Next Article in Journal / Special Issue
Biomolecules Electrochemical Sensing Properties of a PMo11V@N-Doped Few Layer Graphene Nanocomposite
Previous Article in Journal
Stayin’ Alive—Organoplatinum Complexes
Previous Article in Special Issue
A Heteroaromatically Functionalized Hexamolybdate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Polyoxoniobates and Polyoxotantalates as Ligands—Revisited

by
Pavel A. Abramov
1,2,*,
Maxim N. Sokolov
1,2 and
Cristian Vicent
3
1
Nikolaev Institute of Inorganic Chemistry SB RAS, Novosibirsk 630090, Russia
2
Novosibirsk State University, Novosibirsk 630090, Russia
3
Serveis Centrals d'Instrumentació Científica, Universitat Jaume I, Av. Sos Baynat s/n, 12071 Castelló, Spain
*
Author to whom correspondence should be addressed.
Inorganics 2015, 3(2), 160-177; https://doi.org/10.3390/inorganics3020160
Submission received: 8 April 2015 / Revised: 7 May 2015 / Accepted: 8 May 2015 / Published: 20 May 2015
(This article belongs to the Special Issue Polyoxometalates)

Abstract

:
This short review summarizes our contribution to the coordination chemistry of noble metals (organometallic fragments of Rh, Ir, Ru and hydroxo Pt(IV)) and polyoxocomplexes of niobium and tantalum.

Graphical Abstract

1. Introduction

The group 5 (V, Nb, Ta) polyoxometalates (commonly defined as anionic polynuclear oxocomplexes) form two large but apparently distinct families, which, despite certain resemblance, show almost no overlap [1,2,3,4,5,6,7,8,9]. Vanadium(V) prefers aggregation into decaniobate [V10O28]6− and rearrangement of this very stable structure (search in CBSD yields about 150 structurally characterized salts of this anion with various organic cations) can be realized only in the presence of PO43− as template at low pH, producing a “bicapped” Keggin-type [PV14O42]9− anion [10,11,12,13,14,15]. Vanadium (IV) (but not Nb(IV) or Ta(IV)) can aggregate around differently-shaped anionic templates (like Cl, NCS etc.) producing a family of reduced polyoxovanadates built from tetragonal pyramids, with fascinating magnetic properties [16,17]. The chemistry of polyoxoniobates (PONb) and tantalates is much less studied and for a long time it was confined to preparations of alkali metal salts of hexaniobate and hexatantalate, [Nb6O19]8− and [Ta6O19]8−. The beginning of the modern era in this chemistry can be dated back to the works of Spinner, Pope and Stucky in the late 1960s–1980s [18,19,20,21,22,23,24], who obtained evidence for condensation of [Nb6O19]8− into larger entities, and discovered its ability to coordinate certain transition metals. This chemistry is quite different from polyvanadates and is dominated by Lindqvist hexametalates [M6O19]8−, in spite of the existence of the decaniobate, [Nb10O28]6− [25,26] and recently reported decatantalate [Ta10O28]6− [27]. By contrary, free [V6O19]8− is unknown, and the reason for this is perhaps that, if existed, it would have an extraordinarily high charge density (defined as the anionic charge of the POM divided by the number of non-hydrogen atoms of the POM). Among the known POM, the highest charge density (0.32) is achieved precisely for [M6O19]8− [9]. As V is smaller than Nb, [V6O19]8− would have the highest charge density per volume ratio. It can be stabilized only by coordination of four positively charged {Cp*Rh}2+ units [28,29].
The high charge density means that [M6O19]8− anions are expected to act as efficient ligands towards various metals. This is indeed so, and a few examples of coordination of [Nb6O19]8− to Mn(IV) and Ni(IV) were reported, as early as in 1967–1969 [18,19,20,21,22,23,24], and Co(III), reported only in 2011 [30], giving complexes of the type [M(Nb6O19)2]n, where two hexaniobate units (each acting as tridentate ligand) sandwich an octahedrally coordinated central metal ion. There is also a strong association of hexaniobate with alkali metal cations, even in water, where the existence of neutral [A8Nb6O19] (A = Rb, Cs) species was established by Small Angle X-ray Scattering (SAXS) [31]. Curiously enough, no tantalate analogues of [M(Nb6O19)2]n could be obtained [32]. It is doubtless highly desirable to expand this still rather modest list of existing complexes. However, strongly basic conditions, necessary to prevent hydrolysis of hexaniobate (and precipitation of niobium oxide), severely limit the choice of suitable metal precursors (they would simply give insoluble hydroxides). This drawback can be obviated in two different ways. First, in order to limit the hydrolysis and ensuing condensation, a part of the coordinating positions of a transition metal can be occupied by a polydentate (or polyhaptic) ligand (or simply by a strongly coordinated unidentate ligand, such as CO). This can be achieved for Co(III), Cr(III), Cu(II) and Ni(II) in the presence of polyamines [33,34,35,36,37]. Particularly attractive for this purpose are the isolobal fragments d6-fac-{ML3}, where ML3 stands for {M(CO)3}+ (M = Mn, Re), {(arene)Ru}2+, {Cp*Rh}2+, {Cp*Ir}2+. In fact, a decade ago Pope et al. used [M6O19]8− (M = Nb, Ta) as robust inorganic tridentate ligands in reactions with [Mn(CO)3(CH3CN)3]ClO4 or [Re(CO)5Br], which gave rise to trans-[(M'(CO)3)2M6O19]6− (M’ = Mn, Re; M = Nb, Ta) [38]. Proust et al. reported the syntheses of a series of complexes combining {(arene)Ru}2+ and [Nb6O19]8− with Ru/POM stoichiometries ranging from 1:1–4:1 [39]. Coordination of organometallic fragments, such as {(p-cymene)Ru}2+ or {Cp*Rh}2+, to POM adds extra possibilities for the synthesis of hybrid (i.e., combining chemically very different moieties within the same structure) complexes. This development reflects not only academic interest in such compounds, being a logical development of POM chemistry in general, but also is driven by the well-known catalytic properties of many POM complexes of transition metals, in particular, with noble metals [40].
The second way is to use a metal which forms stable well-defined and water soluble hydroxocomplexes, such as Pt(IV), for which both [Pt(OH)4(H2O)2] and [Pt(OH)6]2− exist as individual species [41]. In this case, a particular attraction for coordinating Pt(IV) to [Nb6O19]8− comes from the reports about use of PONb in photocatalysis. In this case, a photoactivator is necessary, which role is typically fulfilled by [PtCl6]2− [42]. From this point of view, direct coordination of Pt(IV) to PONb is very desirable. Pt(IV), being octahedral and sufficiently oxophilic, could either substitute a Nb atom in the hexaniobate structure, similar to [H2PtIVV9O28]5− which is simply a Pt(IV)-monosubstituted classical decavanadate [43,44], or be “grafted” as a capping atom. As we will see, the latter is the case.
In this short review, we focus on the recent advances in the field of group V (in particular, Nb and Ta) POM complexes, concentrating on the synthesis of Lindqvist-based POM with incorporated noble metal centers and their potential applications as electrocatalysts for water oxidation. The second part addresses related studies with heteropolyniobiate homologues coordinating noble metal fragments. The third part deals with the current use of analytical tools to unravel the integrity and reaction dynamics in the solution and selected examples are discussed.

