Substituted Keggin Type Phosphotungstates - ACS Publications

Aug 10, 2016 - pot reaction procedure on reacting the dilacunary K14[P2W19O69- ...... EMR-II], and the University of Delhi for financial support of th...
0 downloads 0 Views 2MB Size
Article pubs.acs.org/IC

Lanthanoid Template Isolation of the α‑1,5 Isomer of Dicobalt(II)-Substituted Keggin Type Phosphotungstates: Syntheses, Characterization, and Magnetic Properties Rakesh Gupta,† Firasat Hussain,*,† Masahiro Sadakane,‡ Chisato Kato,§ Katsuya Inoue,§ and Sadafumi Nishihara§ †

Department of Chemistry, University of Delhi, North Campus, New Delhi, Delhi 110007, India Department of Applied Chemistry, Graduate School of Engineering, Hiroshima University, 1-4-1 Kagamiyama, Higashi, Hiroshima 732-8527, Japan § Graduate School of Science & Center for Chiral Science, Hiroshima University, 1-3-1 Kagamiyama, Higashi, Hiroshima 732-8526, Japan ‡

S Supporting Information *

ABSTRACT: A new series of heterometallic 3d−4f sandwich type phosphotungstates, [Ln{PCo2W10O38(H2O)2}2]11− (Ln = SmIII, EuIII, GdIII, TbIII, DyIII, HoIII, ErIII, TmIII, YbIII, and LuIII, denoted 1a−10a, respectively), have been synthesized by a onepot reaction procedure on reacting the dilacunary K14[P2W19O69(H2O)]·24H2O precursor with Ln(NO3)3·nH2O and Co(NO3)2· 6H2O in an aqueous potassium chloride solution. All the compounds were isolated as potassium salts and further characterized with different analytical techniques such as single-crystal X-ray diffraction, Fourier transform infrared spectroscopy, highresolution electrospray ionization mass spectrometry, elemental analysis by inductively coupled plasma atomic emission spectroscopy, magnetic measurement, and thermogravimetric analysis. Single-crystal X-ray diffraction analysis of the compounds reveals that all these compounds are isostructural and crystallized in the orthorhombic crystal system in space group Iba2. The polyanions contain the α-1,5 isomer of dicobalt-substituted α-Keggin phosphotungstate, which sandwiched lanthanoid cation and formed novel heterometallic dimer species. The temperaturedependent magnetic susceptibilities of 1a, 2a, 4a, and 7a−10a indicate the dominant contribution of the ferromagnetic interaction between CoII and CoII within the cluster, while the antiferromagnetic interaction between CoII and LnIII dominates for 3a, 5a, and 6a. The isothermal magnetizations of 1a−10a show a gradual increase in magnetization at low fields and do not reach saturation even at 50 kOe.



INTRODUCTION Polyoxometalates (POMs) make up a class of oxo-bridged anionic metal clusters that exhibit great structural variety. The heteropolyoxometalates are an important class of POMs, which consist of early transition metals in their high oxidation states combined with hetero group XOn (X = PV, SiIV, GeIV, AsIII, SbIII, etc.). The level of scientific interest in POMs has increased worldwide because of their remarkable unexpected applications in the fields of catalysis, biomedicine, molecular magnets, imaging techniques, and materials science.1 POMs are labile in aqueous solution and can also isomerize into other POM compounds depending upon the reaction conditions. Lacunary POMs are synthesized from the parent ions by the loss of one or more MO6 octahedra, and these lacunary species are more reactive toward metal ions (transition metals and rare earth cations) than parent ions, forming an enormous number of discrete nanomolecular clusters. Polyoxotungstates form the largest subclass of POM compounds in which Keggin and © XXXX American Chemical Society

Wells-Dawson type polyanions are widely explored. So far, a large number of transition metal-substituted POM compounds have been synthesized, and these compounds form a separate class of POM chemistry. Since 1971,2 several research groups have incorporated rare earth cations into the lacunary POM species, and the level of research interest in the lanthanide cations has increased because of their oxophilic nature, multiple coordination requirements with addition of Lewis acid catalysis, magnetic and luminescence properties.3 We have reacted lanthanoid salts with different lacunary POMs, forming novel molecular clusters with different properties.4 Today, POMs are treated with a combination of transition metals (TMs) (3d or 4d) and lanthanoid cations for the development of a unique family of polyoxometalates, which are known as heterometallic POMs. The chemistry of such mixed-metal Received: December 3, 2015

A

DOI: 10.1021/acs.inorgchem.5b02772 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

