Mixed-Ligand Approach to Changing the Metal Ratio in Bismuth

Apr 7, 2016 - Application of the mixed-ligand approach allows one to change the connectivity pattern within the heterometallic assembly and to isolate...
0 downloads 10 Views 710KB Size
Article pubs.acs.org/IC

Mixed-Ligand Approach to Changing the Metal Ratio in Bismuth− Transition Metal Heterometallic Precursors Craig M. Lieberman, Zheng Wei, Alexander S. Filatov, and Evgeny V. Dikarev* Department of Chemistry, University at Albany, SUNY, Albany, New York 12222, United States S Supporting Information *

ABSTRACT: A new series of heteroleptic bismuth−transition metal βdiketonates [BiM(hfac)3(thd)2] (M = Mn (1), Co (2), and Ni (3); hfac = hexafluoroacetylacetonate, thd = tetramethylheptanedionate) with Bi:M = 1:1 ratio have been synthesized by stoichiometric reactions between homometallic reagents [BiIII(hfac)3] and [MII(thd)2]. On the basis of analysis of the metal− ligand interactions in heterometallic structures, the title compounds were formulated as ion-pair {[BiIII(thd)2]+[MII(hfac)3]−} complexes. The direct reaction between homometallic reagents proceeds with a full ligand exchange between main group and transition metal centers, yielding dinuclear heterometallic molecules. In heteroleptic molecules 1−3, the Lewis acidic, coordinatively unsaturated BiIII centers are chelated by two bulky, electrondonating thd ligands and maintain bridging interactions with three oxygen atoms of small, electron-withdrawing hfac groups that chelate the neighboring divalent transition metals. Application of the mixed-ligand approach allows one to change the connectivity pattern within the heterometallic assembly and to isolate highly volatile precursors with the proper Bi:M = 1:1 ratio. The mixed-ligand approach employed in this work opens broad opportunities for the synthesis of heterometallic (main group−transition metal) molecular precursors with specific M:M′ ratio in the case when homoleptic counterparts either do not exist or afford products with an incorrect metal:metal ratio for the target materials. Heteroleptic complexes obtained in the course of this study represent prospective single-source precursors for the low-temperature preparation of multiferroic perovskite-type oxides.



INTRODUCTION Over the years, bismuth-based oxides have received considerable attention as prospective materials for a variety of applications including oxidation catalysts,1 high-Tc superconductors,2 high-temperature electrolytes,3 nonlinear optical materials,4 and ferroelectrics.5 A field of growing importance is the synthesis of multiferroics,6 a class of materials that simultaneously possess two or more of the so-called “ferroic” order parameters: ferroelectricity, ferromagnetism, and/or ferroelasticity. Metal oxides have been the focus of active research in the field of multiferroics, specifically those with a perovskite-type structure formulated as ABO3, in which the A site is a heavy metal cation (such as Tb, Y, Lu Bi, Pb, Ba) in conjunction with a small B site cation composed of 3d transition metals (Ti−Ni). Among those, bismuth−transition metal oxides represent an attractive group of materials providing an environmentally friendly alternative to lead and barium counterparts. The synthesis and characterization of various bismuth−transition metal oxides of the formula BiMO3 (M = Cr−Ni)7−11 with desirable properties have been reported. Thus, BiMnO3 simultaneously undergoes ferroelectric and ferromagnetic transitions when cooled below 500 K (TE) and 100 K (TC), respectively.12 BiNiO3 has been shown to exhibit phase transitions at elevated temperatures and pressures resulting in a volume reduction of up to 2.6%, a process known as negative thermal expansion, which makes it of © XXXX American Chemical Society

fundamental interest to design and investigate the composites with zero or controlled thermal expansion values.13 The most widely studied oxide is bismuth ferrite, BiFeO3, which belongs to a class of materials known as “lone-pair” multiferroics, which make use of the stereochemically active electron pair of the large A-site cation to induce ferroelectricity.14 This oxide exhibits antiferromagnetic and ferroelectric behaviors, and it has been reported15 that epitaxial BiFeO3 thin films display large room-temperature spontaneous polarization, making it an attractive material for advanced technological devices. The majority of these bismuth−transition metal oxides (M = Ni, Co, Mn) were synthesized through conventional solid-state reactions between multiple molecular precursors requiring high temperatures and pressures up to 6 GPa16 and often contain some additional impurities.17 Nevertheless, single-phase thin films of BiMnO3 have been obtained under ambient pressure conditions by pulsed laser deposition, in which the highpressure requirement for phase stabilization was replaced by the epitaxial strain imposed by the substrate.18 Milder conditions associated with MOCVD processes have also been used for the preparation of various ferroelectic and/or multiferroic thin films and nanostructures based on perovskite-type oxides.19 To the best of our knowledge, reports on the use of single-source Received: January 27, 2016