2. Reactivity of [M6O19]8− towards Noble Metal Complexes

In this section, studies of reactivity of the Lindqvist [M6O19]8− (M = Ta and Nb) POM towards a noble metal complex are discussed, with an emphasis on finding suitable experimental conditions for noble metal grafting, the main structural features of the resulting solids, as well as their behavior in aqueous solution.

2.1. Reaction [M6O19]8− with {Cp*Rh}2+ and {Cp*Ir}2+

Heating of [Cp*RhCl2]2 with K7H[Nb6O19]·13H2O or Cs8[Ta6O19]·14H2O in water at 80–90 °C in 2:1 molar ratio yields yellow-orange solutions, from which K4[{Cp*Rh}2Nb6O19]·20H2O (1) and Cs4[{Cp*Rh}2Ta6O19]·18H2O (2) can be readily isolated in high yields (80% for Nb and 70% for Ta) [45]. Crystal structures of both complexes contain hybrid bi-capped organometallic-POM anions trans-[{Cp*Rh}2M6O19]4− (Figure 1) with the typical Lindqvist-type oxide metal core (metal octahedron with one μ6-, twelve μ2- and six terminal oxygen ligands) decorated at an opposite pair of the triangular {M3O3} faces with two capping organometallic fragments {Cp*Rh}2+. In both cases, exclusively trans-isomers crystallize. No cis-isomers were observed in solution or in solid. Both 1 and 2 produce basic solutions in water; 3 mM solutions have the pH value of 8.8, which is much lower in comparison to that produced by alkali metal salts of parent [M6O19]8− of comparable concentration (10.5–11.0). The ESI(-) mass spectra of aqueous solutions of 1 and 2 display prominent peaks assigned to the [{Cp*Rh}2M6O19 + 2H]2− dianions, accompanied by minor signals from the singly-charged [{Cp*Rh}2M6O19 + 3H] anions. However, the protonation degree observed by ESI-MS does not necessarily mirrors that in the solution because of the influence of the ESI process itself on the protonation [46]. For example, the most efficient ionization mechanism observed by ESI-MS for the decaniobate [Nb10O28]6− ion was the uptake of up to three protons, whereas solution studies indicated no protonation [26,47].
Figure 1. View of trans-[{Cp*Rh}2M6O19]4− in the crystal structure of K4[(Cp*Rh)2Nb6O19]·20H2O (1) and Cs4[(Cp*Rh)2Ta6O19]·18H2O (2) (ball and stick model).
Figure 1. View of trans-[{Cp*Rh}2M6O19]4− in the crystal structure of K4[(Cp*Rh)2Nb6O19]·20H2O (1) and Cs4[(Cp*Rh)2Ta6O19]·18H2O (2) (ball and stick model).
Inorganics 03 00160 g001
The reactions of [M6O19]8− with [Cp*RhCl2]2 in 1:1 molar ratio under the same conditions gave solutions, from which only poorly diffracting yellow solids could be isolated. These solids can be extracted with methanol yielding yellow solutions. The ESI mass spectra of the methanolic extracts were consistent with the presence of POM/Rh complexes featuring 1:1 stoichiometry of general formula ([{Cp*Rh}M6O19]6− (3, M = Nb; 4, M = Ta).
Grafting of {Cp*Ir}2+ fragments onto the [M6O19]8− (M = Nb, Ta) proceeds essentially in the same manner with [Cp*IrCl2]2 as the source of the {Cp*Ir}2+ fragment. Both 2:1 [{Cp*Ir}2Nb6O19]4− (5, M = Nb; 6, M = Ta) and 1:1 [{Cp*Ir}Nb6O19]6− (7, M = Nb; 8, M = Ta) complexes can be produced in high yields by varying the reagent ratio. In this case, we succeeded in isolation and X-ray characterization of potassium (M = Nb) and sodium (M = Nb, Ta) salts of 2:1 trans-[{Cp*Ir}2M6O19]4−, and sodium salts of the single-capped [{Cp*Ir}M6O19]6− complexes (both for Nb and Ta) (Figure 2).
Compound 54− was tested as electrocatalyst for water oxidation. The cyclic voltammogram of the trans-[{Cp*Ir}2Nb6O19]4− (54−) in 1 M Na2SO4 exhibited a strong anodic current due to water oxidation already at 0.8 V versus Ag/AgCl. The current (296 μA) at 1.6 V in the presence of the complex was 5.1 times higher than that (57.7 μA) for the blank solution.
Figure 2. Ball and stick model of single-capped [{Cp*Ir}M6O19]6− hybrid anions (M = Nb, Ta).
Figure 2. Ball and stick model of single-capped [{Cp*Ir}M6O19]6− hybrid anions (M = Nb, Ta).
Inorganics 03 00160 g002

2.2. Complexation of [Ta6O19]8− with {(C6H6)Ru}2+

Na8[Ta6O19]∙24.5H2O reacts with [(C6H6)RuCl2]2 (POM/Ru ratio 1:1) in water, yielding a yellow solution which contains single-capped [(C6H6)RuTa6O19]6− (96−) complex as the main product [48]. The progress of the reaction was monitored by ESI-MS as shown in Figure 3.
Figure 3. ESI(−) mass spectrum of aqueous solutions recorded at Uc = 10 V of the products of reaction between Na8[Ta6O19]·24.5H2O and [(C6H6)RuCl2]2 with POM / Ru ratio 1:1.
Figure 3. ESI(−) mass spectrum of aqueous solutions recorded at Uc = 10 V of the products of reaction between Na8[Ta6O19]·24.5H2O and [(C6H6)RuCl2]2 with POM / Ru ratio 1:1.
Inorganics 03 00160 g003
The ESI(−) mass spectrum of this reaction features the [(C6H6)RuTa6O19 + 5H] anion (m/z 1574.5) as the base peak, accompanied by a minor signal due to the doubly charged [(C6H6)RuTa6O19 + 4H]2− species (m/z 786.7). A minor peak at m/z 1751.5 was attributed to the double-capped [{(C6H6)Ru}2Ta6O19 + 3H] anion with two coordinated {(C6H6)Ru}2+ units. Evaporation of the reaction solution gave yellow plates of Na10[{(C6H6)RuTa6O18}2(μ-O)]·39.4H2O (10) in 80% yield, with unexpected [{(C6H6)RuTa6O18}2(μ-O)]10− (1010−) anion (see Scheme 1) built from two fragments {(C6H6)RuTa6O18} joined via a common linear µ2-O bridge. The formation of [{(C6H6)RuTa6O18}2(μ-O)]10− can be viewed as coordination-induced condensation of two hexatantalates according to the equilibrium:
[{(C6H6)Ru}Ta6O19]6− + H2O = [{(C6H6)Ru}Ta6O19H]5− + OH
2[{(C6H6)Ru}Ta6O19H]5− = [{(C6H6)RuTa6O18}2(μ-O)]10− + H2O
It is worth stressing that such condensation is unknown for free, non-coordinated hexatantalate. Perhaps, the reduced charge of the 1:1 complex [{(C6H6)Ru}Ta6O19]6− facilitates interaction between two negatively charged species; coordination of electron-accepting fragment {(C6H6)Ru}2+ may affect electronic density on the oxygen atoms of the POM part as well.
Decreasing the POM/Ru ratio to 1:2 leads to the bis-capped complex trans-[{(C6H6)Ru}2Ta6O19]4− (11), isolated as Na4[{(C6H6)Ru}2Ta6O19]∙20H2O (Na4([11]∙20H2O) in high yield. The ESI(−) mass spectrum displays the signal from the triply protonated [{(C6H6)Ru}2Ta6O19 + 3H] anion (m/z 1751.5) as the main peak, accompanied by a minor signal from the doubly protonated [{(C6H6)Ru}2Ta6O19 + 2H]2−dianion (m/z 875.3). A small amount of monocapped [{(C6H6)Ru}Ta6O19 + 5H] species coexists with the [{(C6H6)Ru}2Ta6O19]4− anion, thus suggesting a plausible dynamic equilibrium between mono- and bis-capped species. The relationship between the Ru/Ta species is summarized in Scheme 1:
Scheme 1. Formation of 1:1 and 1:2 complexes from [Ta6O19]8− and [(C6H6)RuCl2]2.
Scheme 1. Formation of 1:1 and 1:2 complexes from [Ta6O19]8− and [(C6H6)RuCl2]2.
Inorganics 03 00160 g009