1:1:3 molar ratio in an aqueous potassium chloride solution at room temperature. The molecular complexes were isolated as potassium salts and used for further analysis. All molecular nanoclusters were structurally characterized by single-crystal X-ray diffraction (XRD), Fourier transform infrared (FT-IR) spectroscopy, magnetism, thermogravimetric analysis (TGA), electrospray ionization mass spectrometry (ESI-MS), and elemental analysis.

polyoxometalates is still incipient. 3d−4f or 4d−4f heterometallic compounds are more interesting than the 3d- or 4f-substituted POM compounds because of the application of the former in molecular magnets, biometallic catalysis, and molecular absorption.5 Moreover, development of this class combines the properties of 3d and 4f compounds and is highly important in the field of catalysis. Nevertheless, it is comparatively difficult to find suitable reaction conditions for synthesizing 3d−4f heterometallic derivatives, as 4f cations show more oxophilic reactivity with POMs. On the other hand, interactions between TM cations and polyoxoanions are relatively weak; thus, under the same reaction conditions, oxophilic 4f cations and relatively less reactive 3d metal cations can create competition toward highly negative polyoxoanions, hence leading to an amorphous precipitate instead of a crystalline material.6 Therefore, it is very difficult and challenging to develop suitable synthetic conditions for synthesizing heterometallic compounds based on lacunary POMs. The recent reviews show that isolation of the 3d−4f heterometallic class of POMs remains unexplored.7 The few examples of heterometallic species are reported in which 3d and 4f metals are an intrinsic part of the POM skeleton.8 In most of the 3d−4f heterometallic compounds, 4f metal-substituted POMs are decorated with 3d-metal organic complexes or vice versa.9 Merca et al. reported GdIII-substituted Hervé-like tungstoarsenite [((VO)2Gd(H2O)4K2(H2O)2Na(H2O)2)(α-BAsW9O33)2]8−, which was composed of two {α-B-AsW9O33} Keggin subunits through belts of two square pyramidal (VO)2+ groups and square antiprismatic Ln3+ ion.10 Wang et al. described a series of heterometallic nanoclusters, [{LnIIIMnIII4(μ3-O)2(μ2OH)2(H2O)(CO3)}(β-SiW8O31)2]n− (Ln = SmIII, GdIII, HoIII, ErIII, TmIII, and YbIII, n = 13; Ln = CeIV, n = 12), in which appended cubane {LnMnIII4} was sandwiched between two (β-SiW8O31) units.11 Recently, they also synthesized sandwich type tungstoantimonite {K2Dy2Cu2(H2O)8(SbW9O33)2}6− in which a ringlike 3d−4f−4p cluster {K2Dy2Cu2(H2O)8}12+ was sandwiched by two (SbW9O33) units.12 Ibrahim et al. reported the largest discrete heterometallic polyoxotungstate, [Dy30Co8Ge12W108O408(OH)42(OH2)30]56−, by using an [A-αGeW9O34] precursor, which shows single-molecule magnet (SMM) behavior.13 Reinoso et al. isolated a series of nine heterometallic [Sb7W36O133Ln3M2(OAc)(H2O)8]17− anions (Ln3M2; Ln = LaIII, CeIII, PrIII, NdIII, SmIII, EuIII, and GdIII, M = CoII; Ln = CeIII, M = NiII and ZnII) using 3d-disubstituted Krebs type tungstoantimonates(III).14 Recently, Mizuno et al. synthesized five sandwich type POMs of 3d−3d′−4f heptanuclear clusters [FeM4{Ln(L)2}2O2(A-α-SiW9O34)2] [M = MnIII and CuII; Ln = GdIII, DyIII, and LuIII, where L = acac (acetylacetonate) and hfac (hexafluoroacetylacetonate)] and also modulated the magnetic interactions between Mn3+ - Mn3+ and Cu2+ - Cu2+ ions.15 Very recently, the Yang group described a series of organic−inorganic hybrid heteropolyoxometalates, [Fe2Ln(β-PW10O37)2(Tart)]9− (Ln = LaIII, CeIII, SmIII, and TbIII), in which Fe3+ and Ln3+ ions were situated in the cavity of divacant [β-PW10O37]9− with the help of an organic tartaric ligand.16 We studied the reaction of 3d and 4f mixed metal cations with lacunary POM precursors in the absence of an organic moiety under mild reaction conditions. We successfully obtained crystalline materials of the series of 3d−4f sandwich type phosphotungstates [Ln{PCo2W10O38(H2O)2}2]11− (Ln = SmIII, EuIII, GdIII, TbIII, DyIII, HoIII, ErIII, TmIII, YbIII, and LuIII) denoted as 1a−10a, respectively. All these complexes were synthesized by a one-pot reaction of dilacunary K14[P2W19O69(H2O)]·24H2O, Ln(NO3)3·nH2O, and Co(NO3)2·6H2O in a