A

DOI: 10.1021/acs.inorgchem.6b00209 Inorg. Chem. XXXX, XXX, XXX−XXX

Inorganic Chemistry

Article



RESULTS AND DISCUSSION Our recent work27 on heterometallic lead−iron β-diketonates introduced a mixed-ligand approach that allows one to effectively change the connectivity pattern within the heterometallic assembly in order to obtain the target singlesource precursor with molecular structure and with a proper metal:metal ratio. The present work extends an application of the mixed-ligand approach to the family of bismuth−transition metal precursors. Using a combination of two different diketonate ligands, volatile heterometallic complexes [BiM(hfac)3(thd)2] with 1:1 metal ratio have been isolated as the sole products of the solid-state stoichiometric reactions between [Bi(hfac)3] and [M(thd)2] (M = Mn (1), Co (2), Ni (3); for synthetic details, see Supporting Information, Table S1):

precursors (SSP) for the preparation of perovskite-type bismuth−transition metal oxides of the formula BiMO3 (M = Ti−Ni) are limited to the organometallic complex [CpFe(CO)2BiCl2], which was employed to grow BiFeO3 thin films by an aerosol-assisted CVD process.20 The preparation of heterobimetallic complexes incorporating bismuth and transition metals is quite challenging due to the differences in electronic and coordination requirements for the metal centers. These problems are further exaggerated when attempting to design bismuth-containing complexes with discrete molecular structures, since bismuth is a strong Lewis acid that can readily expand its coordination sphere to adopt hypervalent coordination.21 Nevertheless, the preparation of certain molecular heterobimetallic bismuth−transition metal coordination complexes has been achieved for various metal carbonyl complexes dating back over 20 years ago.22 More recently, the synthesis of molecular heterobimetallic bismuth complexes incorporating early transition metals (Ti, Ta, Nb, and V) and, for the most part, constructed from alkoxides or carboxylates has been reported.21,23 Furthermore, the application of heterometallic bismuth complexes incorporating group IV and V metals as single-source precursors for the preparation of mixed-metal oxide materials with ionic conducting24 and ferroelectric25 properties has been successfully demonstrated. Previously, we have reported26 the synthesis of the first family of heterobimetallic diketonates, [Bi2M(hfac)8] (M = Mn−Zn), having trinuclear molecular structure (Scheme 1).

[Bi(hfac)3 ] + [M(thd)2 ] → [BiM(hfac)3 (thd)2 ]

(1)

A certain advantage of the above synthetic approach is an availability of starting reagents. While nickel diketonate is commercially available, the diketonates of bismuth and cobalt can be readily obtained with high yields by simple one-step reactions. The sole exception is manganese diketonate, which is extremely moisture-/air-sensitive and requires special precautions for its preparation and handling. Crystals of heterometallic products 1−3 can be conveniently collected from the cold end of the containers with nearly quantitative yields. The purity of the products was confirmed by comparison of the X-ray powder diffraction patterns with theoretical ones calculated on the basis of single crystal data (SI, Figures S1−S4). Room-temperature solution synthesis in noncoordinating solvent (dichloromethane) was shown to afford fast, high-yield preparation of heterometallic precursors 2 and 3 on a large scale. Heterometallic β-diketonates 1−3 were found to be relatively stable in open air and can be handled outside the glovebox for a reasonable period of time in the course of characterization. All compounds are volatile and can be quantitatively resublimed at 85 (1) or 105 °C (2, 3) in an evacuated ampule. The sublimation starts at ca. 75 (1) or 85 °C (2, 3), and the precursors do not show any signs of decomposition until about ca. 95 (1) or 115 °C (2, 3). There is no visible indication of heterometallic complex dissociation into homometallic [Bi(βdik)3] and [M(β-dik)2] components prior to decomposition. The results of DART (direct analysis in real time) mass spectrometric investigation of [BiNi(hfac)3(thd)2] (3) in the gas phase support the above observations. While severe fragmentation of the heterometallic molecule occurs, the peak corresponding to parent ion [M + H]+ (M = [BiNi(hfac)3(thd)2]) with a characteristic isotope distribution pattern (SI, Figure S17, Table S8) can be clearly detected in the spectra obtained at the temperatures below compound 3’s decomposition point. X-ray structural investigation of heteroleptic diketonates 1−3 revealed dinuclear Bi:M = 1:1 molecules {[Bi(thd)2][M(hfac)3]} with two chelating thd ligands on the bismuth center and three chelating hfac ligands on the transition metal atom (Figure 1), indicating a full ligand exchange between starting reagents during the course of reaction. Such ligand exchange is a well-documented event in metal diketonate chemistry.28 The discrete molecular structure is held together by three bridging interactions between bismuth and oxygens from each of the transition metal−chelating hfac groups. These interactions are in a range of 2.83−3.48 Å (Table 1), which