2.3. Reaction [Nb6O19]8− and Pt4+

Hydrothermal reaction (150 °C) of [Pt(OH)4(H2O)2] with K7H[Nb6O19]·13H2O in 1:1 molar ratio gives a yellow solution, which shows only one signal in 195Pt NMR spectrum at 3189 ppm. Evaporation of this solution yields yellow crystals of Cs2K10[Nb6O19{Pt(OH)2}]2·13H2O (Cs2K10-12). Compound Cs2K10-12 was characterized by X-ray structural analysis (see Figure 4 left) [49].
Platinum niobate 12 is a unique dimeric complex, unprecedented in the PONb chemistry, and consists of two Pt(IV) centers coordinated each by three oxygen atoms from a {Nb3O3} face of a hexaniobate (d(Pt-O)av 2.028(13), 2.010(14), and 2.124(14) Å), a terminal oxygen of the adjacent [Nb6O19]8− anion (d(Pt-O) 1.998(11) Å), and bearing two terminal OH groups (d(Pt-O) = 2.015(16) and 1.994(14) Å). Freshly prepared aqueous solution of 12 gives 195Pt NMR spectrum with only one signal at 3189 ppm, the same as is observed in the mother liquor. The negative ESI mass spectrum also strongly suggests that the integrity of the dimeric [Nb6O19{Pt(OH)2}]212− polyanion is preserved in solution.
Figure 4. Structure of [Nb6O19{Pt(OH)2}]212− (1212−) (left) and structure of [Pt(Nb6O19)2]12−(1312−) (right), ball and stick models.
Figure 4. Structure of [Nb6O19{Pt(OH)2}]212− (1212−) (left) and structure of [Pt(Nb6O19)2]12−(1312−) (right), ball and stick models.
Inorganics 03 00160 g004
Reaction of [Pt(OH)4(H2O)2] and K7H[Nb6O19]·13H2O in the 1:2 molar ratio at 160–190 °C gives a yellow solution, which has two 195Pt NMR signals at 3189 and 3422 ppm with relative intensities 0.75/0.25. Slow evaporation of this solution in a glass beaker produces light-yellow crystals of Na2K10[Pt(Nb6O19)2]·18H2O (Na2K10-13) in 20% yield (Na+ comes from glass leaching). The sandwich-type anion [Pt(Nb6O19)2]12− (1312−) is a new (fourth) member of the well-known [M(Nb6O19)2]n+ (M = Mn, Ni, Co) family (Figure 4 right). All the Pt-O distances are equivalent, being 2.042(1) Å. The salt Na2K10-13 is isotypic with K6Na2[CoIIIH5(Nb6O19)2]·26.5H2O [30].
Further evaporation of the solution gives Cs2K10-12. A freshly prepared solution of Na2K10-13 demonstrates only one, slightly broadened, signal at 3422 ppm. Aged solutions give rise to increasingly intense signal at 3189 ppm, corresponding to the formation of [Nb6O19{Pt(OH)2}]212− (12). This observation implies a dynamics in solution, where [Pt(Nb6O19)2]12− (13) slowly converts into [Nb6O19{Pt(OH)2}]212− (12). Taking advantage of the versatility and high sensitivity of the ESI-MS technique for tracing solution speciation in the POM chemistry [47,50], we recorded ESI-MS spectrum of aqueous solution of Cs2K10-12 equilibrated for one week. ESI-MS results suggest coexistence of at least four different species in solution. Beside 12 and 13, the ESI-MS data suggest the presence of free Lindqvist [Nb6O19]8− polyanion and a species that corresponds to the [Nb6O19Pt(OH)3]7− anion whose structure might putatively correspond to capping of [Nb6O19]8− with a {Pt(OH)3}+ group. Remarkably, the platinum niobate Cs2K10-12 exhibits electrocatalytic activity for water oxidation in 0.1 M Na2SO4. Oxidation process generates a significant rise in the current for a catalytic process mediated by the complex. The current in the presence of the platinum niobate is several times higher than for the blank solution. This electrochemical behavior is similar to the reported for the nonatungstate complex with a tetrarhodium-oxo core [51].

3. Reactivity of Heteropolyniobates towards Noble Metal Complexes

In 1979, S.K. Ray reported that reaction of K7H[Nb6O19]·13H2O with H6TeO6 gave a heteropolyniobate formulated as K8H2[TeNb12O38]·13H2O [52]. This formula was deduced uniquely from elemental analysis. Such stoichiometry could imply that this compound would be analogous with the known heteropolyniobates [(Nb6O19)2M]n (M = Co, Ni, Mn, Pt) with octahedral TeVI sandwiched between two hexaniobates. We were curious if it were really so and repeated these experiments, only to find out that the purported “dodecaniobate” was in fact [(OH)TeNb5O18]6−, which can be regarded simply as a substituted hexaniobate with one {NbO}3+ vertex replaced with a {TeOH}5+ vertex. In fact, this complex had been already identified by W. Casey and co-workers, who prepared it from Nb2O5·xH2O, Me4NOH and H6TeO6 in methanolic solution [53]. Perhaps, the authors of [52] dealt with an approximately stoichiometric mixture of [(OH)TeNb5O18]6− and [Nb6O19]8−, which led them to the wrong conclusion about the composition of the product. Surface charge density of telluroniobate is lower than in the hexaniobate, which means that their reactivity might be different. In order to check the significance of this difference, reactivity studies towards noble metal complexes have been extended in our group to the telluroniobate [(OH)TeNb5O19]6−.

3.1. Reaction [(OH)TeNb5O19]6− with {Cp*Rh}2+ and {Cp*Ir}2+

We investigated reaction of [(OH)TeNb5O18]6− with [Cp*MCl2]2 as sources of {Cp*M}2+ (M = Rh, Ir). Despite the difference in the surface charge density between [(OH)TeNb5O18]6− and [Nb6O19]8−, formation of double-capped complexes takes place as easily as with hexaniobate. Crystal structure of Na2[{Cp*Rh}2Te(OH)Nb5O18]·24H2O (14) shows coordination of two {Cp*Rh}2+ fragments, with the Te atom localized in the central M4 plane of Lindqvist-type anion as illustrated in Figure 5. Two other metal positions are fully occupied by niobium. However, this disorder corresponds to four different orientations of a single isomer which is the only possible structure for the trans complex in the solid state.
Figure 5. Structure of [{Cp*Rh}2Te(OH)Nb5O19]2− anion (14), positions of Te atom with different occupancies are colored in green and purple (ball and stick model).
Figure 5. Structure of [{Cp*Rh}2Te(OH)Nb5O19]2− anion (14), positions of Te atom with different occupancies are colored in green and purple (ball and stick model).
Inorganics 03 00160 g005

