K3MoPO7: the first molybdenum phosphate with edge-sharing MoO6 octahedra and PO4 tetrahedra

Siyang Luoab, Lei Kangabc, Zheshuai Lin*ab and Chuangtian Chenab
aBeijing Center for Crystal Research and Development, Technical Institute of Physics and Chemistry, Chinese Academy of Sciences, Beijing 100190, China. E-mail: zslin@mail.ipc.ac.cn
bKey Laboratory of Functional Crystals and Laser Technology, Technical Institute of Physics and Chemistry, Chinese Academy of Sciences, Beijing 100190, China
cUniversity of the Chinese Academy of Sciences, Beijing 100049, People's Republic of China

Received 21st April 2014 , Accepted 11th June 2014

First published on 11th June 2014


Abstract

Herein we report the discovery of a novel molybdenum phosphate K3MoPO7, in which the basic building unit, [P2Mo2O14]6− ring cluster anions, contains unique edge-sharing MoO6 and PO4 polyhedra. The structural stability of K3MoPO7 is confirmed by both experimental and first-principles studies.


Corner-, edge- and face-sharing patterns are three basic polyhedra connection modes in inorganic chemistry.1 These three ways of linking and their combinations significantly enrich the structural chemistry and result in diverse polymorphisms. In fact, the linkage mainly depends on the ionic charge, ionic radii and coordination number.2 In inorganic oxides, the corner-sharing pattern is the most common way to be adopted by the MxOy (M is cation) polyhedra, such as in borates,3 silicates,4 aluminates5 and phosphates.6 The edge- or face-sharing patterns, on the other hand, are much less likely to exist in the above compounds. In general, both patterns could be obtained in extreme conditions only.7

The molybdenum phosphate system also has rich crystal chemistry with a large number of structures containing tunnels, cages and micropores.8 Till now hundreds of molybdenum phosphates have been synthesized, and many of them have large surface areas which can be used in ion exchange, catalysis, and separations.9 It is observed that in all the discovered molybdenum phosphates, the MoO6 octahedra always share their corners with PO4 tetrahedra to build up frameworks.8 According to Pauling's third and fourth rules,2 polyhedra with high charge cations and low coordination numbers would be more stable if they are connected by sharing vertices while less stable if sharing edges or faces, due to the cation–cation electrostatic repulsion between centroids. Thus, for MoO6 octahedra and PO4 tetrahedra they are very likely to share minimal vertices (i.e., corner-shared) to build frameworks since the Mo6+ and P5+ cations have quite high positive charges. The edge- or face-sharing of MoO6 octahedra and PO4 tetrahedra have not been found yet. In this work we synthesize the first molybdenum phosphate K3MoPO7 in which the MoO6 octahedra are sharing-edged with PO4 tetrahedra so form the unique [P2Mo2O14]6− anion rings. The structural stability is investigated by thermal measures and first-principles calculations.

K3MoPO7 was synthesized through solid-state reaction in stoichiometric ration with K2CO3, MoO3 and K2HPO4 as the starting materials. It can also be obtained from stoichiometric composition melt of above raw materials with slow cooling since it is a congruent compound. The proof will show in thermal characterization aspect below. Small transplant single crystals with light yellow color were obtained through spontaneous crystallization. The crystal structure was solved and refined on the basis of single-crystal data. The XRD patterns of as-synthesized samples show good agreement with the calculated one derived from the single crystal data.

K3MoPO7 crystallizes in a monoclinic system with the centrosymmetric space group C2/m. In an asymmetric unit, K, Mo, P and O occupy three, one, one, and five crystallographically unique positions, respectively. The structure of K3MoPO7 is featured by isolated [P2Mo2O14]6− anions, with K+ cations filled up the space around the anions (Fig. 1a). The [P2Mo2O14]6− anion is constructed by two MoO6 octahedra and two PO4 tetrahedra in which the Mo and P atoms are coplanar. These polyhedra alternately connect with each other via either corner-sharing or edge-sharing patterns to form an 8-membered polyanion ring with the symmetry of C2h point group (Fig. 1b). The simple [P2Mo2O14]6− fundamental building unit is first found in all inorganic compounds.


image file: c4ra03612c-f1.tif
Fig. 1 (a) Polyhedral view of the K3MoPO7 structure projected along the b axis. Pink octahedrons and cyan tetrahedrons stand for MoO6 anion groups and PO4 anion groups, respectively. (b) Ball-and-stick model for a [P2Mo2O14]6− anion. The K, Mo, P, O atoms are shown as gray, teal, pink and red, respectively.