Scheme 1

These homoleptic complexes were shown to be volatile and to retain the heterometallic structure in solution of noncoordinating solvents. However, the 2:1 bismuth to transition metal ratio in the precursor was incorrect for the target oxides. We were particularly interested in targeting heterometallic bismuth− transition metal diketonate complexes with the proper 1:1 metal ratio as potential SSPs for the perovskite-type mixedmetal oxides. Herein, we report the preparation and characterization of a new class of volatile and soluble heterometallic heteroleptic compounds with dinuclear structures [BiM(hfac)3(thd)2] (hfac = hexafluoroacetylacetonate, thd = tetramethylheptanedionate), which contain BiIII and divalent first-row transition metals Mn (1), Co (2), and Ni (3). Prospective precursors were synthesized by a simple stoichiometric reaction that proceeds via full ligand exchange between homometallic reactants. B

DOI: 10.1021/acs.inorgchem.6b00209 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

dionate hydrate [Bi(thd)3·H2O]31 (2.26−2.47 Å). Three additional bridging interactions in molecules 1−3 result in a coordination number of 7 around the bismuth center, the polyhedron of which can be described as a distorted monocapped trigonal prism (SI, Figure S10). In addition, there are Bi···F interactions in 1 and 2 of 3.32−3.36 Å, which are shorter than the sum of van der Waals radii. The coordination of Bi in the title complexes along with the fact that Ni(III) diketonates are unknown supports the formulation of heterometallic diketonates 1−3 as ion pairs {[Bi(thd)2]+[M(hfac)3]−}. Indeed, the octahedrally coordinated tris-chelated anions [M(hfac)3]− are well-known in the chemistry of the first-row transition metal diketonates. Thus, the [Mn(hfac)3] fragment in 1, having three chelating (2.14 Å) and three chelating-bridging (2.17 Å) diketonate oxygens, is significantly different from [MnIII(hfac)3] diketonate, which features a characteristic Jahn−Teller distortion with four short (1.92 Å) and two long (2.14 Å) Mn−O bonds (Table 2). On the other hand, the above fragment matches well with all [MnII(hfac)3]− anions previously reported in the literature.32 For example, a similar anion in the structure of heterometallic lead−transition metal diketonate [PbMn2(hfac)6]33 also features three chelating and three chelating-bridging oxygens that bridge to the Pb center with Mn−O distances of 2.13 and 2.22 Å, respectively. Likewise, while the Co−O distances in 2 (2.06 and 2.07 Å) are much longer than those in the structure of [CoIII(hfac)3] (1.87 Å), they correspond well with the bond lengths in known [CoII(hfac)3]− anions.32a,34 The direct analogue is [BaCo2(hfac)6], which features cobalt−ligand distances of 2.05 and 2.08 Å for three chelating and three chelating-bridging (to Ba center) oxygens, respectively. Similarly, the Ni−O distances for chelating (2.02 Å) and chelating-bridging (2.04 Å) diketonates in heterometallic complexes 3a and 3b are virtually the same as those in reported [NiII(hfac)3]− anions.32a,35,37 X-ray structural investigation of the Bi−Ni complex 3 revealed two polymorph modifications depending on the synthetic method used for its preparation. From the solidstate/gas-phase synthesis, the monoclinic modification 3a was

Figure 1. Molecular structure of {[Bi(thd)2][M(hfac)3]} (M = Mn (1), Co (2), and Ni (3)) heterometallic diketonates. All atoms are represented by spheres of arbitrary radii. Bridging interactions to bismuth are shown as dashed lines.

Table 1. Metal−Oxygen Distances (Å) in the Structures of Heterometallic Diketonates 1−3a 1 2 3a 3b a

Bi−Oc (4×)

Bi−Ob (3×)

M−Oc (3×)

M−Oc‑b (3×)

2.12−2.27 2.13−2.31 2.13−2.28 2.12−2.31

2.87−3.48 2.83−3.15 2.84−3.18 2.83−3.15

2.13−2.15 2.04−2.07 2.01−2.03 2.01−2.04

2.15−2.20 2.07−2.08 2.03−2.04 2.03−2.04

c, chelating; b, bridging; c-b, chelating-bridging.

are shorter than the sum of the corresponding van der Waals radii (3.59 Å).29 Two chelating thd ligands attached to bismuth have an angle of ca. 90° between the diketonate planes. This principal coordination number of 4 represents an unusually low coordination for bismuth ion, even taking into account the bulkiness of the thd groups. The Bi−O distances seem somewhat short even for the BiIII ion. The chelating Bi−O bonds of 2.12−2.31 Å in 1−3 (Table 1) are significantly shorter than those observed in six-coordinated aryl-bismuth bisdiketonate adducts [BiPh(β-dik)2L]30 (2.42−2.52 Å; L = donor solvent) as well as in bismuth(III) tetramethylheptane-