3.2. Reaction [(OH)TeNb5O19]6− with {(C6H6)Ru}2+

Reactivity studies on [(OH)TeNb5O18]6− were also extended to [(C6H6)RuCl2]2 as source of the {(C6H6)Ru}2+. ESI-MS can be conveniently used to monitor the advance of the reaction. Grafting of one and two {(C6H6)Ru}2+ fragments can be immediately inferred from the characteristic m/z shifts to higher values, as illustrated in Figure 6. Moreover, the uptake of one or two Ru-containing fragments confers a new characteristic isotopic pattern due to multiple isotopes of the Ru at the natural abundance. Formation of single and double capped complexes [(OH)TeNb5O18{(C6H6)Ru}]4− (15) and [(OH)TeNb5O18{(C6H6)Ru}2]2− (16) takes place as smoothly as with hexaniobate. Characteristic peaks at m/z 539.1 and 628.1 correspond to formulations [(OH)TeNb5O18{(C6H6)Ru} + 2H]2− and [(OH)TeNb5O18{(C6H6)Ru}2]2−, respectively. Other species based upon 15 and 16 associated Na+ or K+ adducts are also evident.
Figure 6. ESI mass spectrum of the reaction mixture between [(OH)Nb5O18]6− and [(C6H6)RuCl2]2 with POM / Ru ratio 1:1.
Figure 6. ESI mass spectrum of the reaction mixture between [(OH)Nb5O18]6− and [(C6H6)RuCl2]2 with POM / Ru ratio 1:1.
Inorganics 03 00160 g006

3.3. Reaction [SiNb12O40]16− with {(C6H6)Ru}2+

Studies on the complexation with the siliconiobate [SiNb12O40]16− [54] are particularly relevant. Because of its high negative charge, it is expected to be much more “sticky” than well-known Group 6 Keggin complexes [EM12O40]n (n = 2–6) (M = Mo, W) which are often regarded as non-coordinating [55]. It can be anticipated that quite a large number of metal fragments could be grafted to [SiNb12O40]16−. Following this rationale, we undertook preliminary exploration of the reaction of [SiNb12O40]16− with [(C6H6)RuCl2]2, and ESI-MS results indicate that up to four {(C6H6)Ru}2+ fragments can be grafted to [SiNb12O40]16−, leading to the species formulated as [{(C6H6)Ru}4SiNb12O40]8− (17).

4. Analytical Tools to Study Solution Speciation in Nb and Ta POM Chemistry

One of the general features of the reactivity of POMs is their lability in aqueous solutions up to the point that such solutions can be regarded as virtual libraries of different building blocks, which equilibrate together and can form different structures depending on counter cations, pH and etc. The simplest case is protonation. The hexametalate anions [M6O19]8− (M = Nb, Ta) are very basic. For example, hexaniobate in water exists as [H3Nb6O19]5− (рН = 8), [H2Nb6O19]6− (рН = 10), [HNb6O19]7− (рН = 12), and [Nb6O19]8− (рН = 14) [56]. Similarly, for hexatantalate three protonated species, [HTa6O19]7−, [H2Ta6O19]6−, and [H3Ta6O19]5− were found [57,58,59]. Even in aqueous NBu4OH hexatantalate exists as [H2Ta6O19]6− [60]. Small-angle X-ray scattering (SAXS) studies of [Ta6O19]8− and [Nb6O19]8− in water have been reported and important differences in the solution behavior have been observed due to the distinctive propensity to aggregation with counteractions for M = Ta and Nb [61]. Uv-vis spectroscopy complemented with DFT calculations have been also reported [62]. More complicated processes connected with rearrangement of POM backbone can take place [63,64,65]. It is obvious that proving the identity between species existing in solid state and solution is a very important part of this chemistry. However, the study of solution behavior of polyniobates and polytantalates is a challenging task. These species are colorless, diamagnetic, show no redox behavior and are, under most conditions, in a dynamic equilibrium. In this section, useful analytical tools that can provide complementary signatures for the Nb and Ta POM characterization in solution are discussed. In particular, the usefulness of 1H NMR, 1H NMR Diffusion Ordered Spectroscopy (DOSY), capillary electrophoresis and ESI-MS in concert is underlined. An illustrative example is given to demonstrate the complementarily of all techniques to determine stoichiometry and solution dynamics.
The possibilities of NMR are restricted to the use of 17O, which unfortunately has very low natural abundance and sensitivity. It was used, for example, to follow the conversion of several niobates in water [66,67]. However, coordination of organometallic fragments to niobates and tantalates provides new opportunities for use of NMR for solution studies. In particular, structural information can be gained from 1H NMR signals of the organic ligands attached to the grafted heterometal, which would tell us about the overall symmetry of the compound. Moreover, it is possible to estimate diffusion coefficients from 1H NMR DOSY of the organic part that can be subsequently used to calculate molecular weight of a particular species. This 1H DOSY NMR technique is becoming a powerful tool for the determination of the size (hydrodynamic radius) and the molecular weight of inorganic clusters [68,69,70]. This technique has not been virtually used in the chemistry of Nb and Ta POM prior to our work.
Another useful tool for solution studies is capillary electrophoresis (CE). It is rarely used in the chemistry of POMs [71,72]. The possibilities opened up by the use of CE for POMs mixtures were studied in a few cases [73]. Although in these works CE revealed itself as a promising analytical tool, its potential for POM chemistry is far from being fully exploited. Combination of CE, HPLC and other methods (in particular ESI-MS and NMR) can provide sufficient information about quantitative and qualitative composition of solutions, numbers of POM species, and their stability and evolution with time.
Accurate determination of the formulae and structure of new compounds remains an important issue in POM chemistry. In this respect, use of soft ionization mass spectrometric techniques, and in particular, ESI techniques, is becoming widespread. Excellent reviews on the use of soft ionization MS techniques can be found elsewhere; for example, Cronin et al. illustrated the versatility of ESI and cryospray ionization techniques not only for characterization of final POM assemblies, but also to prove mechanisms of self-assembly of supramolecular POM-based architectures in solution [47,50]. Ohlin et al. reported an important work on this topic highlighting different approaches to obtain reliable information on the dynamic solution of POM on the basis of ESI-MS [47]. However, there are several potential limitations to using ESI-MS alone for the investigation of POM. This is the case when dynamic speciation involving hydroxy-bridged POM species, the pH value, ionic strength or the counteraction plays an important role. In addition, during the ESI process itself, the samples undergo dramatic changes in pH and concentration, which may perturb the speciation equilibrium. Moreover, due to the high charge state and molecular weight, the use of ESI-MS spectrometers equipped with high resolution analysers is important to facilitate unambiguous molecular compositions’ determination. In this part, we will discuss a case study, where we used ESI-MS, 1H DOSY NMR and CE techniques, combined together, in order to elucidate the solution behavior of the {(C6H6)Ru}2+/[Ta6O19]8− system to afford compounds 9 and 10 [48].

