Next Article in Journal
Long-Distance High-Power Wireless Optical Energy Transmission Based on VECSELs
Next Article in Special Issue
Mixed Metal Phosphonates: Structure and Proton Conduction Manipulation through Various Alkaline Earth Metal Ions
Previous Article in Journal
New Voltammetric Sensor Based on LDH and H2O2 for L-Proline Determination in Red and White Wines
Previous Article in Special Issue
Coordination of a Pyrazole Functionalized Acetylacetone to the Coinage Metal Cations: An Unexpected Packing Similarity and a Trinuclear CuII/AuI Complex
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New As-Rich Arsenato-Polyoxovanadate Clusters: Solvothermal Synthesis and Selected Properties of [V6IVAs8IIIO26]4− Cluster-Containing Compounds

Institute for Inorganic Chemistry, Kiel University, Max-Eyth-Str. 2, D-24118 Kiel, Germany
*
Author to whom correspondence should be addressed.
Submission received: 16 September 2022 / Revised: 13 October 2022 / Accepted: 14 October 2022 / Published: 18 October 2022
(This article belongs to the Special Issue Mixed-Metal Coordination Polymers)

Abstract

:
Three new arsenato-polyoxovanadates with the composition [M(en)3]2[V6As8O26] (M = Co2+ (I), Zn2+ (II), and Cd2+ (III)) were synthesized under solvothermal conditions in high yields, thus significantly enhancing the knowledge of As-rich polyoxovanadate cluster chemistry. The compounds are isostructural and feature the very rare [V6IVAs8IIIO26]4− cluster anion. The cluster shell is constructed by interconnection of two trimeric {V3O11} groups consisting of three edge-sharing VO5 polyhedra and four As2O5 units, which are formed by two corner-sharing AsO3 pyramids. While the As2O5 group is a common structural feature in arsenato-polyoxovanadates, the {V3O11} unit is only observed in V-rich high-nuclear heteroatom-containing polyoxovanadates {V14E8} (E = As, Sb, Ge). The complexes adopt the Λ (δδδ) conformation, which is the most stable arrangement. Interestingly, the unit cell parameters do not scale with the volume of the [M(en)3]2+ complexes, assuming a constant volume of the anion. Only a very detailed Hirshfeld surface analysis revealed that the van der Waals volume of the {V6As8O26} moiety is the smallest for the Cd-containing compound, while the volumes of the anions in the other two compounds are very similar. Therefore, the observed trends of the lattice parameters can be explained on the basis of these findings. Furthermore, intermolecular interactions include As⋯H contacts in addition to O⋯H and H⋯H interactions. The electronic spectrum of I contains d–d transitions of the vanadyl group and of the Co2+ cation. As expected only the d–d transitions of the VO2+ unit occur for II and III.

1. Introduction

The synthesis and characterization of polyoxometalates (POMs) with the elements V, Mo and W [1,2,3,4,5,6] is a still growing research field because several POMs show promising properties for applications in technologically important areas such as catalysis [7,8,9,10,11,12], in biomedicine [13,14,15], sorption of molecules [16] or in molecular electronics [17]. The most commonly applied synthesis methods are reactions at room or at moderate temperatures (reflux) in polar media such as water applying small molecular moieties which form larger cluster anions by self-assembly and condensation reactions [18]. However, many compounds were also synthesized using large, preformed POMs and/or applying hydrothermal conditions. It needs to be mentioned that not only artificial POMs are known, but a large number of POMs were also found in minerals [19], with several of them having no counterpart in synthetically produced compounds.
Among the main group element containing polyoxovanadates (POVs) the first cluster with integrated As (III) cations (As-POV), [V15As6O42(H2O)]6−, was already reported in 1988 [20]. The spherical cluster shell contains fifteen VO5 pyramids and three As2O5 handles. The cavity of the shell hosts a water molecule. The V centers are in the oxidation state +4 and the compound contains 15 unpaired electrons leading to highly interesting and unusual magnetic properties. Hence, this cluster was intensively investigated in the past and more than 60 papers were published [21]. The chemistry of As-POVs and other heteroatom-containing POVs was systematically developed during the last decades with the main aims to enrich the structure varieties and to alter the physical properties [20]. A large variety of As-POVs exhibiting different geometries and different V:As ratios were reported and selected examples are [V10AsO26(H2O)] [22], K7[V14AsO40] [23], K3[V12As3O42] [24], [{Ln(H2O)6}2V14As8O42(SO3)] (Ln = La3+, Sm3+, and Ce3+) [25], (NH4)4[V12As8O40(H2O)] [26], [Cd(dien)]2[Cd(dien)(H2O)]2[V16As4O42(H2O)], [27]. The crystal structure of the most As-rich As-POV, [V6As8O26]4−, features two trimeric units {V3O11} which are joined by four As2O5 handles formed by AsO3 pyramids sharing a common oxygen atom [28]. The compound was prepared under ambient conditions and no further compounds with this anion were reported until recently. This is a surprising observation because dozens of As-POVs were reported during the last decades. Interestingly, most of these compounds were synthesized under solvothermal conditions and all of them contain V-rich high nuclear clusters. One possible explanation that the As-rich cluster was not obtained in solvothermal syntheses may be a low stability under these special conditions. By systematic variation of the solvothermal reaction parameters we were able to synthesize [Ni(en)3]2[V6As8O26] (en = ethylenediamine) [29] which represents the second structure comprising the [V6IVAs8IIIO26]4− cluster. Solvothermal reactions are complex, and prediction of the reaction product is very limited. Nevertheless, to enhance the knowledge of As-rich POVs further explorative solvothermal syntheses were performed and we were able to prepare three compounds containing the rare [V6IVAs8IIIO26]4− cluster anion, namely [M(en)3]2[V6As8O26] (M = Co2+ (I), Zn2+ (II), and Cd2+ (III)). Here, we report on the syntheses, crystal structures and selected properties of these three compounds.