Table 2. Average M−O Distances (Å) in Heterometallic Diketonates 1−3 and in Related Compounds +



{[Bi(thd)2] [Mn(hfac)3] } (1) [MnIII(hfac)3]36 [PbMn2(hfac)6]33 [NaMn(hfac)3]∞37 {[PyH]+[Mn(hfac)3]−}36 {[Bi(thd)2]+[Co(hfac)3]−} (2) [CoIII(hfac)3]38 [BaCo2(hfac)6]38 [NaCo(hfac)3]∞37 {[C14Py]+[Co(hfac)3]−}34e {[C4mim]+[Co(hfac)3]−}34c {[Bi(thd)2]+[Ni(hfac)3]−} (3a) {[Bi(thd)2]+[Ni(hfac)3]−} (3b) [PbNi(hfac)4]∞39 [NaNi(hfac)3]∞37 {[Ni(2-PyBN)2hfac]+[Ni(hfac)3]−}32a {[NEt4]+[Ni(hfac)3]−}35b

BiIII−Oca

MII−Oca

2.191(3)

2.136(3) 2.144(3) (2×),1.922(3) (4×) 2.127(4) (3×) 2.157(6)d 2.055(2) 1.872(2) 2.052(2) (3×)

2.205(2)

2.061(2)d 2.067(3)d 2.022(3) 2.023(3) 2.017(6) (2×)

2.202(3) 2.206(3)

MII−Oc‑bb 2.169(3) 2.219(4) (3×) 2.142(2)c 2.071(2) 2.076(2) (3×) 2.066(2)c

2.035(3) 2.037(3) 2.030(6) 2.022(5)c

2.027(5)d 2.031(2)d

a

Chelating. bChelating-bridging. cAll diketonate oxygens are chelating-bridging. dAll diketonates are chelating; [C14Py] = (N-tetradecyl)pyridinium; [C4mim] = 1-butyl-3-methylimidazolium; (2-PyBN) = N-tert-butyl-α-(2-pyridyl)nitrone. C

DOI: 10.1021/acs.inorgchem.6b00209 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

electron-donating substituents on a main group metal forming the complex cation as well as of sterically uncongested diketonates with electron-withdrawing substituents on a back π-donating transition metal making the anion is noteworthy. While the [Bi(thd)2]+ fragment is unique, the appearance of the title complexes is in line with the well-documented tendency of forming very stable [M(hfac)3]− anions that readily provide bridging interactions to coordinatively unsaturated metal centers.37,39,40 The mixed-ligand approach employed in this work opens broad opportunities for the synthesis of other heterometallic (main group−transition metal) molecular precursors with specific M:M′ ratios, for which the homoleptic counterparts do not exist or form the complex having either a wrong M:M′ ratio for the target material or a polymeric structure. The title, highly volatile heterometallic complexes obtained in the course of this study are attractive single-source precursors for the low-temperature preparation of heterometallic perovskite-type oxides.

obtained, while crystals grown from a solution of dichloromethane adhere to triclinic form 3b, which appears to be isomorphous with Bi−Co complex 2, which crystallizes as the same modification from both the gas phase and solution (SI, Table S2). While the metal−oxygen distances in 3a and 3b are similar (Table 1), the crystal packing of bimetallic units was found to be distinctly different (SI, Figure S9). The Bi−Mn complex 1 also conforms to triclinic symmetry and to the same molecular packing as 2 and 3b, although with noticeably different unit cell parameters. Interestingly, the isolation of the monoclinic polymorph modification 3a helped to comprehend the powder pattern of the bulk product obtained from the solid-state synthesis of Bi−Mn complex 1. Successful LeBail fit confirmed the presence of the monoclinic modification as a minor component (SI, Figure S1). However, we were unable to pick up a single crystal of the latter polymorph. In accord with their molecular structures and the nature of diketonate ligands, compounds 1−3 are soluble in both coordinating (acetone, THF, DMSO) and noncoordinating (CH2Cl2, CHCl3, hexanes, benzene) solvents. Furthermore, complexes 2 and 3 can be recrystallized from the latter solvents while keeping heterometallic structure intact. The solution behavior of heterometallic precursors has been studied by NMR. The 1H and 19F NMR spectra of the solutions of 1−3 in coordinating solvents (SI, Figures S14 and S15) indicate the disruption of heterometallic molecules accompanied by partial ligand redistribution. Both proton and fluorine signals correspond to heteroleptic BiIII diketonate complexes [Bi(hfac)(thd)2] and [Bi(hfac)2(thd)]. The second dissociation product, [M(hfac)2(sol)2], is NMR silent, as we have already noted in previous studies.32e,33 Mass spectrometric investigation undoubtedly revealed the presence of heteroleptic homometallic species in an acetone solution of complex 3. Several mixed-ligand fragments are easily identifiable (SI, Figure S18, Table S9) in the spectra obtained in positive mode: [Ni(hfac)(thd)+H]+ (meas/calcd 449.071/449.069), [Bi(hfac)(thd)2+H]+ (meas/calcd 783.238/783.253), and [Bi(hfac)2(thd)+H]+ (meas/calcd 807.106/807.103). At the same time, in polar noncoordinating solvents, only proton signals corresponding to cationic [Bi(thd)2]+ fragment appear in the spectra (SI, Figure S16), while the anionic [M(hfac)3]− counterpart is NMR silent. The absence of any signal in 19F NMR indicates that no ligand redistribution occurs in noncoordinating solvents.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.6b00209. Experimental procedures, IR and NMR spectra, and Xray powder diffraction patterns for precursors 1−3 (PDF) Crystallographic data (CIF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Notes