It is worthy to note that the MoO6/PO4 edge-sharing configuration in the [P2Mo2O14]6− ring is unique and distinguished to any other microscopic pattern in molybdenum phosphates. Our comprehensive survey for all transition-metal phosphates in inorganic crystal structural database reveals that indeed the edge-sharing structure with MO6 octahedra is very rare as the M cations are d0 transition-metal (M = Ti4+, Nb5+, Mo6+, W6+, Ta5+, V5+). There is only one phosphovanadate, Li2VPO6, having the similar edge-sharing structure.10 On the other hand, for the non-d0 transition-metal MO6 octahedra (M = Co3+, Mn2+, Cr3+, V3+, V2+ etc.) they are relatively easy to form the edge-sharing structure with PO4 tetrahedra; tens of these compounds containing this connection pattern have been found. This result actually agrees with the Pauling's rules since the non-d0 transition metals have relatively lower charge which could reduce the electrostatic repulsion between Mn+ and P5+ cations. Therefore, the discovery of K3MoPO7 provides a very rare example that violates the common rules for structural stability, which has significant implications to explore the molybdenum phosphates with new structures and new functions.

In [P2Mo2O14]6− anion ring the Mo6+ cation of MoO6 octahedron is located in the distorted environments due to second-order Jahn–Teller (SOJT) effect.11 The centroid shifts toward a vertex of the octahedron (local C3 [111] direction), which results in the formation of three short (1.749(2)–1.754(2) Å) and three long (2.186(2)–2.353(3) Å) Mo–O bonds, as shown in Fig. 1b, and all the three short bonds are outwards from the [P2Mo2O14]6− anion ring. As a comparison, in the PO4 group P–O bond lengths are ranged from 1.52 to 1.57 Å, and this group almost keeps the regular tetrahedral shape. In K3MoPO7 all bond lengths are comparable to those in other reported molybdenum phosphates. The calculated total bond-valances sums on Mo and P are 5.82 and 4.85, respectively. These values match well with the expected oxidation states.12

It is interesting that each MoO6 octahedron in the [P2Mo2O14]6− cluster anion contains three free vertices. This is also very rarely observed in molybdates. In general, the formation of the structures with MO6 octahedra that contain three (or more) free vertices is very unfavorable due to the strong trans influence of the terminal M–O bonds (i.e., the Lipscomb restriction).13 In K3MoPO7 the ring-like [P2Mo2O14]6− groups are very different from all other microscopic structures in molybdenum phosphates where the MoO6 groups are usually connected to other groups as chain, layer or three-dimensional network. The zero-dimensional topological structure of the [P2Mo2O14]6− cluster anion results in more dangling oxygen ions existed in the MoO6 octahedra. Therefore, this simple group might be an ideal staring building unit to further polymerize with other groups for the construction of more complex solids, e.g., open-framework structures or polyoxometalates, for the applications of catalysis, luminescence, magnetism, medicine, etc.14

The stability of K3MoPO7 was experimentally demonstrated by the thermal behaviour measured with a differential scanning calorimetry (DSC). No obvious endo- or exothermic peak was observed until the melting point of the compound was reached near 490 °C (Fig. 2a). This demonstrates that the [P2Mo2O14]6−anionic structure keeps stable in ambient atmosphere from room temperature to the melting point. After melting, the cooled solid remains were characterized by XRD and were identified that it crystallized into K3MoPO7 (Fig. 2b). A few weak extra peaks were observed since the volatilization of melt and component deviation. The DSC and XRD results clearly reveal that this compound melts congruent.


image file: c4ra03612c-f2.tif
Fig. 2 (a) DSC curves of K3MoPO7. (b) X-ray powder diffraction patterns of K3MoPO7. The middle curve is powder of as-synthesized. The top one is powder after melting. The bottom curve is simulated XRD derived from the single crystal data.