2. Materials and Methods

2.1. Reagents for the Syntheses of [M(en)3]2[V6As8O26] (M = Co, Zn, and Cd)

The chemicals were used as purchased: NH4VO3 (Merck PA, 99%); As2O3 (Merck PA, 99.5%); Zn(CH3COO)2·2H2O (Merck 99%); CoCl2·6H2O (VWR 98%; CdCl2·2.5H2O (Sigma Aldrich, Singapore, 99.995%).

2.2. Syntheses

All syntheses were performed in glass tubes (volume: 11 mL), heating the reaction mixtures at 150 °C for up to 4 d. Subsequently, the glass tubes were cooled down and the solids were filtered off. In all syntheses a green product was obtained containing needle-like crystals suitable for single crystal structure determination. The products were washed with water and ethanol and were stored in a desiccator. Upscaling experiments were performed in steel autoclaves with a 30 mL Teflon liner. The temperature of 150 °C was held for 4 d followed by cooling the autoclave to room temperature. The treatment of the reaction products was identical with that obtained in glass tubes.
For all syntheses in glass tubes, 1.341 mmol NH4VO3 (156.9 mg) and 1.194 mmol As2O3 (236.2 mg) were used. For I, 0.66 mmol (157 mg) CoCl2·6H2O; for II, 0.66 mmol (145 mg) Zn(CH3COO)2·2H2O; and for III, 0.66 mmol (150.8 mg) CdCl2·2.5H2O were mixed with the other solid educts. The liquid was composed of 1.7 mL ethylenediamine, and 2.3 mL H2O was added to the solid and the slurry was shaken. The syntheses in the autoclaves were performed using a quadrupled amount of solids and liquid. The yields based on NH4VO3 were 65% for I, 58% for II and 90% for III.

2.3. X-ray Powder Diffraction

The X-ray powder diffraction patterns were collected with a STOE Stadi-P diffractometer using a MYTHEN 1K detector (DECTRIS) and monochromatized Cu-Kα1 radiation (λ = 1.540598 Å). The experimental and the calculated patterns match perfectly indicating phase purity of the samples (Figure S1).

2.4. Spectroscopic Characterization and Chemical Composition

The infrared spectra were recorded at 295 K with a Bruker Vertex70 FT-IR spectrometer.
The UV/vis diffuse reflectance spectra were measured using the UV/vis two-channel spectrometer Cary 5 from Carian Techtron Pty., Darmstadt and BaSO4 served as white standard. The raw data were converted into the Kubelka–Munk factor.
Energy dispersive X-ray analysis (EDX) on selected single crystals of I–III was performed with an Environmental Scanning Electron Microscope ESEM XL30 from Philips. I: Co:V:As = 1:5.97:8.04; II: Zn:V:As = 1:6.05:7.94; III: Cd:V:As = 1:6.06:8.05.

2.5. Crystal Structure Determination

Single crystal data collections were performed with a STOE Imaging Plate Diffraction System (IPDS-2) using Cu-Kα radiation (λ = 1.54184 Å). Structure solution was performed with SHELXT [30] and structure refinements were performed against F2 using SHELXL-2016 [31]. All non-hydrogen atoms were refined with anisotropic displacement parameters. The C–H and N-H hydrogen atoms were positioned with idealized geometry and were refined isotropically with Uiso (H) = 1.2 Ueq (C) using a riding model. Selected crystal data and refinement results are listed in Table 1.
CCDC-2194249 (I), CCDC-2194248 (II) and CCDC-2194250 (III) contain the supplementary crystallographic data for this paper. These data can be obtained free of charge from the Cambridge Crystallographic Data Centre via https://www.ccdc.cam.ac.uk/data_request/cif (accessed on 15 September 2022).