The authors declare no competing financial interest.

■ ■

ACKNOWLEDGMENTS Financial Support from the National Science Foundation (CHE-1152441) is gratefully acknowledged. REFERENCES

(1) (a) Felthouse, T. R. J. Am. Chem. Soc. 1987, 109, 7567−1569. (b) Mamedov, E. A. Catal. Rev.: Sci. Eng. 1994, 36, 1−23. (c) Arora, N.; Deo, G.; Wachs, I. E.; Hirt, A. M. J. Catal. 1996, 159, 1−13. (d) Grasselli, R. K. Top. Catal. 2002, 21, 79−88. (2) (a) Maeda, H. Materia 2001, 40, 947. (b) Maeda, H.; Tanaka, Y.; Fukuyama, M.; Asano, T. Jpn. J. Appl. Phys. 1988, 27, L209. (c) Vezzoli, G. C.; Chen, M. F.; Craver, F.; Safari, A.; Moon, B. M.; Lalevic, B.; Burke, T.; Shoga, M. J. Magn. Magn. Mater. 1990, 88, 351−364. (d) Wasa, K.; Adachi, H.; Hirochi, K.; Ichikawa, Y.; Matsushima, T.; Setsune, K. J. Mater. Res. 1991, 6, 1595−1604. (e) Giannini, E.; Clayton, N.; Musolino, N.; Gladyshevskii, R.; Flukiger, R. Front. Superconduct. Mater. 2005, 739−764. (3) (a) Wachsman, E. D.; Jayaweera, P.; Jiang, N.; Lowe, D. M.; Pound, B. G. J. Electrochem. Soc. 1997, 144, 233−236. (b) Park, J.-Y.; Wachsman, E. D. Ionics 2006, 12, 15−20. (4) (a) Nasu, H.; Sato, M.; Matsuoka, J.; Kamiya, K. Nippon Seramikkusu Kyokai Gakujutsu Ronbunshi 1996, 104, 777−780. (b) Fok, M. P. IEEE J. Sel. Top. Quantum Electron. 2008, 14, 587− 598. (c) Grobnic, D.; Walker, R. B.; Mikhailov, S. J.; Smelser, C. W.; Lu, P. IEEE Photonics Technol. Lett. 2010, 22, 124−126. (5) (a) Zhang, S.-T.; Sun, B.; Yang, B.; Chen, Y.-F.; Liu, Z.-G.; Ming, N.-B. Mater. Lett. 2001, 47, 334−338. (b) Ando, A.; Kimura, M.; Sawada, T.; Hayashi, K.; Sakabe, Y. Ferroelectrics 2002, 268, 65−70.



CONCLUSIONS A new class of heteroleptic bismuth−transition metal βdiketonates {[BiIII(thd)2]+[MII(hfac)3]−} (M = Mn (1), Co (2), and Ni (3)) with Bi:M = 1:1 ratio have been synthesized by applying the mixed-ligand approach. The interaction between homometallic starting reagents [Bi(hfac)3] and [M(thd)2] results in a ligand exchange between main group and transition metal centers and formation of dinuclear heterometallic assemblies. The reaction represents the most profound example of a full diketonate ligand exchange between different metal atoms without accompanied effects such as a redox process/electron transfer or participation of the third type (nondiketonate) of ligands. In heteroleptic complexes 1− 3, the Lewis acidic, coordinatively unsaturated BiIII centers are chelated by two tetramethylheptanedionate groups and maintain bridging interactions with three oxygen atoms of hexafluoroacetylacetonates that chelate the neighboring MII atoms. The specific location of bulky diketonates with D