4.1. {(C6H6)Ru}2+/[Ta6O19]8− System—Case Study

The 1H DOSY NMR experiment performed on a sample of Na10[10]·39.4H2O in D2O (see Figure 7) yielded the self-diffusion coefficients D = 315 ± 10 µm²·s−1 and D = 325 ± 10 µm²·s−1, respectively, for the 1H NMR signals located at 5.85 ppm and 5.94 ppm. The Stokes-Einstein relation D = k T 6 π π R H (which describes the diffusion coefficient D for spherical objects, where k is the Boltzmann constant, T the temperature taken equal to 300 K in this study, η of the solvent viscosity taken equal to that of water at 300 K (1.002 Pa.s) and RH the hydrodynamic radius) gave hydrodynamic radii of 6.75–6.96 Å which are compatible with the radii expected for solvated monomeric species [{(C6H6)Ru}Ta6O19]6− and [{(C6H6)Ru}2Ta6O19]4− but not with the dimeric [{(C6H6)RuTa6O18}2(μ-O)]10−.
Figure 7. 1H DOSY NMR spectra of Na10-10 in D2O (left) and 1H DOSY NMR spectra of Na4-9 in D2O (right).
Figure 7. 1H DOSY NMR spectra of Na10-10 in D2O (left) and 1H DOSY NMR spectra of Na4-9 in D2O (right).
Inorganics 03 00160 g007
The 13C NMR spectrum of Na4[{(C6H6)Ru}2Ta6O19]∙20H2O (Na4-9) in D2O exhibits two signals at 81.95 ppm and 81.82 ppm, with the complementary set of signals in the 1H NMR spectrum at 5.97 (6%) ppm and 5.92 (93%) ppm and 5.85 (≤1%). This pattern corresponds to a mixture of cis (6%) and trans (93%) isomers of [{(C6H6)Ru}2Ta6O19]4−, similarly to what was observed for [{(p-cymene)Ru}2Nb6O19]4−6, while the minor peak at 5.85 ppm is to be attributed to the 1:1 complex [{(C6H6)Ru}Ta6O19]6−. The 1H DOSY NMR experiment gives the self-diffusion coefficients D = 315 ± 10 µm² s−1 and D = 325 ± 10 µm² s−1 for the signals at 5.92 and 5.97 ppm, respectively, whereas the low intensity of the minor peak has prevented accurate determination of the D value. These values correspond to hydrodynamic radii in the 0.675–0.696 nm range. This is identical to that obtained for the solutions of Na10[{(C6H6)RuTa6O18}2(μ-O)]·39.4H2O and agree well with the attribution of 1H NMR peaks to cis-[{(C6H6)Ru}2Ta6O19]4− (5.97 ppm), trans-[{(C6H6)Ru}2Ta6O19]4− (5.92 ppm), and [{(C6H6)Ru}Ta6O19]6− (5.85 ppm). These results necessarily imply two equilibria. The first corresponds to cis-trans isomerization process, strongly displaced towards the trans isomer. The latter corresponds to the de-coordination of a {(C6H6)Ru}2+ moiety. This hypothesis is supported by the observation of a small broad peak at 5.4 ppm, attributed to [(C6H6)2Ru2(OH)3]+ (see Equation 2).
trans-[{(C6H6)Ru}2Ta6O19]4− = cis-[{(C6H6)Ru}2Ta6O19]4−
2[{(C6H6)Ru}2Ta6O19]4− + 3OH + = 2[{(C6H6)Ru}Ta6O19]6− + [(C6H6)2Ru2(μ-OH)3]+
ESI(−) mass analysis of Na10-10 in water reveals only the presence of the monomeric 96- species, thus indicating that the dimerization depicted in Eq. 1 is a reversible process ongoing from the solution to the solid state and vice versa. The electrophorograms recorded for aqueous solutions of Na10-10 and Na4-9 are presented in Figure 8. The solutions of Na10-10, to which some amount of Na4-9 has been added as reference, give two well-resolved peaks (Figure 7). The second peak corresponds to 9, which was confirmed by UV-spectroscopy and by the standard addition method. The first peak must therefore correspond to the monomeric [{(C6H6)Ru}Ta6O19]6− species (or its dominant protonated form), because the signal from the dimer with its larger size and charge, should appear after the signal of 10, and with a longer retention time. Thus, combination of ESI-MS and CE gives coherent evidence that in aqueous solutions [{(C6H6)RuTa6O18}2(μ-O)]1° dissociates into the monomeric [{(C6H6)Ru}Ta6O19]6−.
Figure 8. The electrophorogram of complex Na4-9 (up). The electrophorogram of complex Na10-10 solution with addition of complex Na4-9 as reference (down). Fused-silica capillary 50µm i.d. × 75 sm; sample buffer: 7.5·10−3М borate (pH 9.18); running buffer: 7,5·10−2 M borate (pH 9.18).
Figure 8. The electrophorogram of complex Na4-9 (up). The electrophorogram of complex Na10-10 solution with addition of complex Na4-9 as reference (down). Fused-silica capillary 50µm i.d. × 75 sm; sample buffer: 7.5·10−3М borate (pH 9.18); running buffer: 7,5·10−2 M borate (pH 9.18).
Inorganics 03 00160 g008

5. Conclusions

The chemistry of niobium and tantalum polyoxoanions still belongs to a little researched area of inorganic chemistry. Many unusual complexes still remain hidden, and isolation of principally new [(Ta6O18)2(µ-O)]14− backbone is a tantalyzing hint at this potential richness. In this short review, we summarized our recent research focused on the coordination of noble metals to hexametalates. Grafting of {Cp*M}2+ (M = Rh, Ir) on the surface of Lindqvist-type hexametalates (including [(OH)TeNb5O19]6−) was studied with full characterization of the hybrid complexes in solid state and solution. Coordination of {(C6H6)Ru}2+ to [Ta6O19]8− gives unprecedented [({(C6H6)Ru}Ta6O18)2(µ-O)]10−. Complexation of Pt(IV) with [Nb6O19]8− was studied with different techniques. New dimeric complex [(Nb6O19{Pt(OH)2})2]12− is obtained when hexaniobate reacts with Pt(IV) in 1:1 molar ratio, while increasing the Pt/Nb6 ratio to 2:1 gives sandwich-type complex [Pt(Nb6O19)2]12−. We used different, both traditional and non-traditional, analytical tools for studying the solution behavior of individual reaction products and reaction mixtures, a challenging task for the systems where few analytical methods are readily available. The 1H, 13C and DOSY NMR techniques combined with ESI-MS and CE highlight are adequate tools for characterization of complex solutions. An illustrative example dealing with the {(C6H6)Ru}2+/[Ta6O19]8− system demonstrates complementarity of all these techniques for determination of stoichiometry and solution dynamics, an approach both promising and highly recommendable to trace unambiguously the identity of large inorganic systems [74].

Acknowledgments

Authors thanks SCIC of the UJI and all members of our Russian-Spanish-French team for very fruitful collaboration. This work was supported by Russian Scientific Foundation (RScF 14-13-00645).

