An Unprecedented M–O Cluster Constructed from Nanosized

Nov 10, 2016 - Two hybrid polyoxometalates constructed from Preyssler P 5 W 30 clusters and Schiff base exhibiting interesting third-order NLO propert...
3 downloads 9 Views 1MB Size
Communication pubs.acs.org/IC

An Unprecedented M−O Cluster Constructed from Nanosized {[C5NH5]9[H31MoV12O24CoII12(PO4)23(H2O)4]}2− Anions Exhibiting Interesting Nonlinear-Optical Properties Yayu Dong,† Gonghao Hu,† Hao Miao,† Xingxiang He,† Min Fang,‡ and Yan Xu*,† †

College of Chemistry and Chemical Engineering, State Key Laboratory of Materials-Oriented Chemical Engineering, Nanjing Tech University, Nanjing 210009, P. R. China ‡ Department of Chemistry, School of Chemistry and Materials Science, Nanjing Normal University, Nanjing 210023, P. R. China S Supporting Information *

in the self-assembly of novel metal oxide (M−O) nanoclusters with the desired properties and distinctive structures. Compared with purely inorganic materials, modification of the surface of the nanoclusters by organic ligands will bring about a variety of novel functionalities by virtue of the excellent synergetic effects between inorganic species and organic molecules. These well-designed clusters can not only produce charming topologies but also manifest specific properties.15 Therefore, it is vital to design novel nanosized clusters modified by organic ligands in order to explore their properties. In our previous work, we realized that conjugated organic ligands improve the delocalized electron effect of clusters, which may give rise to nonlinear-optical (NLO) responses.16 Furthermore, the ligands coordinated to metal ions are favorable for the formation and stability of the metal clusters. Meanwhile, isolated clusters could be obtained because terminal ligands increase the steric hindrance, preventing the connection between the clusters. Taking these factors into account, we introduce pyridine and imidazole ligands to serve as chemical modifiers. Herein, a novel high-nuclear nanosized cluster compound, [C5NH5]8[C3H5N2]2{[C5NH5]9 [H31Mo12O24Co12(PO4)23(H2O)4]}·12H2O (1), has been successfully isolated by modulating pH values. The structural building unit of compound 1 is an novel nanosized cluster with more than 400 atoms composed of 12 CoIIO6 or CoIIO5N and 12 MoVO6 octahedra. These clusters are bridged by 23 {PO4} tetrahedra, and the size of the cluster is approximately 3 nm. Meanwhile, the third-order NLO responses indicated that 1 is an excellent third-order NLO material. Dark-red block crystals of compound 1 were derived from the self-assembly of molybdenum trioxide, cobaltous acetate, and imidazole dissolved in water, ethyl alcohol, and pyridine, in which the pH was adjusted to 4.7 with phosphoric acid (see the Supporting Information for details). Single-crystal X-ray diffraction analysis (Table S1) reveals that compound 1 is constructed from a centrosymmetric {[C 5 NH 5 ] 9 [H31MoV12O24CoII12(PO4)23(H2O)4]}2− nanocluster, two protonated imidazoles, eight free pyridines, and 12 lattice water molecules. Two state that the 2− charge is counterbalanced by two protonated imidazoles. In order to facilitate a better appreciation of the fascinating structure and the size of the

ABSTRACT: A novel high-nuclear nanosized cluster modified by conjugated organic ligands (pyridine and imidazole), [C5NH5]8[C3H5N2]2{[C5NH5]9 [H31Mo12O24Co12(PO4)23(H2O)4]}·12H2O (1), has been successfully isolated under hydrothermal conditions and structurally characterized. Compound 1 consists of 12 CoII and 12 MoV ions linked by 23 {PO4} groups, exhibiting unprecedented nanosized ship-shaped clusters. The magnetic measurements reveal that compound 1 exhibits dominant antiferromagnetic interactions. Additionally, pyridine and imidazole ligands enhance the delocalized electron effects of clusters, and the third-order nonlinearoptical response of compound 1 is excellent.