DOI: 10.1021/acs.inorgchem.6b00209 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry (6) Martin, L. W.; Schlom, D. G. Curr. Opin. Solid State Mater. Sci. 2012, 16, 199−215. (7) (a) Murakami, M.; Fujino, S.; Lim, S.-H.; Long, C. J.; SalamancaRiba, L. G.; Wuttig, M.; Takeuchi, I. Appl. Phys. Lett. 2006, 88, 152902. (8) Chiba, H.; Atou, T.; Syono, Y. J. Solid State Chem. 1997, 132, 139−143. (9) Jia, D.-C.; Xu, J.-H.; Ke, H.; Wang, W.; Zhou, Y. J. Eur. Ceram. Soc. 2009, 29, 3099−3103. (10) Sun, B.; Li, H.; Wei, L.; Chen, P. RSC Adv. 2014, 4, 50102− 50106. (11) Azuma, M.; Carlsson, S.; Rodgers, J.; Tucker, M. G.; Tsujimoto, M.; Ishiwata, S.; Isoda, S.; Shimakawa, Y.; Takano, M.; Attfield, J. P. J. Am. Chem. Soc. 2007, 129, 14433−14436. (12) (a) Moreira dos Santos, A.; Parashar, S.; Raju, A. R.; Zhao, Y. S.; Cheetham, A. K.; Rao, C. N. R. Solid State Commun. 2002, 112, 49− 52. (b) Chi, Z. H.; Xiao, C. J.; Feng, S. M.; Li, F. Y.; Jin, C. Q.; Wang, X. H.; Chen, R. Z.; Li, L. Y. J. Appl. Phys. 2005, 98, 103519. (13) Azuma, M.; Chen, W.-T.; Seki, H.; Czapski, M.; Olga, S.; Oka, K.; Mizumaki, M.; Watanuki, T.; Ishimatsu, N.; Kawamura, N.; Ishiwata, S.; Tucker, M. G.; Shimakawa, Y.; Attfield, J. P. Nat. Commun. 2011, 2, 347. (14) Khomskii, D. Physics 2009, 2, 20. (15) Wang, J.; Neaton, J. B.; Zheng, H.; Nagarajan, V.; Ogale, S. B.; Liu, B.; Viehland, D.; Vaithyanathan, V.; Schlom, D. G.; Waghmare, U. V.; Spaldin, N. A.; Rabe, K. M.; Wuttig, M.; Ramesh, R. Science 2003, 299, 1719−1722. (16) (a) Atou, T.; Chiba, H.; Ohoyama, K.; Yamaguchi, Y.; Syono, Y. J. Solid State Chem. 1999, 145, 639−642. (b) Ishiwata, S.; Azuma, M.; Takano, M.; Nishibori, E.; Takata, M.; Sakata, M.; Kato, K. J. Mater. Chem. 2002, 12, 3733−3737. (17) (a) Kimura, T.; Kawamoto, S.; Yamada, I.; Azuma, M.; Takano, M.; Tokura, Y. Phys. Rev. B: Condens. Matter Mater. Phys. 2003, 67, 180401(R)1−4. (b) Belik, A. A.; Likubo, S.; Kodama, K.; Igawa, N.; Shamoto, S.; Niitaka, S.; Azuma, M.; Shimakawa, Y.; Takano, M.; Izumi, F.; Takayama-Muromachi, E. Chem. Mater. 2006, 18, 798−803. (18) (a) Moreira dos Santos, A. F.; Cheetham, A.; Kian, W.; Pan, X. Q.; Jia, Y. F.; Murphy, N. J.; Lettieri, J.; Schlom, D. G. Appl. Phys. Lett. 2004, 84, 91−93. (b) Gajek, M.; Bibes, M.; Barthélémy, A.; Bouzehouane, K.; Fusil, S.; Varela, M.; Fontcuberta, J.; Fert, A. Phys. Rev. B: Condens. Matter Mater. Phys. 2005, 72, 020406. (c) Eerenstein, W.; Morrison, F. D.; Scott, J. F.; Mathur, N. D. Appl. Phys. Lett. 2005, 87, 101906. (d) Yang, C.-H.; Koo, T. Y.; Lee, S.-H.; Song, C.; Lee, K.B.; Jeong, Y. H. Europhys. Lett. 2006, 74, 348−354. (e) Schmidt, R.; Langenberg, E.; Ventura, J.; Varela, M. Bi Containing Multiferroic Perovskite Oxide Thin Films. In Perovskite: Crystallography, Chemistry and Catalytic Performance; Zhang, J.; Li, H., Ed.; Nova Science Publishers, Inc.: USA, 2013; pp 1−15. (19) (a) Kaiser, D. L.; Vaudin, M. D.; Rotter, L. D.; Wang, Z. L.; Cline, J. P.; Hwang, C. S.; Marinenko, R. B.; Gillen, J. G. Appl. Phys. Lett. 1995, 66, 2801−2803. (b) Moret, M. P.; Schermer, J. J. J. Appl. Phys. 2002, 92, 3947−3957. (c) Martin, L. W.; Crane, S. P.; Chu, Y.H.; Holcomb, M. B.; Gajek, M.; Huijben, M.; Yang, C.-H.; Balke, N.; Ramesh, R. J. Phys.: Condens. Matter 2008, 20, 434220. (d) Weiss, F.; Audier, M.; Bartasyte, A.; Bellet, D.; Girardot, C.; Jimenez, C.; Kreisel, J.; Pignard, S.; Salaun, M.; Ternon, C. Pure Appl. Chem. 2009, 81, 1523−1534. (e) Borodavka, F.; Gregora, I.; Bartasyte, A.; Margueron, S.; Plausinaitiene, V.; Abrutis, A.; Hlinka, J. J. Appl. Phys. 2013, 113, 187216. (f) Ichinose, D.; Nakashima, T.; Ehara, Y.; Oikawa, T.; Shimizu, T.; Sakata, O.; Yamada, T.; Funakubo, H. Appl. Phys. Lett. 2015, 107, 022902. (20) Moniz, S. J. A.; Quesada-Cabrera, R.; Blackman, C. S.; Tang, J.; Southern, P.; Weaver, P. M.; Carmalt, C. J. J. Mater. Chem. A 2014, 2, 2922−2927. (21) Ould-Ely, T.; Thurston, J. H.; Whitmire, K. H. C. R. Chim. 2005, 8, 1906−1921. (22) (a) Zouchoune, B.; Ogliaro, F.; Halet, J.-F.; Saillard, J.-Y.; Eveland, J. R.; Whitmire, K. H. Inorg. Chem. 1998, 37, 865−875. (b) Eveland, J. R.; Saillard, J.-Y.; Whitmire, K. H. Inorg. Chem. 1997, 36, 330−334. (c) Eveland, J. R.; Saillard, J.-Y.; Whitmire, K. H. Inorg.

Chem. 1997, 36, 4387−4396. (d) Eveland, J. R.; Whitmire, K. H.; Saillard, J.-Y. Inorg. Chem. 1995, 34, 1542−1551. (e) Cassidy, J. M.; Whitmire, K. H. Inorg. Chem. 1991, 30, 2788−2795. (f) Albright, T. A.; Yee, K. A.; Saillard, J. Y.; Kahlal, S.; Halet, J. F.; Leigh, J. S.; Whitmire, K. H. Inorg. Chem. 1991, 30, 1179−1190. (g) Whitmire, K. H.; Shieh, M.; Lagrone, C. B.; Robinson, B. H.; Churchill, M. R.; Fettinger, J. C.; See, R. F. Inorg. Chem. 1987, 26, 2798−2807. (h) Whitmire, K. H.; Leigh, J. S.; Gross, M. E. J. Chem. Soc., Chem. Commun. 1987, 0, 926− 927. (i) Bachman, R. E.; Whitmire, K. H. Inorg. Chem. 1995, 34, 1542−1551. (j) Whitmire, K. H.; Eveland, J. R. J. Chem. Soc., Chem. Commun. 1994, 1335−1336. (23) (a) Parola, S.; Papiernik, F.; Hubert-Pfalzgraf, L. G.; Jagner, S.; Hakansson, M. J. Chem. Soc., Dalton Trans. 1997, 4631. (b) Thurston, J. H.; Kumar, A.; Hofmann, C.; Whitmire, K. H. Inorg. Chem. 2004, 43, 8427−8436. (c) Mehring, M. Coord. Chem. Rev. 2007, 251, 974−1006. (24) (a) Thurston, J. H.; Whitmire, K. H. Chem. Mater. 2003, 15, 4407−4416. (b) Thurston, J. H.; Trahan, D.; Ould-Ely, T.; Whitmire, K. H. Inorg. Chem. 2004, 43, 3299−3305. (c) Bulimestru, I.; Shova, S.; Popa, N.; Roussel, P.; Capet, F.; Vannier, R.-N.; Djelal, N.; Burylo, L.; Wignacourt, J.-P.; Gulea, A.; Whitmire, K. H. Chem. Mater. 2014, 26, 6092−6103. (25) (a) Thurston, J. H.; Whitmire, K. H. Inorg. Chem. 2002, 41, 4194−4205. (b) Thurston, J. H.; Whitmire, K. H. Inorg. Chem. 2003, 42, 2014−2023. (26) Dikarev, E. V.; Zhang, H.; Li, B. J. Am. Chem. Soc. 2005, 127, 6156−6157. (27) Lieberman, C. M.; Navulla, A.; Zhang, H.; Filatov, A. S.; Dikarev, E. V. Inorg. Chem. 2014, 53, 4733−4738. (28) (a) Wengrovius, J. H.; Garbauskas, M. F.; Williams, E. A.; Going, R. C.; Donahue, P. E.; Smith, J. F. J. Am. Chem. Soc. 1986, 108, 982−989. (b) Kessler, V. G.; Gohil, S.; Parola, S. Dalton Trans. 2003, 544−550. (c) Lerach, J. O. Investigations into the Gas-Phase Rearrangements of Some Transition Metal β-Diketonate Complexes. M.S. Thesis, Youngstown State University, State College, OH, July 2008. (29) (a) Bondi, A. J. Phys. Chem. 1964, 68, 441−451. (b) Mantina, M.; Chamberlin, A. C.; Valero, R.; Cramer, C. J.; Truhlar, D. G. J. Phys. Chem. A 