3. Results and Discussion

3.1. Crystal Structure of I, II and III

The compounds [M(en)3]2[V6As8O26] (M = Co2+, Zn2+, and Cd2+) are isostructural and crystallize in the tetragonal space group I41/acd with Z = 8. The lattice parameters (Table 1) do not vary in a regular way and the shortest A axis and longest C axis are observed for M = Cd2+, and most probably the reason for this observation are the differing volumes of the cations and anions (see below). The two unique As atoms are located at general positions, while one independent V atom is on a special position and the second on a general position. Six of the seven unique oxygen atoms are also at general positions and the seventh as well as the M2+ cations are at special positions. All C, N, and H atoms of the [M(en)3]2+ complexes are at general positions. The M2+ cations are in an octahedral environment of six N donor atoms of three en ligands (Figure S2). The complexes adopt the Λ configuration and the C-C bonds are nearly parallel to the C3 axis of the molecules, i.e., the en ligands are in the δ conformation leading to Λ(δδδ) [32], which is the most stable conformation. The M2+-N bond lengths follow the ionic radii of the cations with the shortest for Co-N (average: 2.170 Å, Supplementary Materials, Table S1), slightly longer bonds for Zn-N (average: 2.190 Å, Table S1) and the longest bonds for Cd-N (average: 2.365 Å, Supplementary Materials, Table S1). The N-M2+-N angles clearly deviate from the ideal values of an octahedron and the three complexes exhibit different degrees of distortion (Supplementary Materials, Table S1). Using the minimum bounding ellipsoid (MBE) approach [33] the volumes of the polyhedra and the distortion expressed by the shape parameter S were calculated. The smallest volumes are obtained for the Co2+ and Zn2+ centered complexes (42.64 and 43.85 Å3) and the largest one for the Cd-containing moiety (55.06 Å3). The [Co(en)3]2+ and [Zn(en)3]2+ complexes are axially compressed, while [Cd(en)3]2+ is axially stretched. The MN2C2 chelate rings show the gauche conformation. The rings are puckered, and the degree of puckering can be expressed by the N-C-C-N torsion angles of 51.52 and 56.96° for I, 51.08 and 56.34° for II, and 45.00 and 52.60° for III, which all deviate from the expected value of approximately 60° and the deviation is most remarkable for the [Cd(en)3]2+ complex. Overall, the geometric parameters of the three tris(ethylenediamine) complexes match well with data published in the literature [34,35].
The V4+ cations are in a rectangular pyramidal environment (Figure 1), VO5, with the bonding pattern found in POVs: one short V = O2+ bond to the apical O atom (I: 1.600–1.605 Å; II: 1.602–1.606; III: 1.603–1.612 Å) and four longer bonds (I: 1.962–2.001 Å; II: 1.920–2.001 Å; III: 1.932–2.004 Å). The O-V-O angles (Supplementary Materials, Tables S2–S4) clearly show that the VO5 pyramids are distorted. Applying the MBE approach, the volumes of the polyhedra and the shape parameter S were calculated (Table 2). Only small differences can be seen demonstrating the rigidity of the VO5 polyhedra. In all three compounds, the V1 centered polyhedron is axially compressed and that containing V2 is slightly axially stretched.
The As atoms are in a trigonal–pyramidal environment with average As-O bond lengths of 1.769 (Co), 1.772 (Zn) and 1.773 Å (Cd) (Supplementary Materials, Tables S2–S4). Neighboring AsO3 pyramids share a common O atom to form four As2O5 handle-like moieties (Figure 1). The angles As-O-As are 127.3, 126.8, and 126.6° for M = Co, Zn, and Cd (Supplementary Materials, Tables S2–S4), and agree with those reported earlier [28,29]. Normally, this angle is larger in high-nuclear As-POVs and is in the range from 132 to 136°, indicating that the As-O-As angle is slightly strained.
The cluster [V6As8O26]4− is composed of two symmetry-related trimeric units, {V3O11}, comprising three VO5 pyramids sharing common edges (Figure 1). The two {V3O11} moieties are bridged by the As2O5 groups to form the cluster anion, which contains a cavity with dimensions 4.4 · 7 Å2 (measured from coordinate-to-coordinate). The size of the cavity is too small for hosting a guest molecule as is often observed for the larger As-POV cluster anions. All geometric parameters are in good agreement with those published for compounds with the [V6As8O26]4− cluster shell [28,29].
We note that the {V3O11} building unit is a common motif in high-nuclear heteroatom-containing POVs {V14E8} (E = As, Sb, Ge) [36,37,38,39,40,41,42,43,44]. In the crystal structures of these compounds eight edge-sharing VO5 pyramids form a central ring which is capped on both sides by the {V3O11} groups, which are rotated by approximately 90° against each other. Removal of the central ring leads to the structure of the cluster anion in the title compounds. Using the bond valence sum (BVS) method, the valence states of V and As were calculated yielding +4 for the V centers and +3 for the As atoms justifying the formula [V6IVAs8IIIO26]4−.
The arrangement of the cations and anions in the structures is shown in Figure 2. The anions form rods along the c-axis and the [M(en)3]2+ complexes are located in the spaces between the anions in a pairwise manner along [1 0 0] and [0 1 0] with M⋯M separations of ≈7.6 Å within the ‘pairs’ and 13.5 Å between the ‘pairs’ (distances are measured from coordinate-to-coordinate).