To further investigate the mechanism of the structurally stability of the edge-sharing [PO4] and [MoO6] in the [P2Mo2O14] ring, the first-principles studies on the electronic density difference and Mulliken atomic/bond populations15 were performed16 by the plane-wave pseudopotential method implemented in the CASTEP package.17 The electronic density difference calculation produces a density difference field which shows the changes in the electron distribution as all chemical bonds are formed in the system. Fig. 3a and b exhibit the electronic density difference in the edge-sharing and corner-sharing MoO6/PO4 groups, respectively, which shows a different charge redistribution among the edge-sharing O(3) and corner-sharing O(2). In detail, the O(3) atoms obtain more electronic charges from the neighboring Mo atoms compared with the O(2) atoms, but the charge transfer from the neighboring P atoms to the O(3) is less than to the O(2) atoms. This indicates a necessary migration of electronic charges from P–O bond to Mo–O bond due to the modification of the chemical environment around the oxygen atoms as they are charged from corner-sharing to edge-sharing.


image file: c4ra03612c-f3.tif
Fig. 3 Contour plots of the electronic density difference on the planes formed by (a) Mo, P and edge-sharing O(3) atoms and (b) Mo, P and corner-sharing O(2) atoms. The off-plane O atoms are represented by small red balls.

The more quantitative results come from the Mulliken analysis; P–O(3) and P–O(2) bond populations are 0.59 and 0.74, respectively, while those on the Mo–O(3) and Mo–O(2) population is 0.28 and 0.23, respectively. Thus, the increase (decrease) of bond covalence between Mo (P) and edge-sharing O compared with those around the corner-sharing O account for the stability of the whole [Mo2P2O14] group. It should be noted, however, that the Mulliken analysis reveals that the effective charges on O(3) (−0.96) is less than that on O(2) (−1.03). This implies that indeed the edge-sharing oxygen atoms are less stable than the corner-sharing oxygen, in consistence with the Pauling's second rule (i.e., the oxygen anion having more charge tends to be more stable). We suggest that the zero-dimensional closed-ring configuration of the [Mo2P2O14] group would be beneficial to the formation of edge-sharing structure compared with the chain, layer or three-dimensional network.

Conclusions

A unique [P2Mo2O14]6− ring-shape anion containing edge-sharing MoO6 octahedra and PO4 tetrahedra has been discovered in a novel molybdenum phosphate K3MoPO7. The basic building unit with edge-sharing MoO6 and PO4 polyhedra was first found. K3MoPO7 is a congruent compound with melting point at about 490 °C. The first-principles calculations reveal that the unusual edge-sharing feature attributes to the migration of electronic charges from P–O bond to Mo–O bond. In other words, the edge-sharing Mo–O bond is of larger covalence than that of the corner-sharing Mo–O bond. Our studies presented in this work, therefore, would greatly prompt the development of molybdenum phosphates and have implications on the progress of structural chemistry.

Acknowledgements

This work was supported by the National Natural Science Foundation of China under Grant nos 11174297 and 91022036, and the National Basic Research Project of China (nos 2010CB630701 and 2011CB922204), and Foundation of the Director of Technical Institute of Physics and Chemistry, CAS.

