  40. A. F. Voter and J. D. Doll, J. Chem. Phys., 1984, 80, 5832–5838 CrossRef CAS.

Footnotes

Electronic supplementary information (ESI) available: Computational methods details for the results presented. See DOI: 10.1039/c7cc02903a
Ziqin Rong and Penghao Xiao contributed equally to this work.

This journal is © The Royal Society of Chemistry 2017