3.2. Hirshfeld Surface Analysis

Cations and anions are connected by N-H⋯O and C-H⋯O hydrogen bonding interactions (Supplementary Materials, Tables S5–S7) generating a three-dimensional network. For gaining more insight into the intermolecular interactions a Hirshfeld surface analysis was performed [45,46,47,48]. A Hirshfeld surface is constructed applying the equation
d n o r m = d i r i v d W r i v d W + d e r e v d W r e v d W
vdW = van der Waals radius; di represents the distance of a point on the surface to the nearest nucleus inside the surface; de is the distance from such a point to the next neighbor outside the surface. If an interatomic distance is shorter than the sum of the van der Waals radii, a red area occurs on the Hirshfeld surface, and all other colored regions indicate interatomic distances longer than the sum of the van der Waals radii. The analysis reveals that the differences between the three compounds are relatively small and the Hirshfeld surfaces of the anions and cations for the compounds are shown in Figure 3.
All the red areas on the Hirshfeld surfaces are caused by intermolecular O⋯H interactions to NH2/CH2 groups of the en ligands (see also Supplementary Materials, Tables S5–S7 for the geometric parameters of the hydrogen bonds). A further interpretation of the 3D Hirshfeld surface is not straightforward; therefore, a fingerprint plot can be constructed representing a 2D view of the 3D surface (Figure 4). Three different intermolecular interactions are identified for the cations: H⋯H (30.3 (Co, Zn) and 33.4% (Cd)), H⋯O (54.8 (Co, Zn) and 52.3% (Cd)) and H⋯As (15 (Co, Zn) and 14.4% (Cd)). Comparing the fingerprint plots of [M(en)3]2+ complexes (M = Co, Zn) with that of [Cd(en)3]2+ the sharp spike caused by H⋯O interactions appears at the same dedi position, but a second less sharp feature occurs for the Cd-containing complex at somewhat longer de-di values caused by H⋯H interactions, which are located at even longer dedi values in the Zn- resp. Co-containing complexes. For the anions the interactions are 74.4% for O⋯H and 26.6% for As⋯H, and they are virtually identical for the three compounds. For As⋯H interactions the shortest contact is at 2.77 (Co), 2.76 (Zn) and 2.85 Å (Cd), which are shorter than the sum of van der Waals radii of 2.95 Å [49]. The corresponding angles N-H⋯As are 153.4 (Co), 155 (Zn) and 160.3° (Cd) pointing towards significant interactions. A second As⋯H separation at 2.91 (Co), 2.89 (Zn) and 2.87 Å (Cd) with the C-H⋯As angles of 156.5 (Co), 154.6 (Zn) and 145.5° (Cd) indicates weaker intermolecular interactions.
The Hirshfeld surface calculations yield further quantities, such as the volumes and areas of the cations and anions. The volumes of the cations do not differ much: 290.87 for I, 289.28 for II and 290.73 Å3 for III. However, for the anions clearly different volumes were calculated: 649.05 for I, 644.97 for II and 628.89 Å3 for III. These differences may account for the non-regular changes of the lattice parameters across the series of compounds.

3.3. Spectroscopic Investigations

The IR spectra of the three compounds in the fingerprint region are displayed in Figure 5. In the NH2 and CH2 regions (not shown here) all spectra contain absorptions of the asymmetric and symmetric N-H stretching vibrations: 3330, 3309 and 3286 cm−1 for I, 3340, 3315 and 3292 cm−1 for II, and 3353, 3331 and 3288 cm−1 for III. The bands of the CH2 groups are at 2967, 2930, and 2885 cm−1 for I, at 2969, 2931 and 2887 cm−1 for II, and at 2969, 2924, and 2879 cm−1 for III. The occurrence of three bands in the N-H region is a typical sign for NH2 groups involved in hydrogen boding interactions. All these signals observed for the title compounds are in agreement with literature values of tris(ethylenediamine) complexes of the cations in I–III [50,51]. The NH2 bending vibration occurs at 1593 cm−1 for III, and is split into two signals for I and II (1600 and 1580 cm−1). The CH2 scissor vibration is located at around 1465 cm−1 and the CH2 twist occurs at ≈1330 cm−1 for I–III. The strong absorption at ≈960 cm−1 is caused by the V=O stretching vibration and is typical for the VO2+ vanadyl group. For the N-C-C-N skeleton six modes are expected which can be classified as three skeleton stretching vibrations, two N-C-C-N deformation modes and one torsional vibration around the C-C bond. The three stretching bands are between 900 and 1100 cm−1, but the overlap with other absorptions makes a distinct assignment impossible. The remaining absorptions of the skeleton are located at much lower wavenumber around 500 cm−1. Below 900 cm−1 the V-O-V, V-O-As, As-O-As, and As-O vibrations are located which cannot be unambiguously assigned. According to data in the literature, the M-N stretching modes are clearly located below 450 cm−1.
In the UV/vis spectra of the compounds (Figure 6), two absorptions are observed for II and III, which are located at ≈1.5 (12,098 cm−1, 827 nm) and ≈2.0 eV (16,131 cm−1, 620 nm), while for I an additional signal is present at ≈2.65 eV (21,374 cm−1, 468 nm) and the first transition is broader than for II and III. In addition, the absorption maximum of the first signal of I is slightly shifted to lower energy (1.44 eV, 11,614 cm−1, 861 nm). For compounds II and III, no d–d transitions are possible for the d10 electronic configuration; therefore, the absorptions must be caused by the vanadyl groups. For this group the d–d transitions 2B2g2Eg, 2B2g2B1g and 2B2g2A1g are expected [52]. Depending on the environment, i.e., packing in the structure, and on the actual distortion of the VO5 pyramids, the ideal C4v symmetry is reduced, and the degeneracy of the first band is removed. This effect may be the reason that the absorption at ≈1.5 is relatively broad and consists of two overlapping bands. For I the situation is different because for the Co2+ cation the first d–d transition 4T1g(F) → 4T2g(F) occurs in the energetic region of the first absorption band of the vanadyl unit and this a further reason for the broadness of the first absorption band in the spectrum of I. The signal at ≈2.65 eV can be assigned to the 4T1g(F) → 4A2g(F) transition of the Co2+ cation. The strong increase in the absorptions above 3 eV is most probable caused by charge-transfer processes.