Author Contributions

The manuscript was written with equal contributions from all authors. MNS owns the general idea and plan of the publication, and he wrote parts 1 and 3, as well as Conclusions. PAA authored part 2, and CV was the main contributor to part 4.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Livage, J. Synthesis of polyoxovanadates via “chimie douce”. Coord. Chem. Rev. 1998, 178–189, 999–1018. [Google Scholar] [CrossRef]
  2. Hayashi, Y. Hetero and lacunary polyoxovanadate chemistry: Synthesis, reactivity and structural aspects. Coord. Chem. Rev. 2011, 255, 2270–2280. [Google Scholar] [CrossRef]
  3. Müller, A.; Sessoli, R.; Krickemeyer, E.; Bögge, H.; Meyer, J.; Gatteschi, D.; Pardi, L.; Westphal, J.; Hovemeier, K.; Rohlfing, R.; et al. Polyoxovanadates:  High-Nuclearity Spin Clusters with Interesting Host–Guest Systems and Different Electron Populations. Synthesis, Spin Organization, Magnetochemistry, and Spectroscopic Studies. Inorg. Chem. 1997, 36, 5239–5250. [Google Scholar] [CrossRef]
  4. Breen, J.M.; Zhang, L.; Clement, R.; Schmitt, W. Hybrid Polyoxovanadates: Anion-Influenced Formation of Nanoscopic Cages and Supramolecular Assemblies of Asymmetric Clusters. Inorg. Chem. 2012, 51, 19–21. [Google Scholar] [CrossRef] [PubMed]
  5. Zhang, L.; Schmitt, W. From Platonic Templates to Archimedean Solids: Successive Construction of Nanoscopic {V16As8}, {V16As10}, {V20As8}, and {V24As8} Polyoxovanadate Cages. J. Am. Chem. Soc. 2011, 133, 11240–11248. [Google Scholar] [CrossRef] [PubMed]
  6. Gatteschi, D.; Sessoli, R.; Müller, A.; Kögerler, P. Polyoxometalate chemistry: A source for unusual spin topologies. In Polyoxometalate Chemistry; Pope, M.T., Müller, A., Eds.; Kluwer: Dordrecht, The Netherlands, 2001; p. 319. [Google Scholar]
  7. Choi, J.; Sanderson, L.A.W.; Musfeldt, J.L.; Ellern, A.; Kögerler, P. Optical properties of the molecule-based magnet K6[V15As6O42(H2O)]·8H2O. Phys. Rev. B 2003, 68, 064412. [Google Scholar] [CrossRef]
  8. Gatteschi, D.; Pardi, L.; Barra, A.L.; Müller, A. Polyoxovanadates: The Missing Link between Simple Paramagnets and Bulk Magnets? Mol. Eng. 1993, 3, 157–169. [Google Scholar] [CrossRef]
  9. Nymann, M. Polyoxoniobate chemistry in the 21st century. Dalton Trans. 2011, 40, 8049–8058. [Google Scholar] [CrossRef] [PubMed]
  10. Kato, R.; Kobayashi, A.; Sasaki, Yu. 1:14 Heteropolyvanadate of phosphorus: preparation and structure. J. Am. Chem. Soc. 1980, 102, 6571–6572. [Google Scholar] [CrossRef]
  11. Kato, R.; Kobayashi, A.; Sasaki, Yu. The heteropolyvanadate of phosphorus. Crystallographic and NMR studies. Inorg. Chem. 1982, 21, 240–246. [Google Scholar] [CrossRef]
  12. Nomiya, K.; Kato, K.; Miwa, M. Preparation and spectrochemical properties of soluble vanadophosphate polyanions with bicapped-Keggin structure. Polyhedron 1986, 5, 811–813. [Google Scholar] [CrossRef]
  13. Khan, M.I.; Zubieta, J.; Toscano, P. Protonation sites in a heteropolyvanadate of phosphorus: X-ray crystal structure of (Me3NH)4(NH4)[H4PV14O42]. Inorg. Chim. Acta 1992, 193, 17–20. [Google Scholar] [CrossRef]
  14. Nakamura, S.; Yamawaki, T.; Kusaka, K.; Otsuka, T.; Ozeki, T. Hydrogen-bond Networks Involving Protonated Bicapped-Keggin Tetradecavanadophosphate Anions. J. Clust. Sci. 2006, 17, 245–256. [Google Scholar] [CrossRef]
  15. Grabau, M.; Forster, J.; Heussner, K.; Streb, C. Synthesis and Theoretical Hirshfeld Analysis of a Supramolecular Heteropolyoxovanadate Architecture. Eur. J. Inorg. Chem. 2011, 1719–1724. [Google Scholar] [CrossRef]
  16. Fang, X.; Kögerler, P.; Speldrich, M.; Schilder, H.; Luban, M. A polyoxometalate-based single-molecule magnet with an S = 21/2 ground state. Chem. Commun. 2012, 48, 1218–1220. [Google Scholar] [CrossRef]
  17. Monakhov, K.Yu.; Linnenberg, O.; Kozłowski, P.; van Leusen, J.; Besson, C.; Secker, T.; Ellern, A.; López, X.; Poblet, J.M.; Kögerler, P. Supramolecular Recognition Influences Magnetism in [X@HVIV8VV14O54]6− Self-Assemblies with Symmetry-Breaking Guest Anions. Chem. Eur. J. 2015, 21, 2387–2397. [Google Scholar] [CrossRef] [PubMed]
  18. Dale, B.W.; Pope, M.T. The heteropoly-12-niobomanganate(IV) anion. J. Chem. Soc. Chem. Commun. 1967, 792–792. [Google Scholar]
  19. Flynn, C.M., Jr.; Stucky, G.D. Heteropolyniobate complexes of manganese(IV) and nickel(IV). Inorg. Chem. 1969, 8, 332–334. [Google Scholar] [CrossRef]
  20. Flynn, C.M., Jr.; Stucky, G.D. Crystal structure of sodium 12-niobomanganate(IV),Na12MnNb12O38·50H2O. Inorg. Chem. 1969, 8, 335–344. [Google Scholar] [CrossRef]
  21. Kheddar, N.; Spinner, B. Constitution et domaines de stabilite des oxaloniobates en solution aqueuse. Bull. Soc. Chim. Fr. 1972, 2, 502–506. (In French) [Google Scholar]
  22. Marty, A.; Abdmeziem, K.; Spinner, B. Une nouvelle condensation pour des sels du niobium V: les nonaniobates de tetramethyl et de tetraethylammonium. Comptes Rendus C. 1976, 283, 285–288. [Google Scholar]
  23. Marty, A.; Abdmeziem, K.; Spinner, B. De noveaux isopolyanions du niobium V: comportement en solution aqueuse des nonaniobates de tetramethyl et tetraethylammonium. Bull. Soc. Chim. Fr. 1977, 3–4, 231–238. (In French) [Google Scholar]
  24. Goiffon, A.; Philippot, E.; Maurin, M. Structure cristalline du niobate 7/6 de sodium (Na7)(H3O)Nb6O19·14H2O. Rev. Chim. Miner. 1980, 17, 466–476. [Google Scholar]
  25. Graeber, E.J.; Morosin, B. The molecular configuration of the decaniobate ion (Nb10O286−). Acta Cryst. 1977, 33, 2137–2143. [Google Scholar] [CrossRef]
  26. Villa, E.M.; Ohlin, C.A.; Balogh, E.; Anderson, T.M.; Nyman, M.D.; Casey, W.H. Reaction Dynamics of the Decaniobate Ion [HxNb10O28](6−x)− in Water. Angew. Chem. Int. Ed. 2008, 47, 4844–4846. [Google Scholar] [CrossRef]