2009, 113, 5806−5812. (30) Stavila, V.; Dikarev, E. V. J. Organomet. Chem. 2009, 694, 2956− 2964. (31) Armelao, L.; Bandoli, G.; Casarin, M.; Depaoli, G.; Tondello, E.; Vittadini, A. Inorg. Chim. Acta 1998, 275, 340−348. (32) (a) Villamena, F. A.; Dickman, M. H.; Crist, D. R. Inorg. Chem. 1998, 37, 1446−1453. (b) Troyanov, S. I.; Gorbenko, O.-Y.; Bosak, A. A. Polyhedron 1999, 18, 3505−3509. (c) Bryant, J. R.; Taves, J. E.; Mayer, J. M. Inorg. Chem. 2002, 41, 2769−2776. (d) Parsons, S.; Winpenny, R.; Wood, P. A. Private Communication, 2004. (e) Zhang, H.; Li, B.; Sun, J.; Clérac, R.; Dikarev, E. V. Inorg. Chem. 2008, 47, 10046−10052. (33) Zhang, H.; Yang, J.-H.; Shpanchenko, R. V.; Abakumov, A. M.; Hadermann, J.; Clérac, R.; Dikarev, E. V. Inorg. Chem. 2009, 48, 8480− 8488. (34) (a) Sporer, C.; Wurst, K.; Ruiz-Molina, D.; Kopacka, H.; Veciana, J.; Jaitner, P. J. Organomet. Chem. 2003, 684, 44−49. (b) Yoshida, J.; Nishikiori, S.; Kuroda, R. Bull. Chem. Soc. Jpn. 2009, 82, 1377−1385. (c) Mehdi, H.; Binnemans, K.; Van Hecke, K.; Van Meervelt, L.; Nockemann, P. Chem. Commun. 2010, 46, 234−236. (d) Wang, Y.F.; Cheng, M. Q.; Wang, L. Y. Russ. J. Coord. Chem. 2011, 37, 111−116. (e) Zhang, P.; Gong, Y.; Lv, Y.; Guo, Y.; Wang, Y.; Wang, C.; Li, H. Chem. Commun. 2012, 48, 2334−2336. (35) (a) Fursova, E.-Y.; Kuznetsova, O. V.; Ovcharenko, V. I.; Romanenko, G. V.; Bogomyakov, A. S. Russ. Chem. Bull. 2008, 57, 1198−1205. (b) Mack, D. J.; Njardarson, J. T. Chem. Sci. 2012, 3, 3321−3325. (36) Bouwman, E.; Caulton, K. G.; Christou, G.; Folting, K.; Gasser, C.; Hendrickson, D. N.; Huffman, J. C.; Lobkovsky, E. B.; Martin, J. D.; Michel, P.; Tsai, H.-L.; Xue, Z. Inorg. Chem. 1993, 32, 3463−3470. (37) Wei, Z.; Filatov, A. S.; Dikarev, E. V. J. Am. Chem. Soc. 2013, 135, 12216−12219. E

DOI: 10.1021/acs.inorgchem.6b00209 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry (38) Wei, Z.; Dikarev, E. V. Unpublished results. (39) Navulla, A.; Tsirlin, A. A.; Abakumov, A. M.; Shpanchenko, R. V.; Zhang, H.; Dikarev, E. V. J. Am. Chem. Soc. 2011, 133, 692−694. (40) Zhang, H.; Li, B.; Dikarev, E. V. J. Cluster Sci. 2008, 19, 311− 321.

F

DOI: 10.1021/acs.inorgchem.6b00209 Inorg. Chem. XXXX, XXX, XXX−XXX