4. Conclusions

In this manuscript, we have reported the solvothermal synthesis and selected properties of three new As-rich polyoxovanadates with composition [M(en)3]2[V6As8O26] (M = Co, Zn, Cd), a significant step forward in the chemistry of As-rich POV clusters. It is surprising that the number of compounds featuring this small As-POV cluster is comparatively low compared with other high-nuclearity As-POVs, especially keeping in mind that the primary building units such as the As2O5 handle and the trimeric {V3O11} group frequently occur in the structures of other As-POVs. From a synthetic point of view, the occurrence of three isostructural compounds prepared under solvothermal conditions is surprising because, in most cases, synthesis recipes cannot be directly transferred from, e.g., one transition metal cation to another. While the overall structural features of the three isostructural compounds are very similar, the Hirshfeld surface analysis reveals some subtle differences in the intermolecular interactions between I/II and III. The analysis identified the reason for the non-continuous variation of the lattice parameters, because one would on a first glance expect a steady change with the ionic radii of the cations. Surprisingly, the anion in III has the smallest van der Waals volume, allowing a denser packing of the constituents along [1 0 0].

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/cryst12101473/s1, Figure S1: Experimental and calculated X-ray powder patterns of I–III; Figure S2: ORTEP plots of the [M(en)3]2+ complexes in the structures of I, II, and III with labelling. Table S1. Selected bond lengths (Å) and angles (°) for the M2+ coordination in I, II, and III. Table S2. Selected bond lengths (Å) and angles (°) for the [V6As8O26]4− anion in I. Table S3. Selected bond lengths (Å) and angles (°) for the [V6As8O26]4− anion in II. Table S4. Selected bond lengths (Å) and angles (°) for the [V6As8O26]4− anion in III. Table S5. Hydrogen bonds [Å and °] for I. Table S6. Hydrogen bonds [Å and °] for II. Table S7. Hydrogen bonds [Å and °] for III.

Author Contributions

Conceptualization, methodology, funding and writing of draft: W.B.; syntheses and basic characterization: M.R.; single crystal structure determination: C.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are within the article or the supplementary information.