  27. Matsumoto, M.; Ozawa, Y.; Yagasaki, A.; Zhe, Y. Decatantalate—The Last Member of the Group 5 Decametalate Family. Inorg. Chem. 2013, 52, 7825–7827. [Google Scholar] [CrossRef] [PubMed]
  28. Hayashi, Y.; Ozawa, Y.; Isobe, K. The first vanadate hexamer capped by 4 cyclopentadienyl-rhodium or cyclopentadienyl-iridium groups. Chem. Lett. 1989, 425–428. [Google Scholar]
  29. Chae, H.K.; Klemperer, W.G.; Day, V.W. Organometal hydroxide route to [(C5Me5)Rh]4(V6O19). Inorg. Chem. 1989, 28, 1423–1424. [Google Scholar]
  30. Ma, P.T.; Chen, G.; Wang, G.; Wang, J.P. Cobalt—Sandwiched lindqvist hexaniobate dimer [Co(III)H5(Nb6O19)2]8−. Rus. J. Coord. Chem. 2011, 37, 772–775. [Google Scholar] [CrossRef]
  31. Antonio, M.R.; Nyman, M.; Anderson, T.M. Direct Observation of Contact Ion-Pair Formation in Aqueous Solution. Angew. Chem. Int. Ed. 2009, 48, 6136–6140. [Google Scholar] [CrossRef]
  32. Dale, B.W.; Buckley, J.M.; Pope, M.T. Heteropoly-niobates and -tantalates containing manganese(IV). J. Chem. Soc. A. 1969, 301–304. [Google Scholar] [CrossRef]
  33. Flynn, C.M., Jr.; Stucky, G.D. Sodium 6-niobo(ethylenediamine)cobaltate(III) and its chromate(III) analog. Inorg. Chem. 1969, 8, 178–180. [Google Scholar] [CrossRef]
  34. Bontchev, R.P.; Nyman, M. Evolution of Polyoxoniobate Cluster Anions. Angew. Chem., Int. Ed. 2006, 45, 6670–6672. [Google Scholar] [CrossRef]
  35. Niu, J.Y.; Ma, P.T.; Niu, H.Y.; Li, J.; Zhao, J.W.; Song, Y.; Wang, J.P. Giant Polyniobate Clusters Based on [Nb7O22]9− Units Derived from a Nb6O19 Precursor. Chem.–Eur. J. 2007, 13, 8739–8748. [Google Scholar] [CrossRef]
  36. Chen, G.; Ma, P.T.; Wang, J.P.; Niu, J.Y. A new organic–inorganic hybrid polyoxoniobate based on Lindqvist-type anion and nickel complex. J. Coord. Chem. 2010, 63, 3753–3763. [Google Scholar] [CrossRef]
  37. Hegetschweiler, K.; Finn, R.C.; Rarig, R.S.; Sander, J.; Steinhauser, S.; Worle, M.; Zubieta, J. Surface complexation of [Nb6O19]8− with NiII: solvothermal synthesis and X-ray structural characterization of two novel heterometallic Ni-Nb-polyoxometalates. J. Inorg. Chim. Acta 2002, 337, 39–47. [Google Scholar] [CrossRef]
  38. Dickman, M.H.; Pope, M.T. Robust, Alkali-Stable, Triscarbonyl Metal Derivatives of Hexametalate Anions, [M6O19{M‘(CO)3}n](8−n)− (M = Nb, Ta; M‘ = Mn, Re; n = 1, 2). Inorg. Chem. 2001, 40, 2582–2586. [Google Scholar] [CrossRef] [PubMed]
  39. Laurencin, D.; Thouvenot, R.; Boubekeur, K.; Proust, A. Synthesis and reactivity of {Ru(p-cymene)}2+ derivatives of [Nb6O19]8−: A rational approach towards fluxional organometallic derivatives of polyoxometalates. Dalton Trans. 2007, 1334–1345. [Google Scholar] [CrossRef]
  40. Hill, C.L.; Prosser-McCartha, C.M. Homogeneous catalysis by transition metal oxygen anion clusters. Coord. Chem. Rev. 1995, 143, 407–455. [Google Scholar] [CrossRef]
  41. Vasilchenko, D.; Tkachev, S.; Baidina, I.; Korenev, S. Speciation of Platinum(IV) in Nitric Acid Solutions. Inorg. Chem. 2013, 52, 10532–10541. [Google Scholar] [CrossRef] [PubMed]
  42. Huang, P.; Qin, C.; Su, Z.-M.; Xing, Y.; Lang, X.-L.; Shao, K.-Z.; Lan, Y.-Q.; Wang, E.-B. Self-Assembly and Photocatalytic Properties of Polyoxoniobates: {Nb24O72}, {Nb32O96}, and {K12Nb96O288} Clusters. J. Am. Chem. Soc. 2012, 134, 14004–14010. [Google Scholar] [CrossRef] [PubMed]
  43. Lee, U.; Joo, H.C.; Park, K.M.; Mal, S.S.; Kortz, U.; Keita, B.; Nadjo, V. Facile Incorporation of Platinum(IV) into Polyoxometalate Frameworks: Preparation of [H2PtIVV9O28]5− and Characterization by 195Pt NMR Spectroscopy. Angew. Chem. Int. Ed. 2008, 47, 793–796. [Google Scholar] [CrossRef]
  44. Kortz, U.; Lee, U.; Joo, H.C.; Park, K.M.; Mal, S.S.; Dickman, M.H.; Jameson, G.B. Platinum-Containing Polyoxometalates. Angew. Chem. Int. Ed. 2008, 47, 9383–9384. [Google Scholar] [CrossRef]
  45. Abramov, P.A.; Sokolov, M.N.; Virovets, A.V.; Floquet, S.; Haouas, M.; Taulelle, F.; Cadot, E.; Vicent, C.; Fedin, V. Grafting {Cp*Rh}2+ on the surface of Nb and Ta Lindqvist-type POM. Dalton Trans. 2015, 44, 2234–2239. [Google Scholar] [CrossRef] [PubMed]
  46. Di Marco, V.B.; Bombi, G.G. Electrospray mass spectrometry (ESI-MS) in the study of metal-ligand solution equilibria. Mass Spectrom. Rev. 2006, 25, 347–379. [Google Scholar] [CrossRef] [PubMed]
  47. Ohlin, C.A. Reaction Dynamics and Solution Chemistry of Polyoxometalates by Electrospray Ionization Mass Spectrometry. Chem.-Asian J. 2012, 7, 262–270. [Google Scholar] [CrossRef] [PubMed]
  48. Abramov, P.A.; Sokolov, M.N.; Floquet, S.; Haouas, V.; Taulelle, V.; Cadot, V.; Peresypkina, E.V.; Virovets, A.V.; Vicent, C.; Kompankov, N.B.; Zhdanov, A.A.; Shuvaeva, O.V.; Fedin, V.P. Coordination-Induced Condensation of [Ta6O19]8–: Synthesis and Structure of [{(C6H6)Ru}2Ta6O19]4– and [{(C6H6)RuTa6O18}2(μ-O)]10−. Inorg. Chem. 2014, 53, 12791–12798. [Google Scholar] [CrossRef] [PubMed]
  49. Abramov, P.A.; Vicent, C.; Kompankov, N.B.; Gushchin, A.L.; Sokolov, M.N. Platinum polyoxoniobates. Chem. Commun. 2015, 51, 4021–4023. [Google Scholar] [CrossRef]
  50. Miras, H.N.; Wilson, E.F.; Cronin, L. Unravelling the complexities of inorganic and supramolecular self-assembly in solution with electrospray and cryospray mass spectrometry. Chem. Commun. 2009, 1297–1311. [Google Scholar] [CrossRef]
  51. Sokolov, M.N.; Adonin, S.A.; Abramov, P.A.; Mainichev, D.A.; Zakharchuk, N.F.; Fedin, V.P. Self-assembly of polyoxotungstate with tetrarhodium-core: Synthesis, structure and 183W NMR studies. Chem. Commun. 2012, 48, 6666–6668. [Google Scholar] [CrossRef]