EXPERIMENTAL SECTION

Materials and Analytical Methods. The potassium salt of the dilacunary K14[P2W19O69(H2O)]·24H2O precursor was synthesized according to the literature procedure and confirmed by FT-IR spectroscopy.17 All other chemicals were commercially purchased and used without further purification. FT-IR spectra of the all compounds were recorded on a PerkinElmer BX spectrometer with KBr pellets. The magnetic data of 1a−10a were collected using a superconducting quantum interference device (SQUID) magnetometer (MPMS-5S, Quantum Design, Inc.). The temperature dependencies of the molar magnetic susceptibilities were measured in an applied magnetic field of 5 kOe for 1a−10a. Thermogravimetric analysis (TGA) measurements were performed on a model DTG-60 TG/DTA instrument (Shimadzu) in the temperature range of 25−600 °C with a heating rate of 5 °C min−1 in a nitrogen atmosphere. Elemental analyses were performed on an ICP-AES instrument, ARCOS from M/s Spectro Germany. High-resolution ESI-MS data were recorded on an LTQ Orbitrap XL (Thermo Fisher Scientific) with an accuracy of 3 ppm. Five milligrams of each sample was dissolved in 5 mL of H2O, and the solutions were diluted with CH3CN (final concentration of ∼10 μg/mL). Peak assignments were performed by comparing both isotropic distributions and m/z values (differences of 2σ(I)] no. of parameters R [I > 2σ(I)]a Rw (all data)b

Sm-1a

0.0216 6293 344 0.0452 0.1225

10724

Co4TbK11O88P2W20 5971.68 orthorhombic Iba2 27.1908(6) 19.7409(5) 18.4980(3) 90 9929.2(4) 4 150(2) 3.995 25.00 1.032 10420 22213

Tb-4a

0.0560 11258 360 0.0422 0.1076

73054

Co4DyK7NaO98P2W20 6001.85 orthorhombic Iba2 27.0510(3) 19.8907(3) 18.5159(3) 90 9962.7(2) 4 150(2) 4.001 24.81 1.021 10448 75330

Dy-5a

0.0240 6546 344 0.0542 0.1508

9901

Co4HoK11O88P2W20 5977.69 orthorhombic Iba2 26.9496(4) 19.8237(3) 18.5091(4) 90 9888.3(3) 4 150(2) 4.015 25.19 1.028 10392 11053

Ho-6a

0.0564 8702 356 0.0366 0.0895

63563

Co4ErK10.6O94P2W20 6060.38 orthorhombic Iba2 27.1582(7) 19.7015(6) 18.5347(5) 90 9917.1(5) 4 298(2) 4.059 25.16 1.042 10558 73261

Er-7a

Table 1. Crystallographic Data for Compounds Sm-1a, Eu-2a, Gd-3a, Tb-4a, Dy-5a, Ho-6a, Er-7a, Tm-8a, Yb-9a, and Lu-10a

0.0716 10103 345 0.0477 0.1229

75276

Co4TmK5O88P2W20 5747.09 orthorhombic Iba2 27.1821(10) 19.8250(7) 18.6039(6) 90 10025.4(6) 4 298(2) 3.808 24.69 1.068 9924 77577

Tm-8a

0.0822 10301 355 0.0519 0.1279

69406

Co4YbK10.4O88P2W20 5962.34 orthorhombic Iba2 27.2067(8) 19.7189(5) 18.6379(5) 90 9999.0(5) 4 298(2) 3.961 25.03 1.047 10224 71294

Yb-9a

0.1281 8575 320 0.0749 0.1890

69758

Co4LuK9.6O86P2W20 5900.89 orthorhombic Iba2 27.2030(11) 19.6409(10) 18.6791(6) 90 9980.1(7) 4 298(2) 3.927 25.095 1.034 10238 71955

Lu-10a

Inorganic Chemistry Article

DOI: 10.1021/acs.inorgchem.5b02772 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 2. (a) Ball-and-stick and polyhedral representation of polyanions [Ln{PCo2W10O38(H2O)2}2]11− (Ln = SmIII, EuIII, GdIII, TbIII, DyIII, HoIII, ErIII, TmIII, YbIII, and LuIII). (b) Coordination environment of the lanthanoid and cobalt ions. Color code: bright green for Ln, orange for Co, dark green for W polyhedra, pink for P, red for O, and aqua for H2O.