Acknowledgments

The authors thank the State of Schleswig-Holstein for financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Müller, A.; Kögerler, P. From simple building blocks to structures with increasing size and complexity. Coord. Chem. Rev. 1999, 182, 3–17. [Google Scholar] [CrossRef]
  2. Gouzerh, P.; Proust, A.A. Main-group element, organic, and organometallic derivatives of polyoxometalates. Chem. Rev. 1998, 98, 77–112. [Google Scholar] [CrossRef]
  3. Pope, M.T.; Müller, A. Polyoxometalate chemistry: An old field with new dimensions in several disciplines. Angew. Chem. Int. Ed. Engl. 1991, 30, 34–48. [Google Scholar] [CrossRef]
  4. Monakhov, K.Y.; Bensch, W.; Kögerler, P. Semimetal-functionalised polyoxovanadates. Chem. Soc. Rev. 2015, 44, 8443–8483. [Google Scholar] [CrossRef] [Green Version]
  5. Hill, C.L. Introduction:  Polyoxometalates - Multicomponent Molecular Vehicles To Probe Fundamental Issues and Practical Problem. Chem. Rev. 1998, 98, 1–2. [Google Scholar]
  6. Long, D.-L.; Burkholder, E.; Cronin, L. Polyoxometalate clusters, nanostructures and materials: From self-assembly to designer materials and devices. Chem. Soc. Rev. 2007, 36, 105–121. [Google Scholar] [CrossRef]
  7. Okuhara, T.; Mizuno, N.; Misono, M. Catalysis by heteropoly compounds—Recent developments. Appl. Catal. A Gen. 2001, 222, 63–77. [Google Scholar] [CrossRef]
  8. Kamata, K.; Yonehara, K.; Nakagawa, Y.; Uehara, K.; Mizuno, N. Efficient stereo and regioselective hydroxylation of alkanes catalysed by a bulky polyoxometalate. Nat. Chem. 2010, 2, 478–483. [Google Scholar] [CrossRef]
  9. Samaniyan, M.; Mirzaei, M.; Khajavian, R.; Eshtiagh-Hosseini, H.; Streb, C. Heterogeneous catalysis by polyoxometalates in metal—Organic frameworks. ACS Catal. 2019, 9, 10174–10191. [Google Scholar] [CrossRef]
  10. Kamata, K.; Sugahara, K. Base catalysis by mono and polyoxometalates. Catalysts 2017, 7, 345. [Google Scholar] [CrossRef] [Green Version]
  11. Lian, L.; Zhang, H.; An, S.; Chen, W.; Song, Y.-F. Polyoxometalates-based heterogeneous catalysts in acid catalysis. Sci. China Ser. B Chem. 2021, 64, 1117–1130. [Google Scholar] [CrossRef]
  12. Katsoulis, D.E. A survey of applications of polyoxometalates. Chem. Rev. 1998, 98, 359–388. [Google Scholar] [CrossRef]
  13. Bijelic, A.; Aureliano, M.; Rompel, A. Polyoxometalates as potential next-generation metallodrugs in the combat against cancer. Angew. Chem. Int. Ed. 2019, 58, 2980–2999. [Google Scholar] [CrossRef] [Green Version]
  14. Bijelic, A.; Aureliano, M.; Rompel, A. The antibacterial activity of polyoxometalates: Structures, antibiotic effects and future perspectives. Chem. Commun. 2018, 54, 1153–1169. [Google Scholar] [CrossRef] [Green Version]
  15. Qi, W.; Zhang, B.; Qi, Y.; Guo, S.; Tian, R.; Sun, J.; Zhao, M. The Anti-proliferation activity and mechanism of action of K12[V18O42(H2O)]∙6H2O on breast cancer cell lines. Molecules 2017, 22, 1535. [Google Scholar] [CrossRef] [Green Version]
  16. Uchida, S.; Mizuno, N. Design and syntheses of nano-structured ionic crystals with selective sorption properties. Coord. Chem. Rev. 2007, 251, 2537–2546. [Google Scholar] [CrossRef]
  17. Izarova, N.V.; Kögerler, P. Trends in Polyoxometalates Research; Ruhlmann, L., Schaming, D., Eds.; Nova Science Publishers Inc.: New York, NY, USA, 2015. [Google Scholar]
  18. Hutin, M.; Rosnes, M.H.; Long, D.-L.; Cronin, L. Comprehensive Inorganic Chemistry II; Reedijk, J., Poeppelmeier, K., Eds.; Elsevier: Oxford, UK, 2013; Volume 2, pp. 2241–2269. [Google Scholar]
  19. Krivovichev, S.V. Polyoxometalate clusters in minerals: Review and complexity analysis. Acta Crystallogr. Sect. B Struct. Sci. Cryst. Eng. Mater. 2020, 76, 618–629. [Google Scholar] [CrossRef]
  20. Müller, A.; Döring, J. A novel heterocluster with D3-symmetry containing twenty-one core atoms: [As6IIIV15IVO42(H2O)]6−. Angew. Chem. Int. Ed. Engl. 1988, 27, 1721. [Google Scholar] [CrossRef]
  21. Kögerler, P.; Tsukerblat, B.; Müller, A. Structure-related frustrated magnetism of nanosized polyoxometalates: Aesthetics and properties in harmony. Dalton Trans. 2009, 39, 21–36. [Google Scholar] [CrossRef]
  22. Bu, W.-M.; Ye, L.; Yang, G.-Y.; Shao, M.-C.; Fan, Y.-G.; Xu, J.-Q. A three-dimensional neutral framework of a novel decavanadium cluster bridged by an AsO4 tetrahedron: [AsVIV8VV2O26(μ-H2O)]·8H2O. Chem. Commun. 2000, 14, 1279–1280. [Google Scholar] [CrossRef]
  23. Müller, A.; Döring, J.; Khan, M.I.; Wittneben, V. [AsVV12VV2IVO40]7−, a topologically interesting, mixed valence cluster as model for vanadium minerals formed by weathering. Angew. Chem. Int. Ed. Engl. 1991, 30, 210–212. [Google Scholar] [CrossRef]
  24. Khan, M.I.; Chen, Q.; Zubieta, J. Hydrothermal syntheses and structures of reduced and mixed valence polyoxovanadium clusters with capped cage cores: K3[H12(AsO)2(AsO4)V6IVV6VO36]·12H2O and Rb5[(As2O)4(VO)2V12IVO36Cl]·2H2O. Inorganica Chim. Acta 1993, 212, 199–206. [Google Scholar] [CrossRef]