  52. Ray, S.K. Synthesis of a Te dodecaniobate. J. Ind. Chem. Soc. 1976, 53, 1238–1239. [Google Scholar]
  53. Son, J.-H.; Wang, J.; Osterloh, F.E.; Yu, P.; Casey, W.H. A tellurium-substituted Lindqvist-type polyoxoniobate showing high H2 evolution catalyzed by tellurium nanowires via photodecomposition. Chem. Commun. 2014, 50, 836–838. [Google Scholar] [CrossRef]
  54. Nyman, M.; Bonhomme, F.; Alam, T.M.; Parise, J.B.; Vaughan, G.M.B. [SiNb12O40]16− and [GeNb12O40]16−: Highly Charged Keggin Ions with Sticky Surfaces. Angew. Chem., Int. Ed. 2004, 43, 2787–2792. [Google Scholar] [CrossRef]
  55. Anyushin, A.V.; Smolentsev, A.I.; Mainichev, D.A.; Vicent, C.; Gushchin, A.L.; Sokolov, M.N.; Fedin, V.P. Synthesis and characterization of a new Keggin anion: [BeW12O40]6−. Chem. Commun. 2014, 50, 9083. [Google Scholar] [CrossRef]
  56. Black, J.R.; Nyman, M.; Casey, W.H. Kinetics of 17O-exchange reactions in aqueous metal-oxo nanoclusters. Geochim. Cosmochim. Acta 2006, 70, A53–A53. [Google Scholar]
  57. Nelson, W.H.; Tobias, R.S. Polyanions of the Transition Metals. II. Ultracentrifugation of Alkaline Tantalum(V) Solutions; Comparison with Light Scattering. Inorg. Chem. 1964, 3, 653–658. [Google Scholar] [CrossRef]
  58. Spinner, B.; Kheddar, N. Nouveaux isopolyanions du tantale V. Comp. Rend. 1969, 268C, 1108–1111. [Google Scholar]
  59. Arana, G.; Etxebarria, N.; Fernandez, L.A.; Madariaga, J.M. Hydrolysis of Nb(V) and Ta(V) in aqueous KCl at 25 °C. Part II: Construction of a thermodynamic model for Ta(V). J. Solution Chem. 1995, 24, 611–622. [Google Scholar] [CrossRef]
  60. Matsumoto, M.; Ozawa, Y.; Yagasaki, A. Which is the most basic oxygen in [Ta6O19]8−?—Synthesis and structural characterization of [H2Ta6O19]6−. Inorg. Chem. Commun 2011, 14, 115–117. [Google Scholar] [CrossRef]
  61. Fullmer, L.B.; Molina, P.I.; Antonio, M.R.; Nyman, M. Contrasting ion-association behaviour of Ta and Nb polyoxometalates. Dalton Trans. 2014, 43, 15295–15299. [Google Scholar] [CrossRef] [PubMed]
  62. Deblonde, G.J.P.; Moncomble, A.; Cote, G.; Belair, S.; Chagnes, A. Experimental and computational exploration of the UV-visible properties of hexaniobate and hexatantalate ions. RSC Adv. 2015, 5, 7619–7627. [Google Scholar] [CrossRef]
  63. Maekawa, M.; Ozawa, Y.; Yagasaki, A. Icosaniobate:  A New Member of the Isoniobate Family. Inorg. Chem. 2006, 45, 9608–9609. [Google Scholar]
  64. Matsumoto, M.; Ozawa, Y.; Yagasaki, A. Reversible dimerization of decaniobate. Polyhedron 2010, 29, 2196–2201. [Google Scholar] [CrossRef]
  65. Tsunashima, R.; Long, D.L.; Miras, H.N.; Gabb, D.; Pradeep, C.P.; Cronin, L. The Construction of High-Nuclearity Isopolyoxoniobates with Pentagonal Building Blocks: [HNb27O76]16− and [H10Nb31O93(CO3)]23−. Angew. Chem. Int. Ed. 2010, 49, 113–116. [Google Scholar] [CrossRef]
  66. Villa, E.M.; Ohlin, C.A.; Rustad, J.R.; Casey, W.H. Isotope-Exchange Dynamics in Isostructural Decametalates with Profound Differences in Reactivity. J. Am. Chem. Soc. 2009, 131, 16488–16492. [Google Scholar] [CrossRef]
  67. Villa, E.M.; Ohlin, C.A.; Casey, W.H. Oxygen-Isotope Exchange Rates for Three Isostructural Polyoxometalate Ions. J. Am. Chem. Soc. 2010, 132, 5264–5272. [Google Scholar] [CrossRef] [PubMed]
  68. Bannani, F.; Floquet, S.; Leclerc-Laronze, N.; Haouas, M.; Taulelle, F.; Marrot, J.; Kögerler, P.; Cadot, E. Cubic Box versus Spheroidal Capsule Built from Defect and Intact Pentagonal Units. J. Am. Chem. Soc. 2012, 134, 19342–19345. [Google Scholar] [CrossRef] [PubMed]
  69. Van Lokeren, L.; Cartuyvels, E.; Absillis, G.; Willema, R.; Parac-Vogt, T.N. Phosphoesterase activity of polyoxomolybdates: as a tool for obtaining insights into the reactivity of polyoxometalate clusters. Chem. Commun. 2008, 2774–2776. [Google Scholar] [CrossRef]
  70. Lemonnier, J.-F.; Floquet, S.; Kachmar, A.; Rohmer, M.-M.; Bénard, M.; Marrot, J.; Terazzi, E.; Piguet, C.; Cadot, E. Host–guest adaptability within oxothiomolybdenum wheels: structures, studies in solution and DFT calculations. Dalton Trans. 2007, 3043–3054. [Google Scholar] [CrossRef]
  71. Sakurai, N.; Kadohata, K.; Ichinose, N. Application of high-speed liquid chromatography using solvent extraction of the molybdoheteropoly yellow to the determination of microamounts of phosphorus in waste waters. Fresenius' Z. Anal. Chem. 1983, 314, 634–637. [Google Scholar] [CrossRef]
  72. Kirk, A.D.; Riske, W.; Lyon, D.K.; Rapko, B.; Finke, R.G. Rapid, high-resolution, reversed-phase HPLC separation of highly charged polyoxometalates using ion-interaction reagents and competing ions. Inorg. Chem. 1989, 28, 792–797. [Google Scholar] [CrossRef]
  73. Hettiarachichi, K.; Ha, Y.; Tran, T.; Cheung, A.P. Application of HPLC and CZE to the analysis of polyoxometalates. J. Pharm. Biomed. An. 1995, 13, 515–523. [Google Scholar] [CrossRef]
  74. Oliveri, A.F.; Elliott, E.W.; Carnes, M.E.; Hutchison, J.E.; Johnson, D.W. Elucidating Inorganic Nanoscale Species in Solution: Complementary and Corroborative Approaches. ChemPhysChem 2013, 14, 2655–2661. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Abramov, P.A.; Sokolov, M.N.; Vicent, C. Polyoxoniobates and Polyoxotantalates as Ligands—Revisited. Inorganics 2015, 3, 160-177. https://doi.org/10.3390/inorganics3020160

AMA Style

Abramov PA, Sokolov MN, Vicent C. Polyoxoniobates and Polyoxotantalates as Ligands—Revisited. Inorganics. 2015; 3(2):160-177. https://doi.org/10.3390/inorganics3020160

Chicago/Turabian Style

Abramov, Pavel A., Maxim N. Sokolov, and Cristian Vicent. 2015. "Polyoxoniobates and Polyoxotantalates as Ligands—Revisited" Inorganics 3, no. 2: 160-177. https://doi.org/10.3390/inorganics3020160

Article Metrics

Back to TopTop