isolated in the presence of a lanthanoid ion to form sandwich type complexes. Another isomer that exists is tetra-Co species [Co4(OH2)2(PW9O34)]10− in low yield. Thermogravimetric Analysis. The thermal analysis of polyanions 1a−10a was performed under a N2 atmosphere in the temperature range of 25−600 °C with a heating rate of 5 °C min−1. The TGA curves exhibit overall a two-step weight loss in the temperature range of 25−300 °C. The first-step weight loss (temperature range of 25−130 °C) was associated with the loss of lattice water molecules, whereas coordinated water molecules emit at slight higher temperatures (130−300 °C) (see Figure S6). Magnetic Properties. The temperature dependence of χMT for 1a−10a (χM is the molar magnetic susceptibility per [Ln{PCo2W10O38(H2O)2}2] cluster) was measured on the polycrystalline samples in the temperature range from 1.8 to 300 K. At room temperature, the χMT values for 1a−10a are 12.2, 13.1, 17.0, 19.7, 24.0, 25.4, 21.7, 20.3, 15.1, and 13.3 emu K mol−1, respectively, which are larger than those expected for four isolated CoII ions (1.88 emu K mol−1 for CoII; g = 2) and one isolated LnIII; SmIII (0.09; g = 2), EuIII (0), GdIII (7.88; g = 2), TbIII (11.8; g = 2), DyIII (14.2; g = 2), HoIII (14.1; g = 2), ErIII (11.5; g = 2), TmIII (7.15; g = 2), YbIII (2.57; g = 2), and LuIII (0 emu K mol−1). The susceptibilities of 1a, 2a, 4a, and 8a−10a showed similar behavior (Figure 3a). When we focus on the magnetic susceptibility behavior of 10a, χMT slowly decreases upon cooling to 28 K, which is due to depopulation of the exited Kramers state of the CoII ion. Below 28 K, its χMT increases and reaches a maximum of 11.5 emu K mol−1 at 9.0 K, indicating ferromagnetic interactions between CoII and CoII in one half of the molecule. χMT rapidly decreases below 9.0 K, which indicates the presence of intra/intercluster antiferromagnetic interactions between the ferromagnetically coupled CoII dimers of the two halves of the molecule. The χMT values for 1a, 2a, 4a, 8a, and 9a, slowly decrease to 20−40 K, followed by a sharp increase originating from ferromagnetic CoII−CoII interaction to reach a maximum at 5−10 K; the decrease is due to antiferromagnetic interaction between CoII dimers or between CoII and LnIII. For 7a, the decrease originating from the orbital

compensated by the potassium cations. Elemental analyses also confirm the formula. Tourné et al. first synthesized such a di-Co-substituted Keggin type phosphotungstate in 1988 and characterized it by elemental analysis and X-ray powder diffraction analysis.17 Tourné et al. also reported that K14[P2W19O69(H2O)] is stable in neutral medium, but at lower or higher pH, it isomerizes into other POM compounds in the presence of potassium ions. At lower pH, it isomerizes into monovacant Keggin type [PW11O39K]6−, which reacts with divalent cobalt metal ions and leads to the formation of di-Co-substituted Keggin type phosphotungstate [PCo2W10O38(H2O)2]7−, which was reported to be mixtures of isomers. Dicobalt-substituted Keggin type phosphotungstate [PW10O38Co2(H2O)2]7− is unstable in aqueous solution, and it slowly transforms to tetra-Co species [Co4(OH2)2(PW9O34)2]10−.25 We have performed this reaction in the presence of lanthanoid cations, so the di-Co-substituted phosphotungstate sandwiched the Ln cation (see Figure 1) and crystallized as a single crystal suitable for single-crystal XRD so that the position of the Co can be studied. Single-crystal XRD studies also reveal that the series of compounds described here is explicitly one isomeric species of α-1,5 isomer. The oxygens next to divalent Co have basicity higher than that of oxygens next to hexavalent W, and the oxygens next to divalent Co have a higher affinity for Ln. The α-1,5 isomers are suitable for making the sandwich type complexes with the Ln cation because oxygen atoms belonging to the Co−O−W bridges in a (PCo2W10) subunit are arranged in a square geometry and can coordinate to Ln. After complete crystallization of the title compounds, if the mother liquor had been left for more than 3 weeks, few crystals (purple) of tetramer [Co4(OH2)2(PW9O34)2]10− compounds were isolated as a side product (see Figure S5). We have also reacted an early lanthanoid (La−Nd) salt under the same reaction conditions but isolated tetramer cobalt complexes [Co4(OH2)2(PW9O34)]10− with the salts of PrIII and NdIII, whereas in the case of LaIII and CeIII, an unknown precipitate was obtained. Hence, we can conclude that the reaction of K14[P2W19O69(H2O)] with only divalent Co(II) ions produced a mixture of isomers, in which only the α-1,5 isomer can be F

DOI: 10.1021/acs.inorgchem.5b02772 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 3. Plot of magnetic susceptibility (χMT) vs temperature (T) for (a) Sm-1a, Eu-2a, Tb-4a, Er-7a, Tm-8a, Yb-9a, and Lu-10a, (b) Gd-3a, Dy-5a, and Ho-6a.

Figure 4. (a) Negative ESI mass spectrum of 6a and (b) expanded region and simulated peaks for H8[Ho{Co2PW10O38}2]3−. G

DOI: 10.1021/acs.inorgchem.5b02772 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry contribution of both ErIII and CoII ions is not efficiently compensated by ferromagnetic CoII−CoII interaction and results in a plateau at low temperatures. On the other hand, the χMT values for 3a, 5a, and 6a decrease gradually with decreasing temperature above ∼9.3, ∼8.1, and ∼7.8 K, respectively, due to the orbital contribution of the LnIII ion and/or CoII ion and then decrease more rapidly with the antiferromagnetic interactions between CoII and LnIII (see Figure 3b). The isothermal magnetizations of 1a−10a have been measured at 2 K in the field between −50 and 50 kOe and show a gradual increase in magnetization at low fields, but the magnetizations do not reach saturation even at 50 kOe (see Figures S7−S9). ESI-MS. High-resolution ESI-MS of the POMs 1a−9a were performed after the sample had been dissolved in a CH3CN/ H2O solution. Figure 4 shows an ESI-MS spectrum of 6a; signals that can be assigned to H8[Ho{Co2PW10O38}2]3−, H7[Ho{Co2PW10O38}2]4−, and H6[Ho{Co2PW10O38}2]5− are observed. ESI mass spectra of 1a−10a are presented (see Figure S10). In all cases, signals that can be assigned to H8−x[Ln{Co2PW10O38}2](3+x)− are observed, indicating that the lanthanoid-sandwiched species are stable in solution. Some peaks that can be assigned to fragments such as [W6O19]2−, [Co2PW10O36]3−, and H[PW10O34]2− are observed. In some cases, peaks that can be assigned to Hn[Co4(PW9O34)2]n− are observed. [Co4(PW9O34)2]10− has a tetra-Co core sandwiched by two [PW9O34]9− ions, which might be a side product or a decomposed product in solution because it is known that [PW10O38Co2(H2O)2]7− transforms into [Co4(PW9O34)2]10−.17



AUTHOR INFORMATION

Corresponding Author

*Department of Chemistry, University of Delhi, North Campus, 110007 Delhi, India. Fax: +91 11 2766 6605. Telephone: +91 11 27666646. E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS F.H. gratefully acknowledges the Council of Scientific and Industrial Research (CSIR), New Delhi [Project 01(2712)/13/ EMR-II], and the University of Delhi for financial support of this work. We thank the University Scientific Instrumentation Centre (USIC) of the University of Delhi for providing instrument facilities and IIT Bombay for ICP-AES. R.G. thanks CSIR, New Delhi, for financial support. We thank Ms. T. Amimoto at the Natural Science Center for Basic Research and Development (N-BARD), Hiroshima University, for ESI-MS measurements. M.S. and S.N. are thankful for Grants-in-Aid for Scientific Research (C) 26420787 and (B) 16H04223, respectively, from the Ministry of Education, Culture, Sports, and Science of Japan.



CONCLUSION A series of 3d−4f heterometallic polyoxometalates have been synthesized by using the lacunary POMs under mild reaction conditions. The lanthanoid-sandwiched [Ln{PCo2W10O38(H2O)2}2]11− (Ln = SmIII, EuIII, GdIII, TbIII, DyIII, HoIII, ErIII, TmIII, YbIII, and LuIII) heteropolyanion compounds were synthesized following a one-pot reaction procedure by using dilacunary K14[P2W19O69(H2O)]·24H2O in a potassium chloride solution and characterized by single-crystal XRD, FT-IR, ICP-AES, HR-ESI-MS, and magnetic measurement as well as TGA. Single crystal XRD analyses of these compounds reveal that all titled polyanions are isostructural. The temperature dependence of magnetic susceptibilities for all can be rationalized by the dominant magnetic interactions, and the isothermal magnetizations show a gradual increase in magnetization at low fields and do not reach saturation even at 50 kOe. We have also reacted early lanthanoid; however, we isolated the tetramer cobalt complexes with PrIII and NdIII, whereas in the case of LaIII and CeIII, uncharacterized precipitates were obtained. To the best of our knowledge, such 3d−4f type heterometallic compounds represent the first examples to be isolated using dilacunary polyoxophosphotungstate. We are currently exploring this work and have isolated several complexes with other transition metal ions; results will be published in due course.



Crystallographic data (CIF) Crystallographic data (CIF) Crystallographic data (CIF) Crystallographic data (CIF) Crystallographic data (CIF) Crystalline material figure, a table of P−O, Co−O, and Ln−O bond lengths, a table of O−P−O bond angles, FT-IR spectra, TGA curves, additional figures of magnetic properties, and additional ESI-MS data (PDF) Crystallographic data (CIF) Crystallographic data (CIF)



REFERENCES

(1) (a) Pope, M. T. In Heteropoly and Isopoly Oxometalates; SpringerVerlag: Berlin, 1983. (b) Pope, M. T.; Müller, A. Angew. Chem., Int. Ed. Engl. 1991, 30, 34. (c) Pope, M. T., Müller, A., Eds. Polyoxometalates: From Platonic Solids to Anti-Retroviral Activity; Kluwer: Dordrecht, The Netherlands, 1994. (d) Müller, A.; Reuter, H.; Dillinger, S. Angew. Chem., Int. Ed. Engl. 1995, 34, 2328. (e) Hill, C. L. Chem. Rev. 1998, 98, 1. (f) Clemente-Juan, J. M.; Coronado, E. Coord. Chem. Rev. 1999, 193−195, 361. (g) Pope, M. T., Müller, A., Eds. Polyoxometalate Chemistry: From Topology via Self-Assembly to Applications; Kluwer: Dordrecht, The Netherlands, 2001. (h) Pope, M. T., Yamase, T., Eds. Polyoxometalates Chemistry for Nano-Composite Design; Kluwer: Dordrecht, The Netherlands, 2002. (i) Kamata, K.; Yonehara, K.; Sumida, Y.; Yamaguchi, K.; Hikichi, S.; Mizuno, N. Science 2003, 300, 964. (j) Pope, M. T. Comprehensive Coordination Chemistry II 2003, 635. (k) Hill, C. L. Comprehensive Coordination Chemistry II 2003, 679. (l) Zhao, J. W.; Wang, C. M.; Zhang, J.; Zheng, S. T.; Yang, G. Y. Chem. - Eur. J. 2008, 14, 9223. (m) Sun, C.-Y.; Liu, S.-X.; Liang, D.-D.; Shao, K.-Z.; Ren, Y.-H.; Su, Z.-M. J. Am. Chem. Soc. 2009, 131, 1883. (n) Wang, J. P.; Zhao, J. W.; Ma, P. T.; Ma, J. C.; Yang, L. P.; Bai, Y.; Li, M. X.; Niu, J. Y. Chem. Commun. 2009, 2362. (2) (a) Peacock, R. D.; Weakley, T. J. R. J. Chem. Soc. A 1971, 1836. (b) Peacock, R. D.; Weakley, T. J. R. J. Chem. Soc. A 1971, 0, 1937. (3) (a) Griffith, W. P.; Morley-Smith, N.; Nogueira, H. I. S.; Shoair, A. G. F.; Suriaatmaja, M.; White, A. J. P.; Williams, D. J. J. Organomet. Chem. 2000, 607, 146. (b) Chen, W.; Li, Y.; Wang, Y.; Wang, E.; Su, Z. Dalton Trans. 2007, 4293. (c) Merca, A.; Müller, A.; Van Slageren, J.; Läge, M.; Krebs, B. J. Cluster Sci. 2007, 18, 711. (4) (a) Hussain, F.; Spingler, B.; Conrad, F.; Speldrich, M.; Kögerler, P.; Boskovic, C.; Patzke, G. R. Dalton Trans. 2009, 4423. (b) Hussain,

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.5b02772. Crystallographic data (CIF) Crystallographic data (CIF) Crystallographic data (CIF) H

DOI: 10.1021/acs.inorgchem.5b02772 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry F.; Gable, R. W.; Speldrich, M.; Kögerler, P.; Boskovic, C. Chem. Commun. 2009, 328. (c) Hussain, F.; Conrad, F.; Patzke, G. R. Angew. Chem. 2009, 121, 9252; Angew. Chem., Int. Ed. 2009, 48, 9088. (d) Hussain, F.; Degonda, A.; Sandriesser, S.; Fox, T.; Mal, S. S.; Kortz, U.; Patzke, G. R. Inorg. Chim. Acta 2010, 363, 4324. (e) Hussain, F.; Patzke, G. R. CrystEngComm 2011, 13, 530. (f) Hussain, F.; Sandriesser, S.; Speldrich, M.; Patzke, G. R. J. Solid State Chem. 2011, 184, 214. (g) Gupta, R.; Saini, M. K.; Doungmene, F.; De Oliveira, P.; Hussain, F. Dalton Trans. 2014, 43, 8290. (h) Saini, M. K.; Gupta, R.; Parbhakar, S.; Kumar Mishra, A.; Mathur, R.; Hussain, F. RSC Adv. 2014, 4, 25357. (i) Saini, M. K.; Gupta, R.; Parbhakar, S.; Singh, S.; Hussain, F. RSC Adv. 2014, 4, 38446. (j) Gupta, R.; Saini, M. K.; Hussain, F. Eur. J. Inorg. Chem. 2014, 2014, 6031. (k) Gupta, R.; Hussain, F.; Behera, J. N.; Bossoh, A. M.; Mbomekalle, I. M.; De Oliveira, P. RSC Adv. 2015, 5, 99754. (5) (a) Mishra, A.; Wernsdorfer, W.; Abboud, K. A.; Christou, G. J. Am. Chem. Soc. 2004, 126, 15648. (b) Zhang, J. J.; Xia, S. Q.; Sheng, T. L.; Hu, S. M.; Leibeling, G.; Meyer, F.; Wu, X. T.; Xiang, S. C.; Fu, R. B. Chem. Commun. 2004, 1186. (c) Prasad, T. K.; Rajasekharan, M. V.; Costes, J. P. Angew. Chem., Int. Ed. 2007, 46, 2851. (6) (a) Wu, C. D.; Lu, C. Z.; Zhuang, H. H.; Huang, J. S. J. Am. Chem. Soc. 2002, 124, 3836. (b) Mialane, P.; Dolbecq, A.; Riviére, E.; Marrot, J.; Sécheresse, F. Eur. J. Inorg. Chem. 2004, 2004, 33. (7) (a) Reinoso, S. Dalton Trans. 2011, 40, 6610. (b) Liu, J.; Han, Q.; Chen, L.; Zhao, J. CrystEngComm 2016, 18, 842. (c) Zhao, J.-W.; Li, Y.-Z.; Chen, L.-J.; Yang, G.-Y. Chem. Commun. 2016, 52, 4418. (8) (a) Fang, X.; Kögerler, P. Chem. Commun. 2008, 3396. (b) Fang, X.; Kögerler, P. Angew. Chem., Int. Ed. 2008, 47, 8123. (c) Chen, W.; Li, Y.; Wang, Y.; Wang, E.; Zhang, Z. Dalton Trans. 2008, 865. (d) Yao, S.; Zhang, Z.; Li, Y.; Lu, Y.; Wang, E.; Su, Z. Cryst. Growth Des. 2010, 10, 135. (e) Reinoso, S.; Galán-Mascarós, J. R. Inorg. Chem. 2010, 49, 377. (f) Reinoso, S.; Galán-Mascarós, J. R.; Lezama, L. Inorg. Chem. 2011, 50, 9587. (g) Zhao, J.; Shi, D.; Chen, L.; Li, Y.; Ma, P.; Wang, J.; Niu, J. Dalton Trans. 2012, 41, 10740. (h) Zhao, H. Y.; Zhao, J. W.; Yang, B. F.; He, H.; Yang, G. Y. CrystEngComm 2013, 15, 5209. (i) Wang, J.; Zhao, J. W.; Zhao, H. Y.; Yang, B. F.; He, H.; Yang, G. Y. CrystEngComm 2014, 16, 252. (j) Zhao, H. Y.; Zhao, J. W.; Yang, B. F.; Wei, Q.; Yang, G. Y. J. Cluster Sci. 2014, 25, 667. (k) Zhang, Z.-H.; Zhang, Z.; Yang, B.-F.; He, H.; Yang, G.-Y. Inorg. Chem. Commun. 2016, 63, 65. (9) (a) Pang, H.; Zhang, C.; Shi, D.; Chen, Y. Cryst. Growth Des. 2008, 8, 4476. (b) Shi, D.; Chen, L.; Zhao, J.; Wang, Y.; Ma, P.; Niu, J. Inorg. Chem. Commun. 2011, 14, 324. (c) Zhang, S.; Zhao, J.; Ma, P.; Chen, H.; Niu, J.; Wang, J. Cryst. Growth Des. 2012, 12, 1263. (d) Zhao, J.; Luo, J.; Chen, L.; Yuan, J.; Li, H.; Ma, P.; Wang, J.; Niu, J. CrystEngComm 2012, 14, 7981. (e) Yu, T.; Ma, H.; Zhang, C.; Pang, H.; Li, S.; Liu, H. Dalton Trans. 2013, 42, 16328. (f) Pang, H.; GómezGarcía, C. J.; Peng, J.; Ma, H.; Zhang, C.; Wu, Q. Dalton Trans. 2013, 42, 16596. (g) Zhao, J. W.; Li, Y. Z.; Ji, F.; Yuan, J.; Chen, L. J.; Yang, G. Y. Dalton Trans. 2014, 43, 5694. (h) Zhao, H. Y.; Zhao, J. W.; Yang, B. F.; He, H.; Yang, G. Y. CrystEngComm 2014, 16, 2230. (i) Zhao, J.W; Cao, J.; Li, Y.-Z.; Zhang, J.; Chen, L.-J. Cryst. Growth Des. 2014, 14, 6217. (j) Zhou, W.; Zhang, Z.; Peng, J.; Wang, X.; Shi, Z.; Li, G. CrystEngComm 2014, 16, 10893. (k) Ma, X.; Song, K.; Cao, J.; Gong, P.; Li, H.; Chen, L.; Zhao, J. Inorg. Chem. Commun. 2015, 60, 65. (l) Chen, L.; Zhang, F.; Ma, X.; Luo, J.; Zhao, J. Dalton Trans. 2015, 44, 12598. (m) Chen, L.; Cao, J.; Li, X.; Ma, X.; Luo, J.; Zhao, J. CrystEngComm 2015, 17, 5002. (n) Fan, L.-Y.; Lin, Z.-G.; Cao, J.; Hu, C.-W. Inorg. Chem. 2016, 55, 2900. (o) Sun, L.; Liu, Y.; Wang, X.; Li, H.; Luo, J.; Chen, L.; Zhao, J. Synth. Met. 2016, 217, 256. (10) Merca, A.; Schnack, J.; Van Slageren, J.; Glaser, T.; Bögge, H.; Hoeke, V.; Läge, M.; Müller, A.; Krebs, B. J. Cluster Sci. 2013, 24, 979. (11) Chen, Y.-Z.; Liu, Z.-J.; Zhang, Z.-M.; Zhou, H.-Y.; Zheng, X.-T.; Wang, E.-B. Inorg. Chem. Commun. 2014, 46, 155. (12) Yao, S.; Yan, J.-H.; Duan, H.; Jia, Q.-Q.; Zhang, Z.-M.; Wang, E.B. RSC Adv. 2015, 5, 76206. (13) Ibrahim, M.; Mereacre, V.; Leblanc, N.; Wernsdorfer, W.; Anson, C. E.; Powell, A. K. Angew. Chem. 2015, 127, 15795; Angew. Chem., Int. Ed. 2015, 54, 15574.

(14) Artetxe, B.; Reinoso, S.; San Felices, L.; Lezama, L.; GutiérrezZorrilla, J. M.; Vicent, C.; Haso, F.; Liu, T. Chem. - Eur. J. 2016, 22, 4616. (15) Sato, R.; Suzuki, K.; Minato, T.; Yamaguchi, K.; Mizuno, N. Inorg. Chem. 2016, 55, 2023. (16) Xue, H.; Zhang, Z.; Pan, R.; Yang, B.-F.; Liu, H.-S.; Yang, G.-Y. CrystEngComm 2016, 18, 4643. (17) Tourné, C. M.; Tourné, G. F. J. Chem. Soc., Dalton Trans. 1988, 2411. (18) Xcalibur CCD System; Oxford Diffraction Ltd.: Abingdon, Oxfordshire, England, 2007. (19) CrysAlis Pro software system, version 171.32; Oxford Diffraction Ltd.: Abingdon, Oxfordshire, England, 2007. (20) (a) Sheldrick, G. M. Acta Crystallogr., Sect. A: Found. Crystallogr. 2008, 64, 112. (b) Sheldrick, G. M. Acta Crystallogr. 2015, C71, 3. (21) SCALE3 ABSPACK, CrysAlisPro, version 1.171.36.32; Oxford Diffraction Ltd.: Abingdon, Oxfordshire, England, 2013. (22) (a) Pettersson, L.; Andersson, I.; Selling, A.; Grate, J. H. Inorg. Chem. 1994, 33, 982. (b) Guan, W.; Yan, L.; Su, Z.; Liu, S.; Zhang, M.; Wang, X. Inorg. Chem. 2005, 44, 100. (23) (a) Ozeki, T.; Yamase, T. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1991, 47, 693. (b) Yamase, T.; Ozeki, T.; Motomura, S. Bull. Chem. Soc. Jpn. 1992, 65, 1453. (24) (a) Brown, I. D.; Altermatt, D. Acta Crystallogr., Sect. B: Struct. Sci. 1985, 41, 244. (b) Trzesowska, A.; Kruszynski, R.; Bartczak, T. J. Acta Crystallogr., Sect. B: Struct. Sci. 2004, 60, 174. (25) (a) Weakley, T. J. R.; Evans, H. T.; Showell, J. S.; Tourné, G. F.; Tourné, C. M. J. Chem. Soc., Chem. Commun. 1973, 139. (b) Evans, H. T.; Tourné, C. M.; Tourné, G. F.; Weakley, T. J. R. J. Chem. Soc., Dalton Trans. 1986, 2699.

I

DOI: 10.1021/acs.inorgchem.5b02772 Inorg. Chem. XXXX, XXX, XXX−XXX