  25. Arumuganathan, T.; Das, S.K. Discrete polyoxovanadate cluster into an organic free metal-oxide-based material: Syntheses, crystal structures, and magnetic properties of a new series of lanthanide linked-POV compounds [{Ln(H2O)6}2As8V14O42(SO3)] x 8 H2O (Ln = La3+, Sm3+, and Ce3+). Inorg. Chem. 2009, 48, 496–507. [Google Scholar] [CrossRef]
  26. Wutkowski, A.; Evers, N.; Bensch, W. Synthesis, structure, and solubility studies of a new arsenato-polyoxovanadate cluster compound: (NH4)4[V12As8O40(H2O)]∙4H2O. Z. Anorg. Allg. Chem. 2011, 637, 2205–2210. [Google Scholar] [CrossRef]
  27. Zhou, J.; Zheng, S.-T.; Fang, W.-H.; Yang, G.-Y. A new 2-D network containing {As4V16O42(H2O)} cluster units. Eur. J. Inorg. Chem. 2009, 2009, 5075–5078. [Google Scholar] [CrossRef]
  28. Dumas, E.; Livage, C.; Sabine Halut, S.; Herve, G. Synthesis and molecular structure of a new arsenic rich polyoxovanadium anion, [As8IIIV6IVO26]4−. Chem. Commun. 1996, 21, 2437–2438. [Google Scholar] [CrossRef]
  29. Kondinski, A.; Rasmussen, M.; Mangelsen, S.; Pienack, N.; Simjanoski, V.; Näther, C.; Stares, D.L.; Schalley, C.A.; Bensch, W. Composition-driven archetype dynamics in polyoxovanadates. Chem. Sci. 2022, 13, 6397–6412. [Google Scholar] [CrossRef]
  30. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Cryst. A 2015, 71, 3–8. [Google Scholar]
  31. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. C 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  32. Beattie, J.K. Conformational analysis of tris(ethylenediamine) complexes. Accounts Chem. Res. 1971, 4, 253–259. [Google Scholar] [CrossRef]
  33. Cumby, J.; Attfield, J.P. Ellipsoidal analysis of coordination polyhedra. Nat. Commun. 2017, 8, 14235. [Google Scholar] [CrossRef] [Green Version]
  34. Pham, D.N.K.; Roy, M.; Golen, J.A.; Manke, D.R. The first-row transition-metal series of tris(ethylenediamine) diacetate complexes [M(en)3](OAc)2 (M is Mn, Fe, Co, Ni, Cu, and Zn). Acta Cryst. C 2017, 73, 442–446. [Google Scholar] [CrossRef]
  35. Ma, G.; Fischer, A.; Nieuwendaal, R.; Ramaswamy, K.; Hayes, S.E. Cd(II)–ethylenediamine mono- and bimetallic complexes—Synthesis and characterization by 113Cd NMR spectroscopy and single crystal X-ray diffraction. Inorganica Chim. Acta 2005, 358, 3165–3173. [Google Scholar] [CrossRef]
  36. Huan, G.; Greaney, M.A.; Jacobson, A.J. The synthesis and crystal structure of a new polyoxovandium cluster anion: [AsIII8VIV14O42(0.5H2O)]4−. J. Chem. Soc. Chem. Commun. 1991, 4, 260–261. [Google Scholar] [CrossRef]
  37. Guo, H.-Y.; Li, Z.-F.; Fu, L.-W.; Hu, Y.-Y.; Cui, X.-B.; Liu, B.-J.; Xu, J.-N.; Huo, Q.-S.; Xu, J.-Q. Structural influences of arsenic-vanadium clusters and transition metal complexes on final structures of arsenic–vanadium-based hybrids. Inorganica Chim. Acta 2016, 443, 118–125. [Google Scholar] [CrossRef]
  38. Wang, C.; Zhou, G.; Zhang, Z.; Zhu, D.; Xu, Y. Hydrothermal synthesis, characterization, and luminescence of two new ar-senic-vanadium compounds with a γ-[As8V14O42(H2O)]4− anions. J. Coord. Chem. 2011, 64, 1198–1206. [Google Scholar] [CrossRef]
  39. Wendt, M.; Näther, C.; van Leusen, J.; Kögerler, P.; Bensch, W. Mixing SbIII and GeIV occupancy in the polyoxovanadate {V14E8} archetype. Z. Nat. B 2018, 73, 773–779. [Google Scholar] [CrossRef]
  40. Wendt, M.; Näther, C.; Bensch, W. The high nuclearity antimonato-polyoxovanadate cluster {V15Sb6O42} as synthon for the solvothermal in-situ generation of α- and β-{V14Sb8O42} isomers. Chem. Eur. J. 2016, 22, 7747–7751. [Google Scholar] [CrossRef]
  41. Cui, X.-B.; Xu, J.-Q.; Li, Y.; Sun, Y.-H.; Ye, L.; Yang, G.-Y. Hydrothermal synthesis and characterization of (enH2)3.5[As8V14O42(PO4)]·2H2O. J. Mol. Struct. 2003, 657, 397–403. [Google Scholar] [CrossRef]
  42. Pitzschke, D.; Wang, J.; Hoffmann, R.-D.; Pöttgen, R.; Bensch, W. Two compounds containing the first mixed Germani-um-Vanadium-Polyoxothioanion [V14Ge8O42S8]12−. Angew. Chem. Int. Ed. 2006, 45, 1305–1308. [Google Scholar] [CrossRef]
  43. Qi, Y.; Li, Y.; Wang, E.; Jin, H.; Zhang, Z.; Wang, X.; Chang, S. Syntheses and structures of four organic–inorganic hybrids based on the surface restricted [AsIII8VIV14O42] building unit and various zinc coordination groups. Inorg. Chim. Acta 2007, 360, 1841–1853. [Google Scholar] [CrossRef]
  44. Qi, Y.; Li, Y.; Wang, E.; Zhang, Z.; Chang, S. Metal-controlled self-assembly of arsenic—Vanadium-cluster backbones with organic ligands. Dalton Trans. 2008, 17, 2335–2345. [Google Scholar] [CrossRef]
  45. Spackman, M.A. Molecules in crystals. Phys. Scr. 2013, 87, 48103. [Google Scholar] [CrossRef]
  46. Spackman, M.A.; McKinnon, J.J. Fingerprinting intermolecular interactions in molecular crystals. Cryst. Eng. Comm. 2002, 4, 378–392. [Google Scholar] [CrossRef]
  47. Spackman, M.A.; Jayatilaka, D. Hirshfeld surface analysis. Cryst. Eng. Comm. 2009, 11, 19–32. [Google Scholar]
  48. Spackman, P.R.; Turner, M.J.; McKinnon, J.J.; Wolff, S.K.; Grimwood, D.J.; Jayatilaka, D.; Spackman, M.A. Crystal explorer: A program for Hirshfeld surface analysis, visualization and quantitative analysis of molecular crystals. J. Appl. Crystallogr. 2021, 54, 1006–1011. [Google Scholar] [CrossRef]