Notes and references

  1. (a) U. Mueller, Inorg. Struc. Chem, Wiley, 1993 Search PubMed; (b) I. Lindqvist, O. Hassel, M. Webb and M. Rottenberg, Acta Chem. Scand., 1950, 4, 1066 CrossRef CAS; (c) A. F. Wells, Struct. Inorg. Chem, Oxord, 1962 Search PubMed; (d) D. Long, E. Burkholder and L. Cronin, Chem. Soc. Rev., 2007, 36, 105 RSC; (e) N. V. Izarova, M. T. Pope and U. Kortz, Angew. Chem., Int. Ed., 2012, 51, 9492 CrossRef CAS PubMed.
  2. L. Pauling, J. Am. Chem. Soc., 1929, 51, 1010 CrossRef CAS.
  3. (a) J. D. Grice, P. C. Burns and F. C. Hawthorne, Can. Mineral., 1999, 37, 731 CAS; (b) C. T. Chen, T. Sasaki, R. K. Li, Y. C. Wu, Z. S. Lin, Y. Mori, Z. G. Hu, J. Y. Wang, G. Aka and M. Yoshimura, Nonlinear Optical Borate Crystals, Wiley-VCH, 2012 Search PubMed; (c) P. C. Burns, J. D. Grice and F. C. Hawthorne, Can. Mineral., 1995, 33, 1131 CAS; (d) F. C. Hawthorne, P. C. Burns and J. D. Grice, Rev. Mineral., 1996, 33, 41 CAS; (e) X. Yan, S. Luo, Z. Lin, Y. Yue, X. Wang, L. Liu and C. Chen, J. Mater. Chem. C, 2013, 1, 3616 RSC; (f) X. Yan, S. Luo, Z. Lin, J. Yao, R. He, Y. Yue and C. Chen, Inorg. Chem., 2014, 53, 1952 CrossRef CAS PubMed; (g) C. Chen, S. Luo, X. Wang, G. Wang, X. Wen, H. Xu, X. Zhang and Z. Xu, J. Opt. Soc. Am. B, 2009, 26, 1519 CrossRef CAS; (h) S. C. Wang, N. Ye, W. Li and D. Zhao, J. Am. Chem. Soc., 2010, 132, 8779 CrossRef CAS PubMed.
  4. (a) C. Meade, R. J. Hemley and H. K. Mao, Phys. Rev. Lett., 1992, 63, 1387 CrossRef; (b) G. V. Gibbs, Am. Mineral., 1982, 67, 421 CAS; (c) M. Grimsditch, Phys. Rev. Lett., 1984, 52, 2379 CrossRef CAS; (d) L. Levien, C. T. Prewitt and D. J. Weidner, Am. Mineral., 1980, 65, 920 CAS; (e) R. M. Hazen, L. W. Finger, R. J. Hemley and H. K. Mao, Solid State Commun., 1989, 72, 507 CrossRef CAS.
  5. (a) C. Y. Chen, H. X. Li and M. E. Davis, Microporous Mater., 1993, 2, 17 CrossRef CAS; (b) Z. Luan, C. Cheng, W. Zhou and J. Klinowski, J. Phys. Chem., 1995, 99, 1018 CrossRef CAS; (c) Z. Luan, M. Hartmann, D. Zhao, W. Zhou and L. Kevan, Chem. Mater., 1999, 11, 1621 CrossRef CAS; (d) P. Mondal and J. W. Jeffery, Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem., 1975, 31, 689 CrossRef.
  6. (a) A. K. Cheetham, G. Ferey and T. Loiseau, Angew. Chem., Int. Ed., 1999, 38, 3268 CrossRef CAS; (b) M. Estermann, L. B. Mccusker, C. Baerlocher, A. Merrouche and H. Kesler, Nature, 1991, 352, 320 CrossRef CAS; (c) P. Y. Feng, X. H. Bu, S. H. Tolbert and G. D. Stucky, J. Am. Chem. Soc., 1997, 119, 2497 CrossRef CAS; (d) R. K. Brow, R. J. Kirkpatrick and G. L. Turner, J. Am. Chem. Soc., 1993, 76, 919 CAS.
  7. (a) S. Jin, G. Cai, W. Wang, M. He, S. Wang and X. Chen, Angew. Chem., Int. Ed., 2010, 49, 4967 CrossRef CAS PubMed; (b) H. Huppertz and B. von der Eltz, J. Am. Chem. Soc., 2002, 124, 9376 CrossRef CAS PubMed; (c) H. Emme and H. Huppertz, Z. Anorg. Allg. Chem., 2002, 628, 2165 CrossRef; (d) H. Emme and H. Huppertz, Chem.–Eur. J., 2003, 9, 3623 CrossRef CAS PubMed; (e) H. Huppertz and H. Emme, J. Phys.: Condens. Matter, 2004, 16, S1283 CrossRef CAS; (f) H. Emme and H. Huppertz, Acta Crystallogr., Sect. C: Cryst. Struct. Commun., 2005, 61, i29 Search PubMed; (g) H. Huppertz and W. Schnick, Chem.–Eur. J., 1997, 3, 249 CrossRef CAS PubMed.
  8. (a) R. C. Haushalter, K. G. Strohmaier and F. W. Lai, Science, 1989, 246, 1289 CAS; (b) G. Costentin, A. Leclaire, M. M. Borel, A. Grandin and B. Raveau, Rev. Inorg. Chem., 1993, 13, 77 CrossRef CAS; (c) R. C. Haushalter and L. A. Mundi, Chem. Mater., 1992, 4, 31 CrossRef CAS; (d) A. Leclaire, A. Guesdon, M. M. Borel, F. Berrah and B. Raveau, Solid State Sci., 2001, 3, 877 CrossRef CAS.
  9. (a) S. Natarajan and S. Mandal, Angew. Chem., Int. Ed., 2008, 47, 4798 CrossRef CAS PubMed; (b) R. Murugavel, A. Choudhury, M. G. Walawalkar, R. Pothiraja and C. N. R. Rao, Chem. Rev., 2008, 108, 3549 CrossRef CAS PubMed.
  10. V. C. Korthuis, R. D. Hoffmann, J. Huang and A. W. Sleight, J. Solid State Chem., 1993, 105, 294 CrossRef CAS.
  11. P. S. Halasyamani, Chem. Mater., 2004, 16, 3586 CrossRef CAS.
  12. I. D. Brown and D. Altermatt, Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem., 1985, 41, 244 CrossRef.
  13. M. T. Pope and A. Muller, Angew. Chem., Int. Ed., 1991, 30, 34 CrossRef.
  14. (a) C. L. Hill, Chem. Rev., 1998, 98, 1 CrossRef CAS PubMed; (b) P. Gouzerh and A. Proust, Chem. Rev., 1998, 98, 77 CrossRef CAS PubMed; (c) N. Mizuno and M. Misono, Chem. Rev., 1998, 98, 199 CrossRef CAS PubMed; (d) A. Muller, F. Peters, M. T. Pope and D. Gatteschi, Chem. Rev., 1998, 98, 239 CrossRef PubMed; (e) J. T. Rhule, C. L. Hill, D. A. Judd and R. F. Schinazi, Chem. Rev., 1998, 98, 327 CrossRef CAS PubMed; (f) T. Yamase, Chem. Rev., 1998, 98, 307 CrossRef CAS PubMed; (g) D. E. Katsoulis, Chem. Rev., 1998, 98, 359 CrossRef CAS PubMed; (h) A. M. Khenkin, L. Weiner, Y. Wang and R. Neumann, J. Am. Chem. Soc., 2001, 123, 8531 CrossRef CAS PubMed; (i) A. Dolbecq, E. Dumas, C. R. Mayer and P. Mialane, Chem. Rev., 2010, 110, 6009 CrossRef CAS PubMed.
  15. R. S. Mulliken, J. Chem. Phys., 1955, 23, 1833 CrossRef CAS PubMed.
  16. The local density approximation and ultrasoft pseudopotentials are adopted in these calculations. O 2s22p4, P 3s23p3, K 3s23p64s1 and Mo 4s24p64d55s1 are treated as valence electrons. The kinetic energy cutoff of 500 eV and Monkhorst–Pack 2 × 4 × 3 k-point meshes are used. The choice of these computational parameters is good enough to ensure the accuracy of present purpose.
  17. (a) M. C. Payne, M. P. Teter, D. C. Allan, T. A. Arias and T. A. J. D. Joannopoulos, Rev. Mod. Phys., 1992, 64, 1045 CrossRef CAS; (b) S. J. Clark, M. D. Segall, C. J. Pickard, P. J. Hasnip, M. J. Probert, K. Refson and M. C. Payne, Z. Kristallogr., 2005, 220, 567 CrossRef CAS.

Footnotes

CCDC 996800. For crystallographic data in CIF or other electronic format see DOI: 10.1039/c4ra03612c
Crystal data for K3MoPO7: M = 712.42 g mol−1, monoclinic, a = 13.990(3) Å, b = 5.8600(12) Å, c = 9.6157(19) Å, β = 111.68(3)°, Z = 2, V = 732.54(87) Å3, space group C2/m. 5379 reflections measured, 1545 independent reflections (Rint = 0.0261). The final R1 value was 0.029. The final wR(F2) value was 0.0776 (I > 2σ(I)). The final R1 value was 0.0311 (all data). The final wR (F2) value was 0.0790 (all data). The goodness of fit on F2 was 1.064.

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