T

he design and syntheses of nanosized clusters are of great interest because of their intriguing structural diversity and abundant optical,1 catalytic,2 magnetic,3 and medical4 applications associated with their size effects.5 The past 2 decades have witnessed much growth in the polyoxometalates (POMs) as versatile precursors for the self-assembly of high-nuclearity nanoclusters.6 Also, major interest has been aroused because such clusters represent a paradigm in the discovery of systems that can grow from the molecular scale to the nanoscale.7 Müller’s groups have pioneered routes to polymolybdates of Keplerate-type {Mo132} and giant wheel {Mo154}.8 They are the first parent species exhibiting nanometer-sized clusters and the starting point for the development of nanocluster chemistry.9 In this respect, it should be mentioned that the vast majority of nanoclusters are constructed of the following well-known building blocks with sizes of 2−6 nm: wheel- or nanoringshaped structures ({Mo176}, {Mo128Eu4}2, {Mo144}, and {Mo16TM16P26}),10 a hollow nanometer sphere (keplerates {Mo102}, {Mo72V30}, and {Mo72Fe30}),11 saddle-shaped architecture ({W200Co8}),12 and nanoring−nanosphere molecule ({Mo214V30}).13 These types of nanoclusters based on molybdenum oxide clusters have been evidenced as possessing aesthetically fantastic structures and special properties, which are significant for materials science.14 However, their syntheses involve complex self-assembly processes, which may be unpredictable, restricting the development of nanocluster chemistry. Therefore, it still remains a tremendous challenge © XXXX American Chemical Society

Received: September 8, 2016

A

DOI: 10.1021/acs.inorgchem.6b02160 Inorg. Chem. XXXX, XXX, XXX−XXX

Communication

Inorganic Chemistry cluster, the building block representations from different views are shown in Figures 1 and S1, and a microscopic image of the

Figure 1. Ball-and-stick representation of the nanocluster of compound 1. The H atoms and lattice water molecules are omitted for clarity.

Figure 2. Assembly process of a cluster. (a) {Mo4Co2} subunit; (b) {Mo6Co6} subunit; (c) {Mo2Co4} subunit; (d) nanoclusters of 1, showing the diameter of the nanocluster. Color code: PO4, pink; MoO6, bright green; CoO6 and CoO5N, turquoise.

crystal morphology of compound 1 is displayed in Figure S2. The nanocluster mainly consists of 12 CoII and 12 MoV metal ions, which can be divided into three kinds of building blocks: {Co4} tetramers and {Co2} and {Mo2} dimers. A total of 23 {PO4} groups play important roles in bridging and stabilizing the structure of compound 1. On this basis, the skeletal structure is anionic with the general formula {[C 5 NH 5 ] 9 [H 31 Mo V 12 O 24 Co II 12 (PO 4 ) 23 (H2O)4]}2−, in which all of the Mo and Co centers adopt six-coordinated {MoO6} and {CoO6−xNx} (x = 0 or 1) modes exhibiting octahedral structures, as shown in Figure 1. Moreover, the arrangement patterns of the 12 Mo and Co centers can be divided into six {Mo2} and two {Co2} dimers and two {Co4} tetramers (Figures S3−S5), which are linked together by phosphato groups. Bond valence sum17 calculations have been applied to confirm the valence of the metal and O atoms (Tables S5 and S6). The results indicate that all of the Mo and Co atoms are in the oxidation states of 5 +and 2+, respectively. The Mo−O bond distances fall into the range of 1.667(3)−2.5997(9) Å. The Co−O and Co−N distances are in the ranges of 1.860(6)−2.315(3) and 2.108(3)−2.174(6) Å, which are consistent with the reported compounds.18 The assembly process of the cluster is illustrated in Figure 2. Two {MoO6} octahedra interconnect to construct a {Mo2} dimer by sharing an edge, while two {Mo2} dimers join together with a μ3-O and six μ2-O atoms from three phosphato groups, which is common in the previously reported wheel compounds.19 Analogously, two {CoO5N} octahedra connect to each other to form a {Co2} dimer in the coplanar pattern. Nevertheless, a striking feature is that one {CoO6} and three {CoO5N} octahedra link with each other in an edge-sharing manner, resulting in a novel building block {Co4} tetramer. Specifically, four {PO4} groups bind {Mo2} dimers and a {Co2} dimer to generate a {Mo4Co2} subunit (Figure 2a). Similarly, {Mo2Co4} (Figure 2c) can be viewed as a fusion of a {Mo2} dimer and a {Co4} tetramer via {PO4} units. Notably, {PO4} groups (Figure S6), as crucial bridges or hinges, play significant roles in the formation of different building blocks. {Mo4Co2} tightly couple neighboring {Mo2Co4}, thus leading to a {Co6Mo6} fragment (Figure 2b). Two such {Co6Mo6} building blocks are tied by bridging {PO4} groups to generate the centrosymmetric nanoclusters. Intriguingly, Co ions are connected with pyridine serving as oars and four free pyridines

as guests taking on the board of the ship, as shown in Figure 2d. Additionally, the free pyridines and protonated imidazoles around the cluster are stabilized by hydrogen-bonding interactions (Table S4). Meanwhile, these ligands further bridge isolated clusters together to form a 3D supramolecular network (Figure S7). It is remarkable that compound 1 is entirely distinct from the reported molybdenum(v) cobaltophosphate ({Mo16Co16P26}),19,10e which is a wheel-shaped nanocluster. As depicted in Figures 2d and S8, the length and width of this cluster reach up to 3.05 and 1.5 nm, respectively, which are due to modification of the organic ligands. Intriguingly, we note that the proportion of the nanocluster is analogous to the real ship. In terms of size, it is larger than the reported nanospheres ({Mo132}7a and {Mo72Fe30}11c) and slightly smaller than nanorings ({Mo154}7b and {Mo176}10a). This size is similar to the size magnitude of nanoparticles, making compound 1 a possible candidate for nanomaterial applications. Nonlinear optics is an emerging research field of photonics.20,1 Hitherto, literature about nanoclusters predominantly focuses on magnetic interactions within transition metals, but reports on their NLO properties, especially thirdorder optical properties, are still sporadic. In compound 1, the conjugated organic ligands (pyridine and imidazole) connect the M−O cluster by covalent or ionic bonds to form the conjugated structural motifs, inducing a strong metal−ligand charge-transfer transition.21 Therefore, this may improve the delocalized electron effect of the M−O cluster and organic units, giving rise to a strong NLO response. Herein, the openaperture Z-scan method with a femtosecond laser pulse and a Ti:95 sapphire system was used to measure the two-photon absorption (TPA) values and TPA absorption coefficient (β) of compound 1. Figure 3 shows the open-aperture Z-scan curves of compound 1. The TPA absorption coefficient β and the molecular TPA cross section σ are calculated as 0.00263 cm/ GM and 1058 GM (1 GM = 10−50 cm4 photon−1), respectively (see the Supporting Information for details). To our knowledge, compound 1 has a bigger molecular TPA cross section σ than those reported for other molybdenum-based NLO materials,22,16 which indicates that the combination of POMs B

DOI: 10.1021/acs.inorgchem.6b02160 Inorg. Chem. XXXX, XXX, XXX−XXX

Communication

Inorganic Chemistry

CoII systems may be described as having an effective spin of 1/2 with large anisotropy.25 Thus, for simple systems, it is possible to fit the data at low temperatures.26 In summary, we have successfully isolated a novel highnuclear M−O cluster modified with pyridine and imidazole ligands. Compound 1 consists of 12 CoII and 12 MoV ions bridged by 23 PO4 groups. Meaningfully, it represents the first example of high-nuclear ship-shaped nanoclusters. Additionally, compound 1 has big TPA values; therefore, it has potential applications in nonlinear optics. The result is significant not only for producing the first nanosized M−O cluster but also for providing a new pathway to exploit POM-based nanoclusters with intriguing physical−optical properties. Further works are focused on the design of novel nanoclusters by means of introducing different conjugated organic ligands into POMs, in order to enhance the NLO properties and further effectively apply for a physical−optical field.

Figure 3. Open-aperture Z-scan data at 820 nm for compound 1 in water at 1.0 × 10−3 mol L−1.



with conjugated organic ligands (pyridine and imidazole) gives rise to a strong NLO response and may thus be a good method for making NLO materials. The variable-temperature magnetic susceptibility measurements were measured at 2000 Oe in the temperature range of 2−300 K and plotted in the form of χM and χMT versus T (Figure 4). Among the MoV-containing compounds, the MoV

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.6b02160. Synthesis, single-crystal structure analysis, characterization, supplementary structural figures, powder X-ray diffraction measurements, IR spectrum, TPA, crystallographic data, supplementary calculation details, and crystal refinements (PDF) Crystallographic data for compound 1 (CIF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the NSFC (Grant 21571103), the Major Natural Science Projects of the Jiangsu Higher Education Institution (Grant 16KJA150005), and the Qing Lan Project.

Figure 4. Temperature dependence of χMT (blue) and χM (black) for 1.



ions favor the formation of diamagnetic Mo−Mo pairs as a spin pairing of the d1 electrons in the MoV dimers with short Mo− Mo distances with an average value of 2.592 Å. Therefore, only the CoII ions are responsible for the magnetic property of compound 1.23,10e As shown in Figure 4, the effective magnetic moment (μB) calculated at 300 K is 18.09 μB. It is higher than that of 12 uncoupled high-spin CoII ions (14.22 μB) with S = 3 /2 assuming g = 2.12. With the lowering of the temperature, the effective magnetic moment gradually decreases to a minimum of 10.82 μB at 2 K. The results are reasonable for systems containing high-spin CoII ions, and the shape of the curve reveals the antiferromagnetic behavior of compound 1. A quantitative interpretation of the magnetic data is not straightforward because of the orbital degeneracy of octahedral CoII. The spin−orbit coupling of the CoII ions plays an important role in researching magnets.24 The degeneracy of the 4 T1g ground state of the octahedral CoII ion is removed by the action of the axial and rhombic distortions of the crystal field, as well as by spin−orbit coupling. In the low-temperature region,

REFERENCES

(1) Wang, C.; Zhang, T.; Lin, W. B. Rational synthesis of noncentrosymmetric metal organic frameworks for second-order nonlinear optics. Chem. Rev. 2012, 112, 1084−1104. (2) (a) Kamata, K.; Yonehara, K.; Sumida, Y.; Yamaguchi, K.; Hikichi, S.; Mizuno, N. Efficient epoxidation of olefins with ≥ 99% selectivity and use of hydrogen peroxide. Science 2003, 300, 964−966. (b) Yoon, M.; Srirambalaji, R.; Kim, K. Homochiral metal−organic frameworks for asymmetric heterogeneous catalysis. Chem. Rev. 2012, 112, 1196−1231. (c) Sun, C. Y.; Liu, S. X.; Liang, D. D.; Shao, K. Z.; Ren, Y. H.; Su, Z. M. Highly stable crystalline catalysts based on a microporous metal−organic framework and polyoxometalates. J. Am. Chem. Soc. 2009, 131, 1883−1888. (3) (a) Clemente-Juan, J. M.; Coronado, E.; Gaita-Arino, A. Magnetic polyoxometalates: from molecular magnetism to molecular spintronics and quantum computing. Chem. Soc. Rev. 2012, 41, 7464−7478. (b) Zheng, S. T.; Yang, G. Y. Recent advances in paramagnetic-TMsubstituted polyoxometalates (TM = Mn, Fe, Co, Ni, Cu). Chem. Soc. Rev. 2012, 41, 7623−7646. (4) (a) Long, D. L.; Tsunashima, R.; Cronin, L. Polyoxometalates: building blocks for functional nanoscale systems. Angew. Chem., Int. Ed. 2010, 49, 1736−1758. (b) Yamase, T. Anti-tumor, -viral, and -bacterial

C

DOI: 10.1021/acs.inorgchem.6b02160 Inorg. Chem. XXXX, XXX, XXX−XXX

Communication

Inorganic Chemistry activities of polyoxometalates for realizing an inorganic drug. J. Mater. Chem. 2005, 15, 4773−4782. (5) (a) Sessoli, R.; Gatteschi, D.; Caneschi, A.; Novak, M. A. Magnetic bistability in a metal-ion cluster. Nature 1993, 365, 141−143. (b) Alivisatos, A. P. Semiconductor clusters, nanocrystals, and quantum dots. Science 1996, 271, 933−937. (c) Huang, L.; Han, L. J.; Feng, W. J.; Zheng, L.; Zhang, Z. B.; Xu, Y.; Chen, Q.; Zhu, D. R.; Niu, S. Y. Two 3D coordination frameworks based on nanosized huge Ln26 (Ln = Dy and Gd) spherical clusters. Cryst. Growth Des. 2010, 10, 2548−2552. (6) (a) Liu, H.; Song, C. Y.; Huang, R. W.; Zhang, Y.; Xu, H.; Li, M. J.; Zang, S. Q.; Gao, G. G. Acid−Base-Triggered structural transformation of a polyoxometalate core inside a dodecahedranelike silver thiolate shell. Angew. Chem., Int. Ed. 2016, 55, 3699−3703. (b) Zhan, C.; Cameron, J. M.; Gao, J.; Purcell, J. W.; Long, D. L.; Cronin, L. Time-Resolved assembly of cluster-in-cluster {Ag12}-in{W76} polyoxometalates under supramolecular control. Angew. Chem., Int. Ed. 2014, 53, 10362−10366. (c) Müller, A.; Beckmann, E.; Bögge, H.; Schmidtmann, M.; Dress, A. Inorganic chemistry goes protein size: a Mo368 nano-hedgehog initiating nanochemistry by symmetry breaking. Angew. Chem., Int. Ed. 2002, 41, 1162−1167. (7) (a) Müller, A.; Roy, S. En route from the mystery of molybdenum blue via related manipulatable building blocks to aspects of materials science. Coord. Coord. Chem. Rev. 2003, 245, 153−166. (b) Liu, T. B.; Diemann, E.; Li, H. L.; Dress, A. W. M.; Müller, A. SelfAssembly in aqueous solution of wheel-shaped Mo154 oxide clusters into vesicles. Nature 2003, 426, 59−62. (8) (a) Müller, A.; Krickemeyer, E.; Bögge, H.; Schmidtmann, M.; Peters, F. Organizational forms of matter: an inorganic super fullerene and keplerate based on molybdenum oxide. Angew. Chem., Int. Ed. 1998, 37, 3359−3363. (b) Müller, A.; Krickemeyer, E.; Meyer, J.; Bö gge, H.; Peters, F.; Plass, W.; Diemann, E.; Dillinger, S.; Nonnenbruch, F.; Randerath, M.; Menke, C. [Mo154(NO)14O420(OH)28(H2O)70](25±5)−: a water-soluble big wheel with more than 700 atoms and a relative molecular mass of about 24000. Angew. Chem., Int. Ed. Engl. 1995, 34, 2122−2124. (9) (a) Müller, A.; Serain, C. Soluble molybdenum blues “des pudels kern. Acc. Chem. Res. 2000, 33, 2−10. (b) Müller, A.; Kögerler, P.; Kuhlmann, C. A variety of combinatorially linkable units as disposition:† from a giant icosahedral Keplerate to multi-functional metal−oxide based network structures. Chem. Commun. 1999, 1347− 1358. (10) (a) Müller, A.; Koop, M.; Bögge, H.; Schmidtmann, M.; Beugholt, C. Exchanged ligands on the surface of a giant cluster: [(MoO3)176(H2O)63(CH3OH)17Hn](32−n)−. Chem. Commun. 1998, 1501−1502. (b) Mü l ler, A.; Shah, S. Q. N.; Bö g ge, H.; Schmidtmann, M. Molecular growth from a Mo176 to a Mo248 cluster. Nature 1999, 397, 48−50. (c) Cronin, L.; Beugholt, C.; Krickemeyer, E.; Schmidtmann, M.; Bögge, H.; Kögerler, P.; Luong, T.; Müller, A. Molecular Symmetry Breakers” generating metal-oxide-based nanoobject fragments as synthons for complex structures: [{Mo128Eu4O388H10(H2O)81}2]20−, a giant-cluster dimer. Angew. Chem., Int. Ed. 2002, 41, 2805−2808. (d) Müller, A.; Krickemeyer, E.; Bögge, H.; Schmidtmann, M.; Peters, F.; Menke, C.; Meyer, J. An unusual polyoxomolybdate: giant wheels linked to chains. Angew. Chem., Int. Ed. Engl. 1997, 36, 484−486. (e) du Peloux, C.; Dolbecq, A.; Mialane, P.; Marrot, J.; Rivière, E.; Sécheresse, F. A new family of layered molybdenum(V) cobalto-phosphates built up of [H14(Mo16O32)Co16(PO4)24(H2O)20]10− wheels. Angew. Chem., Int. Ed. 2001, 40, 2455−2457. (11) (a) Müller, A.; Shah, S. Q. N.; Bögge, H.; Schmidtmann, M.; Kögerler, P.; Hauptfleisch, P.; Leiding, S.; Wittler, K. Thirty electrons “Trapped” in a spherical matrix: a molybdenum oxide-based nanostructured keplerate reduced by 36 electrons. Angew. Chem., Int. Ed. 2000, 39, 1614−1616. (b) Müller, A.; Todea, A. M.; van Slageren, J.; Dressel, M.; Boegge, H.; Schmidtmann, M.; Luban, M.; Engelhardt, L.; Rusu, M. Triangular geometrical and magnetic motifs uniquely linked on a spherical capsule surface. Angew. Chem., Int. Ed. 2005, 44, 3857−3861. (c) Müller, A.; Sarkar, S.; Shah, S. Q. N.; Bögge, H.;

Schmidtmann, M.; Sarkar, S.; Kö gerler, P.; Hauptfleisch, B.; Trautwein, A. X.; Schünemann, V. Archimedean synthesis and magic numbers: “Sizing” giant molybdenum-oxide-based molecular spheres of the keplerate type. Angew. Chem., Int. Ed. 1999, 38, 3238−3241. (12) de la Oliva, A. R.; Sans, V.; Miras, H. N.; Yan, J.; Zang, H. Y.; Richmond, C. J.; Long, D. L.; Cronin, L. Assembly of a gigantic polyoxometalate cluster {W200Co8O660} in a networked reactor system. Angew. Chem., Int. Ed. 2012, 51, 12759−12762. (13) Botar, B.; Kögerler, P.; Hill, C. L. A nanoring−nanosphere molecule, {Mo214V30}: pushing the boundaries of controllable inorganic structural organization at the molecular level. J. Am. Chem. Soc. 2006, 128, 5336−5337. (14) (a) Inokuma, Y.; Kawano, M.; Fujita, M. Crystalline molecular flasks. Nat. Chem. 2011, 3, 349−358. (b) Long, D. L.; Burkholder, E.; Cronin, L. Polyoxometalate clusters, nanostructures and materials: from self-assembly to designer materials and devices. Chem. Soc. Rev. 2007, 36, 105−121. (c) Müller, A.; Kögerler, P.; Dress, A. W. M. Giant metal-oxide-based spheres and their topology: from pentagonal building blocks to keplerates and unusual spin systems. Coord. Chem. Rev. 2001, 222, 193−218. (15) Fu, J.; Sun, H. X.; Xu, Y.; Wang, C. L.; Zhu, D. R.; Sun, Q.; Liu, H. K. [Ni(C17H20N4)4][H5PMoVI8VIV4O40(VIVO)2]·8H2O: confinement of heteropoly anions in Ni-containing rigid concave surfaces with high catalytic activity in the oxidation of styrene. CrystEngComm 2012, 14, 5148−5150. (16) (a) Zhang, Y.; Huang, L.; Miao, H.; Wan, H. X.; Mei, H.; Liu, Y.; Xu, Y. Hydrothermal Synthesis, Structure, and Optical Properties of Two nanosized Ln26@CO3 (Ln = Dy and Tb) Cluster-Based Lanthanide−Transition-Metal−Organic Frameworks (Ln MOFs). Chem. - Eur. J. 2015, 21, 3234−3241. (b) Dong, Y. Y.; Xu, X.; Zhou, G. P.; Miao, H.; Hu, G. H.; Xu, Y. Three novel polyoxoanionsupported compounds: confinement of polyoxoanions in Nicontaining rigid concave surfaces with enhanced NLO properties. Dalton Trans. 2015, 44, 18347−18353. (c) Miao, H.; Wan, H. X.; Liu, M.; Zhang, Y.; Xu, X.; Ju, W. W.; Zhu, D. R.; Xu, Y. Two novel organic−inorganic hybrid molybdenum(V) cobalt/nickel phosphate compounds based on isolated nanosized Mo/Co(Ni)/P cluster wheels. J. Mater. Chem. C 2014, 2, 6554−6560. (17) (a) Brown, I. D. Recent developments in the methods and applications of the bond valence model. Chem. Rev. 2009, 109, 6858− 6919. (b) Gabrielle, D.; Geophysical, L. How to recognize O2‑, OH−, and H2O in crystal: structures determined by X-rays. Am. Mineral. 1970, 55, 1003−1015. (c) Brown, I. D.; Altermatt, D. Bond-valence parameters obtained from a systematic analysis of the inorganic crystal structure database. Acta Crystallogr., Sect. B: Struct. Sci. 1985, 41, 244− 247. (d) Liu, W.; Thorp, H. H. Bond valence sum analysis of metalligand bond lengths in metalloenzymes and model complexes. 2. Refined distances and other enzymes. Inorg. Chem. 1993, 32, 4102− 4105. (18) (a) Liang, Z. J.; Zhang, D. D.; Ma, P. T.; Niu, J. Y.; Wang, J. P. A {Co4O4} cubane incorporated within a polyoxoniobate cluster. Chem. Eur. J. 2015, 21, 8380−8383. (b) Jiao, Y. Q.; Qin, C.; Wang, X. L.; Wang, C. G.; Sun, C. Y.; Wang, H. N.; Shao, K. Z.; Su, Z. M. Three cobalt(II)-linked {P8W48} network assemblies: syntheses, structures, and magnetic and photocatalysis properties. Chem. - Asian J. 2014, 9, 470−478. (c) Liu, H. S.; Gómez-García, C. J.; Peng, J.; Sha, J. Q.; Wang, L. X.; Yan, Y. C. A Co-monosubstituted Keggin polyoxometalate with an antenna ligand and three cobalt(II) chains as counterion. Inorg. Chim. Acta 2009, 362, 1957−1962. (19) (a) Zhang, Y. N.; Zhou, B. B.; Li, Y. G.; Su, Z. H.; Zhao, Z. F. A new molybdenum(V) nickel phosphate based on divacant [H 30 (Mo V 16 O 32 )Ni14 (PO 4 ) 26 O 2 (OH) 4 (H 2 O) 8 ] 12‑ wheel. Dalton Trans. 2009, 43, 9446−9451. (b) Ma, Y.; Li, Y. G.; Wang, E. B.; Lu, Y.; Xu, X.; Bai, X. L. Synthesis, characterization and crystal structure of a new cobalt phosphomolybdate that contains [(Mo 16 O 32 )Co16(H2O)18(PO4)6(HPO4)18] wheels. Transition Met. Chem. 2006, 31, 262−267. (c) Yu, K.; Zhou, B. B.; Yu, Y.; Su, Z. H.; Yang, G. Y. A new organic-inorganic hybrid layered molybdenum(V) cobalt phosphate constructed from [H24(Mo16O32)D

DOI: 10.1021/acs.inorgchem.6b02160 Inorg. Chem. XXXX, XXX, XXX−XXX

Communication

Inorganic Chemistry Co16(PO4)24(OH)4(C5H4N)2(H2O)6]4‑ wheels and 4,4′-bipyridine linkers. Inorg. Chem. 2011, 50, 1862−1867. (d) duPeloux, C.; Dolbecq, A.; Mialane, P.; Marrot, J.; Rivière, E.; Sécheresse, F. A new two-dimensional molybdenum(V) nickel phosphate built up of [H18(Mo16O32)Ni16(PO4)26(OH)6(H2O)8]18‑ wheels. Inorg. Chem. 2002, 41, 7100−7104. (20) (a) Liang, K.; Zheng, H. G.; Song, Y. L.; Lappert, M. F.; Li, Y. Z.; Xin, X. Q.; Huang, Z. X.; Chen, J. T.; Lu, S. F. Self-assembly of interpenetrating coordination nets formed from interpenetrating cationic and anionic three-dimensional diamondoid cluster coordination polymers. Angew. Chem., Int. Ed. 2004, 43, 5776−5779. (b) Zhou, Y. S.; Qu, N. N.; Wang, X.; Zhang, L. J.; Shi, Z. H.; Hassan, ul S. Third-order optical nonlinearities of a series of compounds derived from a keplerate type molybdenum-oxide-based polyoxometalate. Chem. Res. Chin. Univ. 2014, 30, 715−719. (21) Das, S.; Nag, A.; Goswami, D.; Bharadwaj, P. K. Zinc(II)- and copper(I)-mediated large two-photon absorption cross sections in a bis-cinnamaldiminato schiff base. J. Am. Chem. Soc. 2006, 128, 402− 403. (b) Kamada, K.; Hara, C.; Ogawa, K.; Ohta, K.; Kobuke, Y. Strong two-photon absorption and its saturation of a self-organized dimer of an ethynylene-linked porphyrin tandem. Chem. Commun. 2012, 48, 7988−7990. (22) Xie, Y. M.; Zhang, Q. S.; Zhao, Z. G.; Wu, X. Y.; Chen, S. C.; Lu, C. Z. New optical supramolecular compound constructed from a polyoxometalate cluster and an organic substrate. Inorg. Chem. 2008, 47, 8086−8090. (23) (a) Vergamini, P. J.; Vahrenkamp, H.; Dahl, L. F. Synthesis, structure, and bonding of [M O3 (h5 -C5 H5)3 S4 ] +[Sn(CH3) 3Cl2] − containing a triangular molybdenum(IV) cluster cation with doubly and triply bridging sulfur atoms. J. Am. Chem. Soc. 1971, 93, 6327− 6329. (b) Haushalter, R. C.; Mundi, L. A. Reduced molybdenum phosphates: octahedral-tetrahedral framework solids with tunnels, cages, and micropores. Chem. Mater. 1992, 4, 31−48. (24) Mabbs, F. E.; Martin, D. J.Magnetism and transition metal complexes; Chapman and Hall: London, 1973. (25) (a) Lan, Q.; Tan, H. Q.; Liu, D.; Wang, E. B. J. Solid State Chem. 2013, 199, 129−133. (b) Caneschi, A.; Dei, A.; Gatteschi, D.; Tangoulis, V. Antiferromagnetic coupling in a six-coordinate high spin cobalt(II)-semiquinonato complex. Inorg. Chem. 2002, 41, 3508−3512. (c) Bu, X. H.; Tong, M. L.; Xie, Y. B.; Li, J. R.; Chang, H. C.; Kitagawa, S.; Ribas, J. Synthesis, structures, and magnetic properties of the copper(II), cobalt(II), and manganese(II) complexes with 9acridinecarboxylate and 4-quinolinecarboxylate ligands. Inorg. Chem. 2005, 44, 9837−9846. (26) Andres, H.; Clemente-Juan, J. M.; Aebersold, M.; Güdel, H. U.; Coronado, E.; Büttner, H.; Kearly, G.; Melero, J.; Burriel, R. Magnetic excitations in polyoxometalate clusters observed by inelastic neutron scattering: evidence for anisotropic ferromagnetic exchange interactions in the tetrameric cobalt(II) cluster [Co4(H2O)2(PW9O34)2]10‑. comparison with the magnetic and specific heat properties. J. Am. Chem. Soc. 1999, 121, 10028−10034.

E

DOI: 10.1021/acs.inorgchem.6b02160 Inorg. Chem. XXXX, XXX, XXX−XXX