  49. Mantina, M.; Chamberlin, A.C.; Valero, R.; Cramer, C.J.; Truhlar, D.G. Consistent van der Waals radii for the whole main group. J. Phys. Chem. A 2009, 113, 5806–5812. [Google Scholar] [CrossRef] [Green Version]
  50. Bennett, A.M.; Foulds, G.A.; Thornton, D.A. The i.r. spectra of ethylenediamine complexes—I. The tris(ethylenediamine) complexes of first transition series metal(II) sulphates. Spectrochim. Acta Part A Mol. Spectrosc. 1989, 45, 219–223. [Google Scholar] [CrossRef]
  51. Krishnan, K.; Plane, R.A. Raman and infrared spectra of complexes of ethylenediamine with Zinc(II), Cadmium(II), and Mercury(II). Inorg. Chem. 1966, 5, 852–857. [Google Scholar] [CrossRef]
  52. Ballhausen, C.J.; Gray, H.B. The electronic structure of the vanadyl ion. Inorg. Chem. 1962, 1, 111–122. [Google Scholar] [CrossRef]
Figure 1. Polyhedral representation of the cluster anion [V6As8O26]4− in the structures of the title compound.
Figure 1. Polyhedral representation of the cluster anion [V6As8O26]4− in the structures of the title compound.
Crystals 12 01473 g001
Figure 2. View of the arrangement of the constituents in the structure of the title compounds. C and H atoms are omitted for clarity.
Figure 2. View of the arrangement of the constituents in the structure of the title compounds. C and H atoms are omitted for clarity.
Crystals 12 01473 g002
Figure 3. The Hirshfeld surfaces of the cation (a) and anion (b) of [M(en)3]2[V6As8O16] (M = Co, Zn); the Hirshfeld surfaces of the cation (c) and anion (d) of [Cd(en)3]2[V6As8O26]. Red areas on the Hirshfeld surface indicate intermolecular distances shorter than the sum of the van der Waals radii.
Figure 3. The Hirshfeld surfaces of the cation (a) and anion (b) of [M(en)3]2[V6As8O16] (M = Co, Zn); the Hirshfeld surfaces of the cation (c) and anion (d) of [Cd(en)3]2[V6As8O26]. Red areas on the Hirshfeld surface indicate intermolecular distances shorter than the sum of the van der Waals radii.
Crystals 12 01473 g003aCrystals 12 01473 g003b
Figure 4. (a) the fingerprint plots of the cations in I, II, and III from left to right; (b) fingerprint plots of the anions in I, II, and III in the same order as the cations.
Figure 4. (a) the fingerprint plots of the cations in I, II, and III from left to right; (b) fingerprint plots of the anions in I, II, and III in the same order as the cations.
Crystals 12 01473 g004aCrystals 12 01473 g004b
Figure 5. The IR spectra of the title compounds in the fingerprint region.
Figure 5. The IR spectra of the title compounds in the fingerprint region.
Crystals 12 01473 g005
Figure 6. The UV/vis spectra of compounds I–III.
Figure 6. The UV/vis spectra of compounds I–III.
Crystals 12 01473 g006
Table 1. Selected crystal data and details of the single crystal structure refinements for I, II, and III.
Table 1. Selected crystal data and details of the single crystal structure refinements for I, II, and III.
CompoundIIIIII
formulaC12H48As8Co2N12O26V6C12H48As8N12O26V6Zn2C12H48As8Cd2N12O26V6
MW/g mol−11799.481812.361906
crystal systemtetragonaltetragonaltetragonal
space groupI41/acdI41/acdI41/acd
a/Å20.2849 (5)20.1813 (3)19.9293 (3)
b/Å20.2849 (5)20.1813 (3)19.9293 (3)
c24.4386 (8)24.5412 (4)24.8786 (5)
α909090
β909090
γ/°909090
V310,055.9 (6)9995.3 (3)9881.2 (4)
T/K200.0 (1)200.0 (1)200.0 (1)
Z888
Dcalc/g cm−32.3772.4092.563
µ/mm−17.0197.3587.333
θmax/deg26.00526.000527.005
measured refl.246843102037145
Rint0.06770.03280.0486
unique refl.248224552687
refl. F0 > 4σ (F0)220823662563
parameter151152152
R1 [F0 > 4σ (F0)]0.04640.02690.0486
wR2 [all data]0.10790.07270.1304
GOF1.1441.1601.159
Δρmax/min/e Å−30.422/−0.6300.373/−0.3610.665/−0.614
Table 2. Volumes of the VO5 polyhedra (Å3) and of the shape parameter S using the MBE method.
Table 2. Volumes of the VO5 polyhedra (Å3) and of the shape parameter S using the MBE method.
CoZnCd
V126.9826.9927.12
V226.8927.0327.04
S (V1)−0.12−0.11−0.12
S (V2)0.0.0430.0460.037
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rasmussen, M.; Näther, C.; Bensch, W. New As-Rich Arsenato-Polyoxovanadate Clusters: Solvothermal Synthesis and Selected Properties of [V6IVAs8IIIO26]4− Cluster-Containing Compounds. Crystals 2022, 12, 1473. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12101473

AMA Style

Rasmussen M, Näther C, Bensch W. New As-Rich Arsenato-Polyoxovanadate Clusters: Solvothermal Synthesis and Selected Properties of [V6IVAs8IIIO26]4− Cluster-Containing Compounds. Crystals. 2022; 12(10):1473. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12101473

Chicago/Turabian Style

Rasmussen, Maren, Christian Näther, and Wolfgang Bensch. 2022. "New As-Rich Arsenato-Polyoxovanadate Clusters: Solvothermal Synthesis and Selected Properties of [V6IVAs8IIIO26]4− Cluster-Containing Compounds" Crystals 12, no. 10: 1473. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12101473

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop